Next Article in Journal
Bidirectional TRP/L Type Ca2+ Channel/RyR/BKCa Molecular and Functional Signaloplex in Vascular Smooth Muscles
Next Article in Special Issue
Fibroblasts—Warriors at the Intersection of Wound Healing and Disrepair
Previous Article in Journal
Strain-Dependent Morphology of Reactive Astrocytes in Human- and Animal-Vole-Adapted Prions
Previous Article in Special Issue
Trametinib-Induced Epidermal Thinning Accelerates a Mouse Model of Junctional Epidermolysis Bullosa
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Extracellular Targets to Reduce Excessive Scarring in Response to Tissue Injury

1
Department of Orthopaedic Surgery, Sidney Kimmel Medical College, Thomas Jefferson University, Philadelphia, PA 19107, USA
2
Rothman Institute of Orthopaedics, Thomas Jefferson University Hospital, Philadelphia, PA 19107, USA
*
Author to whom correspondence should be addressed.
Biomolecules 2023, 13(5), 758; https://doi.org/10.3390/biom13050758
Submission received: 29 March 2023 / Revised: 24 April 2023 / Accepted: 25 April 2023 / Published: 27 April 2023
(This article belongs to the Special Issue Role of Mesenchymal Cells in Wound Healing and Fibrosis)

Abstract

:
Excessive scar formation is a hallmark of localized and systemic fibrotic disorders. Despite extensive studies to define valid anti-fibrotic targets and develop effective therapeutics, progressive fibrosis remains a significant medical problem. Regardless of the injury type or location of wounded tissue, excessive production and accumulation of collagen-rich extracellular matrix is the common denominator of all fibrotic disorders. A long-standing dogma was that anti-fibrotic approaches should focus on overall intracellular processes that drive fibrotic scarring. Because of the poor outcomes of these approaches, scientific efforts now focus on regulating the extracellular components of fibrotic tissues. Crucial extracellular players include cellular receptors of matrix components, macromolecules that form the matrix architecture, auxiliary proteins that facilitate the formation of stiff scar tissue, matricellular proteins, and extracellular vesicles that modulate matrix homeostasis. This review summarizes studies targeting the extracellular aspects of fibrotic tissue synthesis, presents the rationale for these studies, and discusses the progress and limitations of current extracellular approaches to limit fibrotic healing.

1. Introduction

The ability to heal injured tissues is fundamental for survival. Natural healing usually includes scar formation, a process accelerated by inflammation. While scars patch the injury sites, excessive scarring alters critical tissue and organ functions.
Regardless of the injury site, tissue type, and nature of the injury, the healing process includes hemostasis, inflammation, proliferation, and remodeling. For instance, in response to acute or systemic injury, blood-derived and local inflammatory cells migrate to damaged sites and set the stage for tissue repair by producing many growth factors. Subsequently, these growth factors stimulate resident and migratory fibroblastic cells, increasing their proliferation and the biosynthesis of scar tissue materials. Myofibroblasts are crucial producers of scar-building elements, including fibrillar collagens, fibronectin, proteoglycans, glycosaminoglycans, and others (Figure 1) [1,2].
While complete regeneration of injured adult tissues (i.e., returning to their original state) is a rare phenomenon, in some fetal tissues, regeneration may occur [3]. Essential elements for the regeneration of mature tissues include an intact extracellular matrix (ECM) and tissue-specific cells able to synthesize damaged components. For example, hepatocytes can regenerate the liver following acute toxic injury if the ECM is undamaged. Studies have demonstrated that the hepatocytes that perform regeneration are derived from local or circulating stem cells or mature hepatocytes that re-entered the cell cycle [4]. In contrast, chronic or traumatic acute injuries that damage the ECM architecture make healing by regeneration impossible. Consequently, healing by scarring helps to maintain the function and structural integrity of wounded tissues.
Although balanced scar formation maintains tissue integrity, excessive scarring in various tissues and organs is a significant medical problem. According to some estimates, fibrotic disorders are associated with 45% of all deaths [5]. Despite the enormous burden caused by these disorders, attempts to treat them have been largely unsuccessful. Consequently, researchers continue to define anti-fibrotic targets and design relevant inhibitors to block or reverse the fibrotic scarring process. This review focuses on one of these targets, namely the formation of collagen-rich ECM that defines the fibrotic deposits.

2. Collagen-Rich Matrix, a Versatile Biological Patch

Although healthy tissues differ in the cell types, molecular composition, and architecture of their ECM, the scar formation steps and fundamental elements of fibrotic deposits are similar.
For example, healthy cartilage mainly consists of collagen II-rich fibrils, pericellular collagen VI, fibril-associated collagen IX, and proteoglycans that form a structure able to withstand compression forces. Cartilage-specific cells, chondrocytes, control cartilage homeostasis and maintain cartilaginous structure [6].
The cellular makeup of the kidney, however, is more complex. It includes smooth muscle cells, podocytes, pericytes, fibroblasts, and many other cell types. Similarly, the pool of collagenous proteins differs and includes collagen I, collagen III, relatively large amounts of collagen IV, collagen VI, and a few additional collagen types [7].
Despite the differences in the cellular and ECM ingredients of healthy cartilage and kidney, both heal by scar formation. In both tissues, the main component of the scar is collagen I-rich fibrils [8,9]. Because similar mechanisms function in other tissues and organs, the formation of collagen I-rich scars represents a ubiquitous tissue repair mechanism, regardless of the wounded tissue’s injury type or location.
Although scar-based repair appears to be nature’s way to fix injuries without needing multiple unique and tissue-specific repair mechanisms, the tradeoff is that, when formed excessively, significant scarring can severely alter the repaired tissue architecture and function.

3. Excessive Scar Formation That Alters Vital Functions of Affected Tissues and Organs

Excessive scarring of the skin, tendon, muscle, and ligament alters the mechanical functions of these tissues. Similarly, ocular scars may impair vision, while the excessive formation of fibrotic deposits can alter vocal cords and harm speech. Post-traumatic scarring of peripheral nerves prevents the regeneration of the axons and blocks their growth from the proximal toward the distal stump of an injured nerve [10,11,12,13,14,15].
Similar consequences of excessive formation of collagen I-rich deposits have also been observed in organ fibrosis disorders caused by chronic inflammation, often with no defined injury events. For instance, patients with idiopathic pulmonary fibrosis (IPF) develop collagen I-rich thick scar tissue that blocks an efficient oxygen exchange, ultimately leading to death. In addition, as indicated above, fibrotic deposits in the liver and kidney alter these organs’ proper functions and may lead to their failure [16,17]. Further, patients with scleroderma develop stiff skin and, in later stages of the disease, progress toward fibrosis of multiple sites, including joints, lungs, esophagus, heart, and other organs [18].
These examples illustrate that repairing acute local and systemic injuries is a complex balancing act that can rapidly shift from needed tissue repair to unwanted fibrotic scarring, with severe outcomes.

4. Anti-Fibrotic Treatment: A Challenging Task

Even with decades-long studies attempting to define anti-fibrotic treatments, validate anti-fibrotic targets, and produce valuable therapeutics, effective and safe therapies designed to limit excessive scarring have yet to be developed.
Nintedanib and pirfenidone were only recently approved by the Food and Drug Administration (FDA) for patients with IPF. Clinical data indicate that these drugs slow down the rate of decline in forced vital capacity (FVC). However, studies did not show conclusively if these drugs reduce the mortality of patients with IPF [19].
Furthermore, tests of nintedanib applied to reduce pleuroparenchymal fibroelastosis, a subtype of interstitial pneumonia with upper lobe fibrosis, showed limited efficacy of this drug compared with that for the IPF treatment [20]. These studies suggest only a limited utility of nintedanib for treating fibrotic disorders of the lungs.
The anti-fibrotic mechanism of these FDA-approved drugs remains unclear. Some have suggested that they have broad anti-inflammatory effects and reduce the production of pro-fibrotic factors and matrix elements, including transforming growth factor-beta 1 (TGF-β1), tumor necrosis factor-alpha (TNF-α), platelet-derived growth factor (PDGF), interleukin 1 beta (IL-1β), and collagen I [21]. Other studies indicate that the anti-fibrotic mechanisms of these drugs may directly block collagen fibrillogenesis [22].
Nintedanib and pirfenidone were also tested as inhibitors in many other fibrotic conditions in models of excessive skin, eye, and muscle scarring [23,24,25,26]. Although the drugs demonstrated anti-fibrotic properties in some of these tests, they have yet to be applied clinically to treat fibrotic disorders other than IPF.

4.1. Targeting Pro-Fibrotic Cells

Myofibroblasts that elaborate fibrosis, intracellular processes that drive excessive scarring, and the extracellular steps of the scar matrix assembly are recognized as anti-fibrotic targets. Strategies to limit the pro-fibrotic behavior of the myofibroblasts include using anti-proliferative agents, blocking the transition of fibroblastic and epithelial cells to myofibroblasts, and inhibiting circulating pro-fibrotic cells from homing in on injury sites [27].
Since many of these processes are controlled by TGF-β1, this cytokine and associated mediators of its activity, including connective tissue growth factor (CTGF), represent crucial anti-fibrotic targets.
Despite the crucial roles of cells, pro-fibrotic intracellular processes, and growth factors associated with excessive scarring, no adequate specific treatments that aim at these targets have been developed for clinical use. Although the reasons for the poor outcomes are unclear, the literature points to several problems hampering the development of successful anti-fibrotic approaches. One problem is the natural redundancy of injury repair mechanisms that utilize multiple pathways to form collagen I fibril-rich scars. In fibrosis, these mechanisms are preserved and active in all tissues. Therefore, targeting only one mechanism or pathway to block the fibrotic process is likely insufficient [28].

4.2. Stiff ECM: A Crucial Pro-Fibrotic Culprit

The common denominator of different scarring mechanisms is the end product of the scarring process, namely, collagen I-based fibrotic neotissue. Following initial synthesis, this tissue stiffens, altering crucial natural functions of repaired sites.
Studies have demonstrated that stiff tissue is a crucial pro-fibrotic stimulant of fibroblasts and inflammatory cells in injury sites (Figure 2) [28,29,30].

4.2.1. Fibroblasts

Local fibroblasts that reside in wounded sites and fibroblasts that migrate from distant locations perform crucial tasks in balanced wound healing and fibrotic scarring (Figure 2) [31]. Many growth factors modulate these tasks, with TGF-β1 playing the central role [31,32].
As indicated above, contractile myofibroblasts that express α smooth muscle actin (αSMA) incorporated into stress fibrils are a hallmark of fibrosis [31]. Crucial functions of these cells include wound contracture and the production of elements of the ECM, including collagenous proteins.
Myofibroblasts interact with inflammatory cells, including macrophages and mast cells. These cells influence fibroblast activities by secreting TGF-β1, PDGF, vascular endothelial growth factor (VEGF), IL-6, and IL-13. In turn, fibroblasts impact the macrophages’ phenotype and function by changing the physical properties of the ECM [31].
In wound healing without excessive fibrosis, myofibroblasts ultimately cease their functions. They may revert to an “inactive fibroblast” state, enter senescence, or be eliminated via apoptosis at the end of a routine healing process. In contrast, in excessive scarring, myofibroblasts remain active in the fibrotic processes, resisting apoptosis while continuing their pro-fibrotic activities [33].

4.2.2. Inflammatory Cells

Inflammatory cells that drive fibrotic healing include mast cells and macrophages (Figure 2) [31]. Studies have demonstrated that the stiff ECM environment enhances the pro-fibrotic behavior of mast cells and promotes their durotaxis, i.e., migration along stiffness gradients (Figure 2) [34]. Consequently, Hildebrand et al. targeted these cells with ketotifen to reduce the progress of fibrotic healing after an elbow injury. However, clinical trials in Canada demonstrated that this approach did not significantly mitigate post-traumatic elbow stiffness [35].
These studies indicate that aiming at mast cells alone is insufficient to reduce fibrotic healing in a clinically relevant way.
Macrophages show stiffness-dependent pro-fibrotic behavior too [36]. As previously demonstrated, mechano-gated ion channels and the α2β1 integrin, a member of the integrin family of heterodimeric cellular receptors, mediate the mechanical activation of these cells via complex signaling pathways that promote cell migration, proliferation, and ECM production [31,37].

5. Targeting the ECM Stiffness

Considering the crucial pro-fibrotic properties of scar neotissue, some have suggested that targeting the stiff ECM formation and modulating its pro-fibrotic signals may be a game-changer for developing anti-fibrotic therapies [28].
Here, we focus on crucial processes and factors that increase ECM stiffness and contribute to pro-fibrotic mechanotransduction. They include (i) extracellular regulators of scar production, e.g., TGF-β1; (ii) extracellular ECM assembly, e.g., collagen fibrillogenesis; and (iii) cellular receptors that allow the ECM–cell communication, specifically integrins [28,29,30,38].

5.1. Production of a Crucial Precursor of Fibrotic Deposits

Collagen I is the main component of scars formed in the skin, musculoskeletal systems, peripheral nerves, the eye, abdomen, spinal cord, and others [39,40,41,42,43,44]. This protein constitutes the most considerable portion of fibrotic tissues formed due to injury in the internal organs, including the liver, lungs, heart, and kidney [45].
The fibrillar architecture formed by this collagen type provides mechanical stability to the injury sites. Nevertheless, collagen I-rich deposits are the main factor causing harmful consequences associated with excessive scarring. These fibrillar structures are produced in a complex process that includes intracellular and extracellular steps (Figure 3).

5.1.1. Intracellular Steps of Collagen I Synthesis

Each collagen molecule comprises three collagen α-chains, that associate in the endoplasmic reticulum (ER) into a triple-helical structure. In the fibril-forming collagens, including collagen I, each chain consists of approximately 330 uninterrupted repeats of -G-X-Y- triplets, in which the -X- and -Y- positions are frequently occupied by proline residues [46].
The fibril-forming collagens are produced as procollagens, in which the triple-helical domains are flanked by globular N-terminal and C-terminal propeptides (Figure 3). Relatively short non-triple-helical telopeptides separate the propeptides and the triple-helical domain.
Post-translational modifications of nascent procollagen α-chains are vital steps that determine collagen molecules’ proper thermostability and mechanical properties. In particular, proline and lysine residues present in the -Y- positions of the -G-X-Y- triplets are hydroxylated by prolyl 4-hydroxylase (P4H) and lysyl hydroxylase (LH), respectively [46].
P4H is a tetramer formed by two catalytic α subunits (P4Hα) and two non-catalytic β subunits (P4Hβ). P4Hβ also serves as protein disulfide isomerase (PDI) and a protein chaperone that prevents premature aggregation of procollagen chains [47,48]. 3-Hydroxyproline residues are also present in the -X- and -Y- positions of the –G-X-Y- triplets [49]. In procollagen I, only one proline residue of the α1(I) chain is 3-hydroxylated [50].
Mature procollagen chains assemble into a triple-helical conformation by a zipper-like folding mechanism [51]. Specialized chaperone proteins stabilize the procollagen molecules and prevent their aggregation. Chaperones involved in procollagen biosynthesis include (i) heat-shock protein 47 (HSP47), (ii) heat-shock 70 kDa-related luminal binding protein (BiP), and (iii) P4Hβ/PDI [52].

5.1.2. Extracellular Procollagen I Processing

Following secretion to the extracellular space, enzymatic cleavage of procollagen propeptides triggers collagen fibril formation [53]. A group of proteolytic enzymes, including a disintegrin and metalloprotease with thrombospondin motifs (ADAMTS)-2, -3, and -14, cleaves the N-terminal propeptides [54]. Among these enzymes, ADAMTS-2, or procollagen N proteinase (PNP), is abundant in collagen I-rich tissues, including skin, tendon, bone, eye, and others [55].
Another group of enzymes, from the tolloid family of zinc metalloproteinases, cleaves the C-terminal propeptides of fibrillar procollagens. Among these metalloproteases, procollagen C proteinase (PCP), also known as bone morphogenetic protein-1 (BMP-1), plays a pivotal role [56].
Further, PCP enhancer (PCPE) participates in the cleavage of procollagen I C propeptides by increasing the rate of the propeptide cleavage up to 20-fold [57]. Studies have shown that this protein is upregulated in many fibrotic conditions, including hypertrophic, keloid, and ocular scars. The increased production of PCPE was also demonstrated in organ fibrosis models [58,59].
Researchers have demonstrated that other enzymes might also process procollagen propeptides. These enzymes include meprins and mast cell chymase, whose activity increases during inflammation and fibrosis [60,61].

5.1.3. Extracellular Assembly of Collagen Fibrils

Following the cleavage of the procollagen propeptides, collagen I molecules self-assemble to form fibrils in a process driven by site-specific interactions among individual collagen molecules [62]. The binding interaction between the C-terminal telopeptides of one collagen molecule and an interacting partner’s telopeptide-binding region (TBR) facilitates crucial nucleation and a proper staggered alignment of collagen molecules [63,64,65]. This interaction is ground zero for the collagen fibril growth in physiological conditions and during the formation of scar deposits (Figure 3).

5.1.4. Cross-Linking of Collagen Fibrils and its Impact on ECM Stiffness

The assembly of individual collagen I molecules into fibrils is an entropy-driven process [53]. The nascent fibrils are held together by electrostatic and hydrophobic forces. However, these fibrils are unstable and may dissociate back into collagen molecules by changing optimal temperature or solvent conditions [66,67].
Thus, the assembled fibrils must be stabilized by covalent bonds between the individual collagen molecules that build them. Ultimately, these bonds, or cross-links, define the resistance of collagen fibrils to proteolytic degradation and determine their mechanical strength (Figure 3).
The hydroxylation of selected lysine residues, catalyzed by the lysyl oxidase (LOX) family of enzymes, facilitates the formation of collagen cross-links [48]. Transglutaminase 2 (TG2) also catalyzes the formation of fibril-stabilizing cross-links [68,69,70].
Collagen fibril formation is a prerequisite for the formation of the cross-links. The collagen molecules must first be arranged in the staggered, D-periodic pattern to allow the LOX enzymes to catalyze the cross-linking reaction.

5.1.5. Collagen Fibrillogenesis: A Crucial Anti-Fibrotic Target

Ultimately, the number of collagen fibrils, their spatial organization, and the extent of their cross-linking define the stiffness of the scar tissue and impact the severity of fibrotic disorders. Therefore, reducing the number of fibrils is the goal of all anti-fibrotic approaches, regardless of whether they aim at cells, intracellular, or extracellular targets.

6. Mediators of the Stiffness-Dependent Signals

The stiff-ECM-derived mechanical signals upregulate the expression of macromolecules that build the fibrotic scars. The mechanotransduction that modulates this expression involves integrins. These receptors facilitate cell signaling by forming focal adhesion structures that include talin, vinculin, focal adhesion kinase (FAK), and paxillin (Figure 4).
DDRs also participate in pro-fibrotic mechanotransduction [71,72]. Research data indicate that DDR1 regulates cell adhesion and migration through collagen-rich matrices by associating with non-muscle myosin IIA [73]. This protein is a hexameric enzyme with ATPase activity. It can bind to the actin cytoskeleton, generate forces that shape cell architecture, and facilitate cell motility (Figure 4) [74,75].
Key mediators of fibrotic processes are TGF-β receptors (TGFBR), which regulate the ECM–cell signaling via binding its TGF-β ligands and modulate SMAD-mediated processes (Figure 4). As reviewed by Abuammah et al., under low-shear mechanical conditions, the TGF-β1-dependent SMAD-2 signaling is upregulated [76]. The authors suggested that this mechanism may increase epithelial-to-mesenchymal transition, increasing fibrotic responses in some tissues [77].
Additional mechanotransduction pathways depend on regulating mechanosensitive ion channels via the tether force mechanism facilitated by the extracellular and intracellular partners [38,78].
The nucleus also participates in mechanosensing [79]. One of the elements facilitating the mechanosensing functions of the nucleus is lamin-A, whose production correlates positively with collagen-dependent ECM stiffness [80]. Other central players include yes-associated protein (YAP) and transcriptional co-activator with PDZ-binding motif (TAZ). Under low mechanical loading, they are located in the cytoplasm, where proteasomes degrade them. In contrast, YAP and TAZ escape proteasomal degradation in a high-mechanical-loading environment and translocate to the nucleus. There, they upregulate fibroblasts’ activities, including proliferation, differentiation, suppression of apoptosis, and matrix production [81,82,83,84]. Studies have also demonstrated that in addition to YAP/TAZ signaling, myocardin-related transcription factor (MRTF) participates in ECM stiffness-dependent fibroblast activities (Figure 4) [85,86].

7. Mechanotherapeutics

Because of the central role of the factors that mediate the stiffness-dependent pro-fibrotic cell behavior, targeting them is an attractive approach to reduce fibrotic scarring (Figure 4). Efforts to block pro-fibrotic cells and some canonical intracellular mediators of fibrotic healing, however, have been unsuccessful [29]. Significant concerns that hamper the bench-to-bed transition of these efforts include poor specificity of the blockers and unwanted side effects [28].
Targets for the mechanotherapeutics tested thus far are mainly downstream of the central pro-fibrotic physical stimulant, namely the stiff ECM (Figure 3). Aiming at less explored upstream targets directly associated with steps stiffening the matrix, e.g., collagen fibrillogenesis, offers an attractive yet poorly exploited alternative for reducing excessive fibrosis (Figure 3).

7.1. Targeting Procollagen Processing

As indicated above, one of the necessary conditions for forming proper collagen fibrils is extracellular, enzymatic removal of procollagen N-terminal and C-terminal propeptides (Figure 3). Studies have demonstrated that the presence of both propeptides precludes the formation of functional stable fibrils. Even the presence of only one of the propeptides leads to the assembly of abnormal tape-like or sheet-like structures [87]. In dermatosporaxis, where the N propeptides have not been removed from the procollagen I molecules, similar structures weaken the architecture of collagen I-rich tissues, most notably the skin [88].
Because of the importance of procollagen I propeptide cleavage for fibril formation, PNP and PCP were identified as attractive anti-fibrotic targets. The rationale for targeting these enzymes was that inhibiting their activities would prevent the cleavage of procollagen propeptides, thereby preventing collagen fibril formation and limiting excessive scarring (Figure 3).
In one study, Ovens et al. synthesized acidic dipeptide hydroxamate inhibitors of PCP. They demonstrated their utility in inhibiting PCP in vitro [89]. Although inhibitors of PNP were reported, they showed broad inhibitory properties, limiting their potential use as specific inhibitors of the N propeptide cleavage [90].
Mouse-based experiments, in which scientists knocked out genes encoding PCP or PNP, demonstrated that other enzymes may also process the procollagen propeptides. Because of these alternative procollagen I propeptide cleavage mechanisms, the interest in targeting PCP and PNP has diminished [91,92].
Still, scientists are exploring modulating the PCP activity by inhibiting procollagen C proteinase enhancer-1 (PCPE-1, Figure 3). Research has demonstrated that this protein is pivotal in mediating PCP activity in vivo. Further, the biosynthesis of this enhancer is upregulated in fibrotic conditions, including in the skin, heart, liver, kidney, lungs, ligament, muscle, eye, and other tissues and organs [93]. So far, however, no specific PCPE-1 inhibitors have been developed. Controversies exist regarding the safety of inhibiting the activity of this protein, in particular in the context of its roles in broad biological processes, including angiogenesis, cell proliferation, and RNA stabilization [93].

7.2. Blocking Collagen Self-Assembly into the Fibrils

One of the newer concepts for reducing the extracellular buildup of collagen I-rich fibrotic deposits is blocking the assembly of collagen molecules into fibrils [64]. Scientists demonstrated that a rationally engineered monoclonal antibody that targets the C-terminal telopeptide of collagen I prevents the aggregation of blocked collagen molecules into fibrils (Figure 3) [94,95]. The anti-fibrotic activities of this anti-collagen antibody (ACA) have been demonstrated in animal models of arthrofibrosis, skin fibrosis, and lung fibrosis [64,96,97,98].
Recently, Steplewski et al. demonstrated that blocking collagen I fibril formation with the antibody accelerates the degradation of ACA-blocked collagen molecules not incorporated into fibrils and speeds up the remodeling of injury sites. Furthermore, detailed analyses of tissues from animals treated with the ACA continuously for two months found that this antibody was safe and caused no side effects [98].
Therefore, the ACA-based method to limit fibrosis offers an attractive anti-fibrotic approach that targets the early onset of the extracellular process of fibril formation. Since collagen fibrillogenesis is a prerequisite for scar formation in all tissues and organs, targeting this process may be a versatile therapeutic approach to limit excessive scarring.
Other inhibitors of collagen fibrillogenesis have also been tested in vitro. Studies demonstrated that (±)-α lipoic acid, trigonelline hydrochloride, oleuropein, capsaicin, soluble discoidin domain receptors (DDR), and fibromodulin block collagen fibrillogenesis by directly interacting with collagen molecules [99,100,101,102,103,104].
Unlike with the ACA, however, the anti-fibrotic utility of these molecules has not been studied in relevant animal models of fibrotic scarring. However, the mechanism of blocking collagen fibrillogenesis with these molecules warrants future tests, to establish their anti-fibrotic potential and safety.

7.3. Reducing the Collagen Cross-Linking

Tissue repair via scarring involves scar tissue remodeling. Various enzymes, most notably matrix metalloproteinases (MMPs), degrade the scar elements during this process. Simultaneously, biosynthesis of new ECM elements rebuilds tissue architecture. In optimal conditions, this process transforms a stiff scar into a structure more compatible with the surrounding native tissue.
One of the hurdles that alters the degradation stage of remodeling is the resistance of collagen-rich mature matrices to proteolytic degradation. This resistance is mainly caused by the intrinsic stability of collagen molecules, their tight packing in the fibrils, and covalent cross-links that make fibrillar deposits a formidable target for proteolytic degradation (Figure 3).
In addition to negatively impacting the remodeling of scar tissue, cross-linking stiffens the scar’s ECM, further enhancing its pro-fibrotic characteristics.
Studies have found that those characteristics create mechanical, pro-fibrotic signals not only to fibroblasts, that continue to produce the ECM components, but also to immune cells that participate in healing. Research performed in animal models indicates that a stiff matrix promotes a pro-fibrotic inflammatory response in macrophages and prompts them to produce collagen and other macromolecules that contribute to scar formation [105,106]. Studies have also suggested that a key player that enables the mechanosensing properties of macrophages is an ion channel, PIEZO1 [107].
Because of its involvement in creating a pro-fibrotic matrix environment, collagen cross-linking represents an attractive target to limit excessive scar formation and thus soften the neotissue formed in response to injury.

7.3.1. Inhibiting LOX Activity

One of the indications that blocking LOX-mediated cross-linking reduces fibrotic scarring is an experiment with β-aminopropionitrile (BAPN), that chelates the copper ions needed for proper LOX activity. In animal models, treatment with this compound reduced bleomycin-induced pulmonary fibrosis, CCl4-induced liver fibrosis, and esophageal scarring caused by alkali burn [108,109,110].
To create a clinically relevant LOX inhibitor, researchers identified lysyl oxidase-like 2 (LOXL2), a member of the LOX family, as a critical player in the progression of many fibrotic disorders, including pulmonary, cardiac, and tumor-associated fibrosis [111].
Given its pro-fibrotic role, LOXL2 has become a valid anti-fibrotic target (Figure 3). Barry-Hamilton et al. demonstrated that a LOXL2-specific monoclonal antibody inhibits fibrotic changes in cancer, liver, and pulmonary fibrosis models [112]. Subsequently, a humanized IgG4 variant of the anti-LOXL2 antibody (Simtuzumab, Gilead Sciences, Inc., Foster City, CA, USA) was engineered. Its utility to block pulmonary, liver, and tumor-associated fibrosis was tested in clinical trials. These trials, however, were terminated due to the lack of efficacy of simtuzumab [113,114,115,116].
Despite positive outcomes in rodent-based models, it remains unclear why blocking LOXL2 in humans demonstrates no appreciable beneficial anti-fibrotic effects. Puente et al. suggested this could be due to the uninterrupted activity of other collagen cross-linking enzymes, including other members of the LOX family and tissue transglutaminase [117].

7.3.2. Targeting Transglutaminase 2 (TG2)

TG2 is a multifunctional enzyme, able to catalyze the formation of cross-links between the ε-amino group of a lysine residue and a γ-carboxamide group of glutamine residue. TG2-mediated cross-linking of the collagen-rich matrix plays a significant role in fibrosis progression, and blocking this enzyme reduces fibrosis-associated collagen deposition (Figure 3). Oh et al. demonstrated that in TG2-knockout mice, bleomycin-induced pulmonary fibrosis was significantly reduced due to the attenuation of collagen deposition [118]. Similarly, an irreversible TG2 inhibitor reduced fibrosis in a rat model of kidney fibrosis and mouse models of nephrosclerosis, myocardial infarction, and peritoneal fibrosis [119,120,121,122].
Studies have demonstrated that, in addition to stiffening the ECM and making it resistant to proteolysis, TG2 enhances TGF-β1 pro-fibrotic functions. Troilo et al. showed that this enhancement depends on the TG2-mediated multimerization of latent TGF-β binding protein 1 (LTBP1) [123]. The authors suggested that LTBP1 oligomers enhance TGF-β1 binding and activation of this factor by mechanical forces.
Efforts to develop TG2 inhibitors identified a group of molecules that act as competitive amine, reversible allosteric, or irreversible inhibitors [124,125]. These inhibitors, however, lack TG2-blocking specificity, and their use may cause unwanted side effects. Harrison et al. demonstrated that TG inhibitors caused hyperproliferation and parakeratosis in a skin model [124]. Further, Freund et al. demonstrated that targeting TG may inhibit crucial coagulation factor XIIIa [126].
In search of a specific TG2 inhibitor, scientists have developed a therapeutic anti-TG2-antibody and defined its specific binding epitopes. Subsequent studies demonstrated promising results, showing reduced ECM accumulation in a cell-based fibrosis model [127]. Since antibodies are considered effective and safe therapeutics, these results hold promise for developing effective and specific TG2 inhibitors.

8. Integrins-TGF-β Activation Axis

Integrins comprising the αv subunit and the β1, β3, β5, β6, or β8 subunit are crucial players in organ fibrosis disorders [128]. Suggested pro-fibrotic mechanisms of action of these integrins include activation of latent members of the TGF-β family.
TGF-β homodimers are synthesized as pro-TGF-β precursors linked covalently with the latency-associated protein (LAP, Figure 5) [129,130]. Following intracellular cleavage of the pro-domain by furins, the mature TGF-β remains non-covalently associated with LAP, forming the small latent complex (SLC). Dissociation of TGF-β from this complex is needed to activate TGFBR type 1 and type 2 [131].
Studies have found that in most cell types, the SLC is secreted in the insoluble form as the large latent complex (LLC), formed intracellularly by the association of TGF-β with the latent TGF-β-binding proteins (LTBPs) [132,133]. LLC interacts with ECM elements following secretion into the extracellular space [134,135,136]. Consequently, ECM-bound LLC is a reservoir of latent TGF-β, which must be released from the complex to activate the TGFBRs.

8.1. Integrin-Mediated Enzymatic TGF-β Activation

Two essential mechanisms of TGF-β activation release this factor from the LLC complex. The first mechanism involves enzymatic cleavage of the complex. PCP and various MMPs catalyze proteolysis of the complex [137,138,139]. Studies have suggested that the αv integrins participate in this process by combining the LLC and the LLC-digesting enzymes, facilitating the cleavage and release of active TGF-β. These integrins further optimize the TGF-β functions by enabling the proper spatial arrangement of LLC and TGFBR [140,141].

8.2. Integrin-Mediated TGF-β Activation by Cell Traction Forces

TGF-β activation is also achieved by releasing this factor from the LLC without the involvement of proteolytic enzymes [142,143]. In this case, TGF-β is freed from the latent complex by mechanical forces mediated via αv integrins interacting directly with the LAP element of the LLC complex covalently bound to the ECM (Figure 5).
In the stiff ECM environment, the αv-mediated cell binding to LAP transmits cell-generated traction forces in a way that causes LAP deformation and, consequently, TGF-β release. In contrast, cell traction forces are inefficient in deforming LAP in the compliant ECM environment due to the non-resisting matrix. Thus, in the soft ECM, this mechanism of cell traction force-dependent release of TGF-β is less effective than that of the stiff matrix environment (Figure 5).

9. Targeting αv Integrins to Reduce TGF-β-Mediated Pro-Fibrotic Cell Behavior

Targeting TGF-β has been recognized as a potent anti-fibrotic strategy [144]. However, it is now clear that direct TGF-β1 targeting is associated with severe side effects [145]. Consequently, scientists have explored the possibility of indirectly blocking this factor’s activity by targeting the various mediators associated with TGF-β signaling. One potential target is CTGF. Various tests have demonstrated that blocking this growth factor reduces fibrotic processes in many disorders [146,147].
Similarly, discovering the role of the αv integrins in TGF-β activation has opened the possibility of blocking this factor via interfering with the activation process. Experimental studies with CWHM12, a synthetic pan-inhibitor that targets all αv integrins, demonstrated attenuation of fibrosis in mouse models of liver, lung, heart, and skeletal muscle fibrosis [148]. Moreover, MK-0429, an αv integrin inhibitor, effectively blocked kidney fibrosis in rats [149].
Integrin-specific blockers have also been developed and tested (Figure 3). One small-molecule compound 8 (c8), that blocks αvβ1 integrin, demonstrated efficacy in renal, liver, and pulmonary fibrosis in mice [150,151]. Further, a cyclic RGD peptide (cilengitide) has shown anti-fibrotic activity by blocking the αvβ3 and αvβ5 integrins in a murine scleroderma model. Similarly, anti-αvβ5 integrin antibody reduced the pro-fibrotic behavior of fibroblasts derived from patients with localized scleroderma [152].
The αvβ6 integrin was also a target of anti-fibrotic treatments. In one example, Patsenker et al. demonstrated that αvβ6 antagonist EMD527040 reduced biliary fibrosis in a murine model [153]. The authors concluded that this reduction was due to the attenuation of TGF-β1 activation, thereby suggesting the utility of this integrin in blocking the progress of a broad range of fibrotic disorders.
Mouse-based studies have demonstrated that αvβ6 integrin-neutralizing antibodies attenuate renal fibrosis in Alport mice [154]. Consequently, the anti-αvβ6 therapeutic antibody, STX-100, or BG00011, was developed for clinical tests. Phase 2 clinical trials with this antibody were conducted to reduce fibrosis in kidney transplants. However, these trials (NCT00878761) were terminated due to unspecified safety concerns [155].
The STX-100 antibody was also applied in clinical trials (NCT03573505) to inhibit IPF and improve FVC parameters. These tests, however, did not show any benefits of the STX-100 antibody. Patients with IPF who received the antibody showed worsening fibrosis compared to the placebo group [156]. Consequently, the trial was terminated.

Other Integrins as Potential Anti-Fibrotic Targets

Other integrin types may play similar roles in the pathology of fibrotic disorders and should therefore be considered potential anti-fibrotic targets (Figure 3) [157].
Studies have demonstrated that among integrins that recognize the RGD protein motif, including αIIbβ3, α5β1, and α8β1 integrins, the α8β1 integrin is the most attractive target. Blocking this integrin with neutralizing antibodies demonstrated liver and pulmonary fibrosis attenuation in murine models [157,158,159].
Although none of the tested integrin inhibitors have progressed to clinical trials thus far, integrins continue to be an attractive therapeutic target to limit TGF-β-mediated fibrosis [160].

10. DDRs as an Anti-Fibrotic Target

DDR1 and DDR2 are collagen-specific receptors that belong to a family of receptor tyrosine kinases (RTKs) (Figure 3) [161,162].
In physiological conditions, these receptors play a pivotal role in embryonic development, growth, wound healing, and tissue homeostasis. They are widely distributed in various tissues. For instance, DDR1 has been detected on epithelial cells in normal tissues and fibrotic areas of the skin, liver, lung, and kidney [163]. Interstitial collagens and collagen types IV and VIII activate this receptor. DDR2 is expressed explicitly on mesenchymal cells and is activated by collagen types II, X, and interstitial collagens [164].
Research data indicate that DDRs are upregulated in fibrotic disorders, including IPF [165]. Studies found that deleting DDR1 reduced fibrosis in adipose tissue in a murine model of cardiometabolic disease. Similarly, knocking out DDR2 attenuated fibrotic changes in renal interstitial fibrosis [166,167]. These results indicate that DDRs participate in pro-fibrotic mechanisms and are valid targets for anti-fibrotic treatments.
As shown by Tao et al., the CQ-061 inhibitor of DDR1 effectively reduced the accumulation of collagen and fibronectin in TGF-β1-activated cultures of human lung fibroblasts [165]. In another example, small-molecule inhibitors of DDR1, imatinib and disatinib, pan-kinase inhibitors, were used as prototypes to develop more specific DDR1 inhibitors [168,169,170].
Although progress has been made in developing DDR inhibitors using existing general RTK blockers as molecular templates, Moll et al. identified some associated challenges [163]. The authors suggested that these challenges occur due to the conserved nature of ATP-binding pockets in all RTKs. As these pockets serve as targets for competitive inhibition of RTKs, blocking the DDR activity in a specific way is difficult. It may therefore be necessary to consider tests of type IV kinase inhibitors that are substrate-competitive rather than ATP-competitive [171,172].
Furthermore, novel drug-screening approaches, such as screening DNA-encoded libraries, may lead to future discoveries of DDR-specific inhibitors. Moll et al. reported some progress in identifying DDR1-specific inhibitors utilizing this screening approach [163].

11. Other ECM Anti-Fibrotic Targets

11.1. ED-A Fibronectin

Additional potential high-value targets have also been identified, including a fibronectin variant containing the type III extra domain A (ED-A fibronectin). Although ED-A is expressed commonly during non-fibrotic wound healing, it has also been detected in fibrotic lesions [173,174,175,176]. Various studies suggested that ED-A fibronectin activates pro-fibrotic myofibroblasts that produce a stiff ECM, providing a physical environment facilitating the integrin-mediated release of active TGF-β1 [177]. As discussed above, this creates a perfect storm for the acceleration of fibrosis.
Because of its role in fibrosis, the utility of ED-A fibronectin was evaluated in various experimental models. Studies utilizing the ED-A fibronectin function-blocking antibodies or synthetic peptides demonstrated a reduction in TGF-β1 activation and differentiation toward myofibroblasts, indicating that the ED-A variant could serve as a potentially helpful target to limit fibrosis [177,178].

11.2. Matricellular Proteins

Matricellular proteins (MCPs) are also considered extracellular anti-fibrotic targets (Figure 3). These proteins belong to a diverse family of matrix molecules that do not contribute directly to building the mechanical structure of the ECM [179]. Although in healthy adult tissues, the expression of MCPs is relatively low, during wound healing, their expression increases significantly [180,181,182].
MCPs fulfill their functions via their ability to modulate the communication between the structural elements of the ECM and cells. They bind to the ECM components and cellular receptors [183]. The receptors that participate in the MCPs’ functions include many integrins, syndecan-4, CD44, endoglin, and others [181].
MCPs have been recognized as crucial players in normal wound healing and fibrosis. Although mechanisms of their involvement are complex and poorly understood, evidence exists for their influence on myofibroblasts’ pro-fibrotic functions. In one example, an increased expression of CCN1 MCP was observed in IPF patients’ myofibroblasts in fibrotic lesions. This increase in CNN1 expression correlated with increased production of fibrotic proteins, including collagen I and fibronectin [182].
One of the critical pro-fibrotic players among MCPs is CTGF, also known as CCN2. This facilitator’s role in TGF-β1 pro-fibrotic functions has been described in many fibrotic disorders, including scleroderma and IPF [184,185].
Accordingly, an anti-CTGF antibody (FG-3019, pamrevlumab) has been selected for clinical tests that target IPF patients. The results of a phase 2 trial demonstrated improvements in the antibody-treated group compared to the placebo control.
Pamrevlumab attenuated the decline in FVC by 60% at week 48 of the treatment. Moreover, the tests indicated that the antibody was well tolerated, and its safety was similar to placebo [186]. At present, pamrevlumab is undergoing a phase 3 clinical trial (NCT03955146).
This encouraging example of CTGF targeting in patients with IPF suggests that MCPs may be attractive targets to limit the progress of fibrosis in other tissues and organs. Therefore, the utility of blocking CTGF in other fibrotic conditions has also been analyzed. Barbe et al. demonstrated a reduction in skeletal muscle fibrosis in rats treated with pamrevlumab [187]. Similarly, Vainio et al. utilized this antibody to attenuate fibrosis in myocardial infarction and improve the repair of the cardiac muscle in a murine model [188].
Other MCPs, including osteopontin (OPN), also play pro-fibrotic roles. While in healthy adult tissues the OPN expression is relatively low, in fibrotic conditions, the expression of this protein increases significantly. This increase is associated with myofibroblasts activation and increased collagen deposition [189,190,191,192].
Clinical and experimental data suggest the pro-fibrotic function of OPN in cardiovascular diseases with fibrotic features, including dilated cardiomyopathy and post-myocardial infarction injuries [193,194,195,196]. The role of this protein in cardiac fibrosis was strongly supported by various studies of cardiac fibrosis performed in OPN-null mice. These studies demonstrated that, in contrast to the wild-type mice, the mice lacking OPN had a significantly attenuated fibrotic response to pro-fibrotic stimuli [197,198,199].
Subsequently, OPN was included in the list of anti-fibrotic targets that blocked MCP-associated functions. In one study, Dai et al. showed that the OPN-neutralizing antibodies attenuated functional decline in heart functions in a murine model of dilated cardiomyopathy [200].
The above data suggest that MCPs may remain an attractive anti-fibrotic target worth exploiting beyond the MCP candidates identified thus far.

11.3. Targeting the Extracellular Vesicles in Organ Fibrosis

Extracellular vesicles (EVs) have recently been identified as targets for limiting fibrosis (Figure 3). These structures include exosomes, microvesicles, and apoptotic bodies, each characterized by distinct formation mechanisms and specific cargo [201].
EV formation includes the budding of cellular membranes via evagination or invagination. These budding mechanisms form vesicles that envelope various materials, including proteins, lipids, mRNA, and micro(mi)RNA. As EVs are released to the extracellular space, they occur in bodily fluids and the ECM.
The EVs modulate the behavior of cells by fusing with them and releasing specific cargo. In fibrotic conditions, EVs may carry pathological products formed due to injury, inflammation, and pro-fibrotic cell activation.
Some have suggested that EVs may help to develop practical anti-fibrotic approaches due to their properties of cell homing and the ability to reprogram the behavior of target cells [202]. Lenzini et al. suggested that, due to the aquaporin-1-dependent regulation of EV hydration, these vesicles are uniquely suited to penetrate a dense fibrotic tissue structure, facilitating efficient cargo delivery [203].
In support of the EVs’ involvement in fibrosis, research has demonstrated their increase in many fibrotic organs, including the lung, kidney, heart, pancreas, skin, and others [204,205,206,207,208]. Consequently, various research groups have explored the possibility of utilizing EVs to attenuate fibrotic responses to organ and tissue injuries. They proposed that applying EVs from non-fibrotic sources to fibrotic tissues and organs would provide therapeutic effects.
Many preclinical studies have indicated the potential utility of EV-based therapies. In one study, exosomes extracted from human bone marrow mesenchymal stem cells (BM-MSC) prevented and reversed pulmonary fibrosis in mice treated with bleomycin [209]. The authors demonstrated that these positive outcomes were due to changes in the macrophage population, that switched their phenotype from pro-inflammatory to homeostatic. Similarly, utilizing a hyperoxia-induced bronchopulmonary dysplasia model, Wills et al. demonstrated that applying exosomes isolated from human MSCs improved lung function via macrophage-associated mechanisms [210].
Other cell sources of the EVs suitable for reducing pulmonary fibrosis tested in animal models included amnion epithelial cells. Exosomes isolated from these cells attenuated the inflammation, epithelial damage, deposition of fibrotic ECM, and expression of TGF-β and reduced the number of myofibroblasts [211,212].
Other sources of EVs tested to improve fibrotic lung functions included macrophages [213]. Studies have established that exosomes isolated from macrophages suppress the biosynthesis of TGFBRs and collagen I. One study documented that this reduction was enabled by mechanisms involving anti-fibrotic miRNA-142-3p present in the macrophage-derived exosomes.
Similar therapeutic approaches utilizing EVs were evaluated in other organ fibrosis models [201]. Wang et al. employed human BM-MSC-derived EVs loaded with anti-fibrotic miRNA-101a. They demonstrated their protective functions in a mouse model of myocardial infarction. In particular, they observed the miRNA-101a-mediated reduction in pro-fibrotic TGF-β1, TGF-β2, and collagen [214].
In a comprehensive review, Brigstock presented further details on the clinical utility of EVs derived from various sources [201]. The author pointed out the significant potential of EVs to play a positive role in anti-fibrotic approaches. He also emphasized that our understanding of EVs’ functions in biology and pathology, including fibrosis, is still in its infancy. One factor limiting our understanding of these functions in vivo is that most of the data generated thus far are derived from cell-based studies. Therefore, more biologically relevant preclinical studies are needed to fully comprehend the prospects and limitations of EV-based approaches to reduce fibrosis.

12. Conclusions

Localized and systemic fibrotic disorders continue to result in severe medical illness and, thus, social burden. Despite substantial scientific efforts to mitigate fibrotic disease, few fibrosis-specific therapeutics have been approved for limited clinical use.
Although various tissues and organs have distinct biological functions and molecular and cellular compositions, they equally respond to fibrotic stimuli by synthesizing collagen-rich scars. Consequently, universal anti-fibrotic targets must be defined against shared pro-fibrotic mechanisms to develop broad-use anti-fibrotic therapeutics. In this context, extracellular targets, potentially limiting fibrotic healing in response to tissue injury, have become a focal point of many anti-fibrotic approaches. The central premise of these approaches, is that by modulating crucial elements of mechanisms that propagate the formation of the pro-fibrotic stiff matrix, it may be possible to reduce excessive scarring. Several potential targets associated with matrix stiffening, including LOX, TG2, PCP, and others, were identified. Their therapeutic utility has been tested at both the preclinical and clinical levels. However, despite promising preliminary results, most of these targets failed to meet the required expectations to be considered therapeutically valid.
Nevertheless, because of the essential role of the ECM in the structure and function of distinct tissues and organs during excessive scar production, targeting the ECM remains an attractive strategy to limit fibrosis. Such an approach offers the possibility of developing therapeutics to treat various fibrotic disorders, regardless of the injury site or location.
We propose that to improve the outcomes of studies targeting extracellular scarring mechanisms, it will be necessary to (i) employ relevant animal models of fibrotic disorders to test anti-fibrotic approaches in biologically relevant conditions, (ii) apply stringent criteria to describe the outcomes of anti-fibrotic approaches at the molecular, cellular, and tissue levels simultaneously, (iii) target early stages of stiff matrix formation, (iv) aim concomitantly at multiple targets, and (v) focus on mitigating excessive fibrosis rather than resolving established fibrotic tissue.

Author Contributions

Conceptualization, J.F., A.F. and M.L.W.; methodology, A.F.; data curation, J.F.; writing—original draft preparation, A.F.; writing—review and editing, J.F., M.L.W., M.R., P.K.B., J.A. and W.V.A.; visualization J.F.; supervision, A.F. All authors have read and agreed to the published version of the manuscript.

Funding

The intramural Joan and John Mullen Spine Injury Research Innovation Fund partly supported this work.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful to Liz Declan for revising the article. All illustrations were created with BioRender.com.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hinz, B. Myofibroblasts. Exp. Eye Res. 2016, 142, 56–70. [Google Scholar] [CrossRef]
  2. Ghatak, S.; Maytin, E.V.; Mack, J.A.; Hascall, V.C.; Atanelishvili, I.; Moreno Rodriguez, R.; Markwald, R.R.; Misra, S. Roles of proteoglycans and glycosaminoglycans in wound healing and fibrosis. Int. J. Cell Biol. 2015, 2015, 834893. [Google Scholar] [CrossRef] [PubMed]
  3. Moore, A.L.; Marshall, C.D.; Barnes, L.A.; Murphy, M.P.; Ransom, R.C.; Longaker, M.T. Scarless wound healing: Transitioning from fetal research to regenerative healing. Wiley Interdiscip. Rev. Dev. Biol. 2018, 7, e309. [Google Scholar] [CrossRef] [PubMed]
  4. Carr, N.J. The pathology of healing and repair. Surgery 2022, 40, 13–19. [Google Scholar] [CrossRef]
  5. Wynn, T. Cellular and molecular mechanisms of fibrosis. J. Pathol. J. Pathol. Soc. Great Br. Irel. 2008, 214, 199–210. [Google Scholar] [CrossRef] [PubMed]
  6. Eyre, D.R.; Wu, J.J.; Fernandes, R.J.; Pietka, T.A.; Weis, M.A. Recent developments in cartilage research: Matrix biology of the collagen II/IX/XI heterofibril network. Biochem. Soc. Trans. 2002, 30, 893–899. [Google Scholar] [CrossRef]
  7. Bülow, R.D.; Boor, P. Extracellular matrix in kidney fibrosis: More than just a scaffold. J. Histochem. Cytochem. 2019, 67, 643–661. [Google Scholar] [CrossRef]
  8. Ecke, A.; Lutter, A.-H.; Scholka, J.; Hansch, A.; Becker, R.; Anderer, U. Tissue specific differentiation of human chondrocytes depends on cell microenvironment and serum selection. Cells 2019, 8, 934. [Google Scholar] [CrossRef]
  9. Buchtler, S.; Grill, A.; Hofmarksrichter, S.; Stöckert, P.; Schiechl-Brachner, G.; Gomez, M.R.; Neumayer, S.; Schmidbauer, K.; Talke, Y.; Klinkhammer, B.M. Cellular origin and functional relevance of collagen I production in the kidney. J. Am. Soc. Nephrol. 2018, 29, 1859–1873. [Google Scholar] [CrossRef]
  10. Lim, X.; Tateya, I.; Tateya, T.; Munoz-Del-Rio, A.; Bless, D.M. Immediate inflammatory response and scar formation in wounded vocal folds. Ann. Otol. Rhinol. Laryngol. 2006, 115, 921–929. [Google Scholar] [CrossRef]
  11. Kwok, S.S.; Shih, K.C.; Bu, Y.; Lo, A.C.; Chan, T.C.; Lai, J.S.; Jhanji, V.; Tong, L. Systematic Review on Therapeutic Strategies to Minimize Corneal Stromal Scarring after Injury. Eye Contact Lens 2019, 45, 347–355. [Google Scholar] [CrossRef]
  12. Cholok, D.; Lee, E.; Lisiecki, J.; Agarwal, S.; Loder, S.; Ranganathan, K.; Qureshi, A.T.; Davis, T.A.; Levi, B. Traumatic muscle fibrosis: From pathway to prevention. J. Trauma Acute Care Surg. 2017, 82, 174–184. [Google Scholar] [CrossRef]
  13. Graham, J.G.; Wang, M.L.; Rivlin, M.; Beredjiklian, P.K. Biologic and mechanical aspects of tendon fibrosis after injury and repair. Connect. Tissue Res. 2018, 60, 10–20. [Google Scholar] [CrossRef]
  14. Grabowski, G.; Pacana, M.J.; Chen, E. Keloid and Hypertrophic Scar Formation, Prevention, and Management: Standard Review of Abnormal Scarring in Orthopaedic Surgery. J. Am. Acad. Orthop. Surg. 2020, 28, e408–e414. [Google Scholar] [CrossRef]
  15. Wang, M.L.; Rivlin, M.; Graham, J.G.; Beredjiklian, P.K. Peripheral nerve injury, scarring, and recovery. Connect. Tissue Res. 2018, 60, 3–9. [Google Scholar] [CrossRef]
  16. Berumen, J.; Baglieri, J.; Kisseleva, T.; Mekeel, K. Liver fibrosis: Pathophysiology and clinical implications. WIREs Mech. Dis. 2021, 13, e1499. [Google Scholar] [CrossRef]
  17. Panizo, S.; Martínez-Arias, L.; Alonso-Montes, C.; Cannata, P.; Martín-Carro, B.; Fernández-Martín, J.L.; Naves-Díaz, M.; Carrillo-López, N.; Cannata-Andía, J.B. Fibrosis in chronic kidney disease: Pathogenesis and consequences. Int. J. Mol. Sci. 2021, 22, 408. [Google Scholar] [CrossRef]
  18. Asano, Y. Systemic sclerosis. J. Dermatol. 2018, 45, 128–138. [Google Scholar] [CrossRef]
  19. Finnerty, J.P.; Ponnuswamy, A.; Dutta, P.; Abdelaziz, A.; Kamil, H. Efficacy of antifibrotic drugs, nintedanib and pirfenidone, in treatment of progressive pulmonary fibrosis in both idiopathic pulmonary fibrosis (IPF) and non-IPF: A systematic review and meta-analysis. BMC Pulm. Med. 2021, 21, 411. [Google Scholar] [CrossRef]
  20. Kinoshita, Y.; Miyamura, T.; Ikeda, T.; Ueda, Y.; Yoshida, Y.; Kushima, H.; Ishii, H. Limited efficacy of nintedanib for idiopathic pleuroparenchymal fibroelastosis. Respir. Investig. 2022, 60, 562–569. [Google Scholar] [CrossRef]
  21. Antar, S.A.; Saleh, M.A.; Al-Karmalawy, A.A. Investigating the possible mechanisms of pirfenidone to be targeted as a promising anti-inflammatory, anti-fibrotic, anti-oxidant, anti-apoptotic, anti-tumor, and/or anti-SARS-CoV-2. Life Sci. 2022, 309, 121048. [Google Scholar] [CrossRef]
  22. Knuppel, L.; Ishikawa, Y.; Aichler, M.; Heinzelmann, K.; Hatz, R.; Behr, J.; Walch, A.; Bachinger, H.P.; Eickelberg, O.; Staab-Weijnitz, C.A. A Novel Antifibrotic Mechanism of Nintedanib and Pirfenidone. Inhibition of Collagen Fibril Assembly. Am. J. Respir. Cell Mol. Biol. 2017, 57, 77–90. [Google Scholar] [CrossRef]
  23. Bicaklioglu, G.; Pirhan, D.; Yazir, Y.; Duruksu, G.; Furat Rencber, S.; Yuksel, N. Evaluation of nintedanib as a new postoperative antiscarring agent in experimental extraocular muscle surgery. Int. J. Ophthalmol. 2022, 15, 914–923. [Google Scholar] [CrossRef]
  24. Cristodor, P.L.; Nechifor, A.; Fotea, S.; Nadasdy, T.; Bahloul, Y.; Nicolescu, A.C.; Tatu, A.L. New Antifibroblastic Medication in Dermatology: Could Nintedanib Treat Scarring? Int. J. Gen. Med. 2022, 15, 7169–7172. [Google Scholar] [CrossRef] [PubMed]
  25. Hall, C.L.; Wells, A.R.; Leung, K.P. Pirfenidone reduces profibrotic responses in human dermal myofibroblasts, in vitro. Lab. Investig. 2018, 98, 640–655. [Google Scholar] [CrossRef] [PubMed]
  26. Lin, X.; Wen, J.; Liu, R.; Gao, W.; Qu, B.; Yu, M. Nintedanib inhibits TGF-beta-induced myofibroblast transdifferentiation in human Tenon’s fibroblasts. Mol. Vis. 2018, 24, 789–800. [Google Scholar]
  27. Rosenbloom, J.; Mendoza, F.A.; Jimenez, S.A. Strategies for anti-fibrotic therapies. Biochim. Biophys. Acta 2013, 1832, 1088–1103. [Google Scholar] [CrossRef]
  28. Walraven, M.; Hinz, B. Therapeutic approaches to control tissue repair and fibrosis: Extracellular matrix as a game changer. Matrix Biol. 2018, 71–72, 205–224. [Google Scholar] [CrossRef] [PubMed]
  29. Lampi, M.C.; Reinhart-King, C.A. Targeting extracellular matrix stiffness to attenuate disease: From molecular mechanisms to clinical trials. Sci. Transl. Med. 2018, 10, eaao0475. [Google Scholar] [CrossRef]
  30. Tschumperlin, D.J.; Lagares, D. Mechano-therapeutics: Targeting Mechanical Signaling in Fibrosis and Tumor Stroma. Pharmacol. Ther. 2020, 212, 107575. [Google Scholar] [CrossRef]
  31. Schuster, R.; Rockel, J.S.; Kapoor, M.; Hinz, B. The inflammatory speech of fibroblasts. Immunol. Rev. 2021, 302, 126–146. [Google Scholar] [CrossRef]
  32. Watson, R.S.; Gouze, E.; Levings, P.P.; Bush, M.L.; Kay, J.D.; Jorgensen, M.S.; Dacanay, E.A.; Reith, J.W.; Wright, T.W.; Ghivizzani, S.C. Gene delivery of TGF-beta1 induces arthrofibrosis and chondrometaplasia of synovium in vivo. Lab. Investig. 2010, 90, 1615–1627. [Google Scholar] [CrossRef]
  33. Darby, I.A.; Zakuan, N.; Billet, F.; Desmouliere, A. The myofibroblast, a key cell in normal and pathological tissue repair. Cell. Mol. Life Sci. 2016, 73, 1145–1157. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, H.W.; Liu, X.Y.; Shen, Z.F.; Yao, W.; Gong, X.B.; Huang, H.X.; Ding, G.H. An investigation of the distribution and location of mast cells affected by the stiffness of substrates as a mechanical niche. Int. J. Biol. Sci. 2018, 14, 1142–1152. [Google Scholar] [CrossRef] [PubMed]
  35. Hildebrand, K.A.; Schneider, P.S.; Mohtadi, N.G.H.; Ademola, A.; White, N.J.; Garven, A.; Walker, R.E.A.; Sajobi, T.T.; Investigators, P. PrEvention of Posttraumatic contractuRes with Ketotifen 1 (PERK 1): A Randomized Clinical Trial. J. Orthop. Trauma 2020, 34, e442–e448. [Google Scholar] [CrossRef] [PubMed]
  36. Pakshir, P.; Alizadehgiashi, M.; Wong, B.; Coelho, N.M.; Chen, X.; Gong, Z.; Shenoy, V.B.; McCulloch, C.A.; Hinz, B. Dynamic fibroblast contractions attract remote macrophages in fibrillar collagen matrix. Nat. Commun. 2019, 10, 1850. [Google Scholar] [CrossRef]
  37. Kechagia, J.Z.; Ivaska, J.; Roca-Cusachs, P. Integrins as biomechanical sensors of the microenvironment. Nat. Rev. Mol. Cell Biol. 2019, 20, 457–473. [Google Scholar] [CrossRef]
  38. Yang, S.; Plotnikov, S.V. Mechanosensitive Regulation of Fibrosis. Cells 2021, 10, 994. [Google Scholar] [CrossRef]
  39. Xue, M.; Jackson, C.J. Extracellular matrix reorganization during wound healing and its impact on abnormal scarring. Adv. Wound Care 2015, 4, 119–136. [Google Scholar] [CrossRef]
  40. Fertala, J.; Rivlin, M.; Wang, M.L.; Beredjiklian, P.K.; Steplewski, A.; Fertala, A. Collagen-rich deposit formation in the sciatic nerve after injury and surgical repair: A study of collagen-producing cells in a rabbit model. Brain Behav. 2020, 10, e01802. [Google Scholar] [CrossRef]
  41. Steplewski, A.; Fertala, J.; Beredjiklian, P.K.; Abboud, J.A.; Wang, M.L.; Namdari, S.; Barlow, J.; Rivlin, M.; Arnold, W.V.; Kostas, J.; et al. Auxiliary proteins that facilitate formation of collagen-rich deposits in the posterior knee capsule in a rabbit-based joint contracture model. J. Orthop. Res. 2016, 34, 489–501. [Google Scholar] [CrossRef]
  42. Al Shamrani, M.; Al Hati, K.; Alkatan, H.; Alharby, M.; Jastaneiah, S.; Song, J.; Edward, D.P. Pathological and immunohistochemical alterations of the cornea in congenital corneal opacification secondary to primary congenital glaucoma and peters anomaly. Cornea 2016, 35, 226–233. [Google Scholar] [CrossRef]
  43. Ayazi, M.; Zivkovic, S.; Hammel, G.; Stefanovic, B.; Ren, Y. Fibrotic scar in CNS injuries: From the cellular origins of fibroblasts to the molecular processes of fibrotic scar formation. Cells 2022, 11, 2371. [Google Scholar] [CrossRef]
  44. Lieber, R.L.; Ward, S.R. Cellular mechanisms of tissue fibrosis. 4. Structural and functional consequences of skeletal muscle fibrosis. Am. J. Physiol. Cell Physiol. 2013, 305, C241–C252. [Google Scholar] [CrossRef]
  45. Weiskirchen, R.; Weiskirchen, S.; Tacke, F. Organ and tissue fibrosis: Molecular signals, cellular mechanisms and translational implications. Mol. Asp. Med. 2019, 65, 2–15. [Google Scholar] [CrossRef]
  46. Prockop, D.J.; Berg, R.A.; Kivirikko, K.I.; Uitto, J. Intracellular Steps in the Biosynthesis of Collagen. In Biochemistry of Collagen; Ramachandran, G.N., Reddi, A.H., Eds.; Plenum: New York, NY, USA, 1976; pp. 163–237. [Google Scholar]
  47. Bottomley, M.J.; Batten, M.R.; Lumb, R.A.; Bulleid, N.J. Quality control in the endoplasmic reticulum: PDI mediates the ER retention of unassembled procollagen C-propeptides. Curr. Biol. 2001, 11, 1114–1118. [Google Scholar] [CrossRef]
  48. Prockop, D.J.; Kivirikko, K.I. Collagens: Molecular biology, diseases, and potentials for therapy. Annu. Rev. Biochem. 1995, 64, 403–434. [Google Scholar] [CrossRef]
  49. Fietzek, P.P.; Rexrodt, F.W.; Wendt, P.; Stark, M.; Kuhn, K. The covalent structure of collagen. Amino-acid sequence of peptide 1-CB6-C2. Eur. J. Biochem. 1972, 30, 163–168. [Google Scholar] [CrossRef]
  50. Morello, R.; Bertin, T.K.; Chen, Y.; Hicks, J.; Tonachini, L.; Monticone, M.; Castagnola, P.; Rauch, F.; Glorieux, F.H.; Vranka, J.; et al. CRTAP is required for prolyl 3-hydroxylation and mutations cause recessive osteogenesis imperfecta. Cell 2006, 127, 291–304. [Google Scholar] [CrossRef]
  51. Engel, J.; Prockop, D.J. The zipper-like folding of collagen triple helices and the effects of mutations that disrupt the zipper. Annu. Rev. Biophys. Biophys. Chem. 1991, 20, 137–152. [Google Scholar] [CrossRef]
  52. Lamande, S.R.; Bateman, J.F. Procollagen folding and assembly: The role of endoplasmic reticulum enzymes and molecular chaperones. Semin. Cell Dev. Biol. 1999, 10, 455–464. [Google Scholar] [CrossRef] [PubMed]
  53. Kadler, K.E.; Hojima, Y.; Prockop, D.J. Assembly of collagen fibrils de novo by cleavage of the type I pC-collagen with procollagen C-proteinase. Assay of critical concentration demonstrates that collagen self-assembly is a classical example of an entropy-driven process. J. Biol. Chem. 1987, 262, 15696–15701. [Google Scholar] [CrossRef] [PubMed]
  54. Colige, A.; Vandenberghe, I.; Thiry, M.; Lambert, C.A.; Van Beeumen, J.; Li, S.W.; Prockop, D.J.; Lapiere, C.M.; Nusgens, B.V. Cloning and characterization of ADAMTS-14, a novel ADAMTS displaying high homology with ADAMTS-2 and ADAMTS-3. J. Biol. Chem. 2002, 277, 5756–5766. [Google Scholar] [CrossRef]
  55. Colige, A. 216—Procollagen N-endopeptidase, ADAMTS2. In Handbook of Proteolytic Enzymes, 2nd ed.; Barrett, A.J., Rawlings, N.D., Woessner, J.F., Eds.; Academic Press: London, UK, 2004; pp. 737–740. [Google Scholar] [CrossRef]
  56. Li, S.W.; Sieron, A.L.; Fertala, A.; Hojima, Y.; Arnold, W.V.; Prockop, D.J. The C-proteinase that processes procollagens to fibrillar collagens is identical to the protein previously identified as bone morphogenic protein-1. Proc. Natl. Acad. Sci. USA 1996, 93, 5127–5130. [Google Scholar] [CrossRef]
  57. Adar, R.; Kessler, E.; Goldberg, B. Evidence for a protein that enhances the activity of type I procollagen C-proteinase. Collagen Relat. Res. 1986, 6, 267–277. [Google Scholar] [CrossRef] [PubMed]
  58. Hassoun, E.; Safrin, M.; Ziv, H.; Pri-Chen, S.; Kessler, E. Procollagen C-proteinase enhancer 1 (PCPE-1) as a plasma marker of muscle and liver fibrosis in mice. PLoS ONE 2016, 11, e0159606. [Google Scholar] [CrossRef]
  59. Wong, V.W.; You, F.; Januszyk, M.; Gurtner, G.C.; Kuang, A.A. Transcriptional profiling of rapamycin-treated fibroblasts from hypertrophic and keloid scars. Ann. Plast. Surg. 2014, 72, 711. [Google Scholar] [CrossRef]
  60. Broder, C.; Becker-Pauly, C. The metalloproteases meprin alpha and meprin beta: Unique enzymes in inflammation, neurodegeneration, cancer and fibrosis. Biochem. J. 2013, 450, 253–264. [Google Scholar] [CrossRef]
  61. Kofford, M.W.; Schwartz, L.B.; Schechter, N.M.; Yager, D.R.; Diegelmann, R.F.; Graham, M.F. Cleavage of type I procollagen by human mast cell chymase initiates collagen fibril formation and generates a unique carboxyl-terminal propeptide. J. Biol. Chem. 1997, 272, 7127–7131. [Google Scholar] [CrossRef]
  62. Prockop, D.J.; Fertala, A. The collagen fibril: The almost crystalline structure. J. Struct. Biol. 1998, 122, 111–118. [Google Scholar] [CrossRef]
  63. Shayegan, M.; Altindal, T.; Kiefl, E.; Forde, N.R. Intact Telopeptides Enhance Interactions between Collagens. Biophys. J. 2016, 111, 2404–2416. [Google Scholar] [CrossRef]
  64. Chung, H.J.; Steplewski, A.; Chung, K.Y.; Uitto, J.; Fertala, A. Collagen fibril formation. A new target to limit fibrosis. J. Biol. Chem. 2008, 283, 25879–25886. [Google Scholar] [CrossRef]
  65. Prockop, D.J.; Fertala, A. Inhibition of the self-assembly of collagen I into fibrils with synthetic peptides. Demonstration that assembly is driven by specific binding sites on the monomers. J. Biol. Chem. 1998, 273, 15598–15604. [Google Scholar] [CrossRef]
  66. Zou, M.; Yang, H.; Wang, H.; Wang, H.; Zhang, J.; Wei, B.; Zhang, H.; Xie, D. Detection of type I collagen fibrils formation and dissociation by a fluorescence method based on thioflavin T. Int. J. Biol. Macromol. 2016, 92, 1175–1182. [Google Scholar] [CrossRef]
  67. Yeh, A.T.; Choi, B.; Nelson, J.S.; Tromberg, B.J. Reversible dissociation of collagen in tissues. J. Investig. Dermatol. 2003, 121, 1332–1335. [Google Scholar] [CrossRef] [PubMed]
  68. Grenard, P.; Bresson-Hadni, S.; El Alaoui, S.; Chevallier, M.; Vuitton, D.A.; Ricard-Blum, S. Transglutaminase-mediated cross-linking is involved in the stabilization of extracellular matrix in human liver fibrosis. J. Hepatol. 2001, 35, 367–375. [Google Scholar] [CrossRef] [PubMed]
  69. Olsen, K.C.; Sapinoro, R.E.; Kottmann, R.M.; Kulkarni, A.A.; Iismaa, S.E.; Johnson, G.V.; Thatcher, T.H.; Phipps, R.P.; Sime, P.J. Transglutaminase 2 and its role in pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 2011, 184, 699–707. [Google Scholar] [CrossRef]
  70. Philp, C.J.; Siebeke, I.; Clements, D.; Miller, S.; Habgood, A.; John, A.E.; Navaratnam, V.; Hubbard, R.B.; Jenkins, G.; Johnson, S.R. Extracellular Matrix Cross-Linking Enhances Fibroblast Growth and Protects against Matrix Proteolysis in Lung Fibrosis. Am. J. Respir. Cell Mol. Biol. 2018, 58, 594–603. [Google Scholar] [CrossRef]
  71. Guerrot, D.; Kerroch, M.; Placier, S.; Vandermeersch, S.; Trivin, C.; Mael-Ainin, M.; Chatziantoniou, C.; Dussaule, J.-C. Discoidin domain receptor 1 is a major mediator of inflammation and fibrosis in obstructive nephropathy. Am. J. Pathol. 2011, 179, 83–91. [Google Scholar] [CrossRef] [PubMed]
  72. Song, S.; Shackel, N.A.; Wang, X.M.; Ajami, K.; McCaughan, G.W.; Gorrell, M.D. Discoidin domain receptor 1: Isoform expression and potential functions in cirrhotic human liver. Am. J. Pathol. 2011, 178, 1134–1144. [Google Scholar] [CrossRef]
  73. Huang, Y.; Arora, P.; McCulloch, C.A.; Vogel, W.F. The collagen receptor DDR1 regulates cell spreading and motility by associating with myosin IIA. J. Cell Sci. 2009, 122, 1637–1646. [Google Scholar] [CrossRef]
  74. Even-Ram, S.; Doyle, A.D.; Conti, M.A.; Matsumoto, K.; Adelstein, R.S.; Yamada, K.M. Myosin IIA regulates cell motility and actomyosin–microtubule crosstalk. Nat. Cell Biol. 2007, 9, 299–309. [Google Scholar] [CrossRef]
  75. Giannone, G.; Dubin-Thaler, B.J.; Rossier, O.; Cai, Y.; Chaga, O.; Jiang, G.; Beaver, W.; Döbereiner, H.-G.; Freund, Y.; Borisy, G. Lamellipodial actin mechanically links myosin activity with adhesion-site formation. Cell 2007, 128, 561–575. [Google Scholar] [CrossRef]
  76. Abuammah, A.; Maimari, N.; Towhidi, L.; Frueh, J.; Chooi, K.Y.; Warboys, C.; Krams, R. New developments in mechanotransduction: Cross talk of the Wnt, TGF-β and Notch signalling pathways in reaction to shear stress. Curr. Opin. Biomed. Eng. 2018, 5, 96–104. [Google Scholar] [CrossRef]
  77. Mahler, G.J.; Frendl, C.M.; Cao, Q.; Butcher, J.T. Effects of shear stress pattern and magnitude on mesenchymal transformation and invasion of aortic valve endothelial cells. Biotechnol. Bioeng. 2014, 111, 2326–2337. [Google Scholar] [CrossRef]
  78. Goswami, C.; Kuhn, J.; Heppenstall, P.A.; Hucho, T. Importance of non-selective cation channel TRPV4 interaction with cytoskeleton and their reciprocal regulations in cultured cells. PLoS ONE 2010, 5, e11654. [Google Scholar] [CrossRef] [PubMed]
  79. Kirby, T.J.; Lammerding, J. Emerging views of the nucleus as a cellular mechanosensor. Nat. Cell Biol. 2018, 20, 373–381. [Google Scholar] [CrossRef] [PubMed]
  80. Swift, J.; Ivanovska, I.L.; Buxboim, A.; Harada, T.; Dingal, P.C.; Pinter, J.; Pajerowski, J.D.; Spinler, K.R.; Shin, J.W.; Tewari, M.; et al. Nuclear lamin-A scales with tissue stiffness and enhances matrix-directed differentiation. Science 2013, 341, 1240104. [Google Scholar] [CrossRef]
  81. Haak, A.J.; Kostallari, E.; Sicard, D.; Ligresti, G.; Choi, K.M.; Caporarello, N.; Jones, D.L.; Tan, Q.; Meridew, J.; Diaz Espinosa, A.M.; et al. Selective YAP/TAZ inhibition in fibroblasts via dopamine receptor D1 agonism reverses fibrosis. Sci. Transl. Med. 2019, 11, eaau6296. [Google Scholar] [CrossRef]
  82. Lagares, D.; Santos, A.; Grasberger, P.E.; Liu, F.; Probst, C.K.; Rahimi, R.A.; Sakai, N.; Kuehl, T.; Ryan, J.; Bhola, P.; et al. Targeted apoptosis of myofibroblasts with the BH3 mimetic ABT-263 reverses established fibrosis. Sci. Transl. Med. 2017, 9, eaal3765. [Google Scholar] [CrossRef]
  83. Liu, F.; Lagares, D.; Choi, K.M.; Stopfer, L.; Marinkovic, A.; Vrbanac, V.; Probst, C.K.; Hiemer, S.E.; Sisson, T.H.; Horowitz, J.C.; et al. Mechanosignaling through YAP and TAZ drives fibroblast activation and fibrosis. Am. J. Physiol. Lung Cell. Mol. Physiol. 2015, 308, L344–L357. [Google Scholar] [CrossRef] [PubMed]
  84. Szeto, S.G.; Narimatsu, M.; Lu, M.; He, X.; Sidiqi, A.M.; Tolosa, M.F.; Chan, L.; De Freitas, K.; Bialik, J.F.; Majumder, S.; et al. YAP/TAZ Are Mechanoregulators of TGF-beta-Smad Signaling and Renal Fibrogenesis. J. Am. Soc. Nephrol. 2016, 27, 3117–3128. [Google Scholar] [CrossRef] [PubMed]
  85. Shiwen, X.; Stratton, R.; Nikitorowicz-Buniak, J.; Ahmed-Abdi, B.; Ponticos, M.; Denton, C.; Abraham, D.; Takahashi, A.; Suki, B.; Layne, M.D.; et al. A Role of Myocardin Related Transcription Factor-A (MRTF-A) in Scleroderma Related Fibrosis. PLoS ONE 2015, 10, e0126015. [Google Scholar] [CrossRef]
  86. Velasquez, L.S.; Sutherland, L.B.; Liu, Z.; Grinnell, F.; Kamm, K.E.; Schneider, J.W.; Olson, E.N.; Small, E.M. Activation of MRTF-A-dependent gene expression with a small molecule promotes myofibroblast differentiation and wound healing. Proc. Natl. Acad. Sci. USA 2013, 110, 16850–16855. [Google Scholar] [CrossRef]
  87. Holmes, D.F.; Mould, A.P.; Chapman, J.A. Morphology of sheet-like assemblies of pN-collagen, pC-collagen and procollagen studied by scanning transmission electron microscopy mass measurements. J. Mol. Biol. 1991, 220, 111–123. [Google Scholar] [CrossRef]
  88. Smith, L.T.; Wertelecki, W.; Milstone, L.M.; Petty, E.M.; Seashore, M.R.; Braverman, I.M.; Jenkins, T.G.; Byers, P.H. Human dermatosparaxis: A form of Ehlers-Danlos syndrome that results from failure to remove the amino-terminal propeptide of type I procollagen. Am. J. Hum. Genet. 1992, 51, 235–244. [Google Scholar]
  89. Ovens, A.; Joule, J.A.; Kadler, K.E. Design and synthesis of acidic dipeptide hydroxamate inhibitors of procollagen C-proteinase. J. Pept. Sci. 2000, 6, 489–495. [Google Scholar] [CrossRef]
  90. Bekhouche, M.; Colige, A. The procollagen N-proteinases ADAMTS2, 3 and 14 in pathophysiology. Matrix Biol. 2015, 44–46, 46–53. [Google Scholar] [CrossRef]
  91. Suzuki, N.; Labosky, P.A.; Furuta, Y.; Hargett, L.; Dunn, R.; Fogo, A.B.; Takahara, K.; Peters, D.M.; Greenspan, D.S.; Hogan, B.L. Failure of ventral body wall closure in mouse embryos lacking a procollagen C-proteinase encoded by Bmp1, a mammalian gene related to Drosophila tolloid. Development 1996, 122, 3587–3595. [Google Scholar] [CrossRef] [PubMed]
  92. Li, S.W.; Arita, M.; Fertala, A.; Bao, Y.; Kopen, G.C.; Langsjo, T.K.; Hyttinen, M.M.; Helminen, H.J.; Prockop, D.J. Transgenic mice with inactive alleles for procollagen N-proteinase (ADAMTS-2) develop fragile skin and male sterility. Biochem. J. 2001, 355, 271–278. [Google Scholar] [CrossRef]
  93. Lagoutte, P.; Bettler, E.; Vadon-Le Goff, S.; Moali, C. Procollagen C-proteinase enhancer-1 (PCPE-1), a potential biomarker and therapeutic target for fibrosis. Matrix Biol. Plus 2021, 11, 100062. [Google Scholar] [CrossRef]
  94. Fertala, J.; Steplewski, A.; Kostas, J.; Beredjiklian, P.; Williams, G.; Arnold, W.; Abboud, J.; Bhardwaj, A.; Hou, C.; Fertala, A. Engineering and characterization of the chimeric antibody that targets the C-terminal telopeptide of the alpha2 chain of human collagen I: A next step in the quest to reduce localized fibrosis. Connect. Tissue Res. 2013, 54, 187–196. [Google Scholar] [CrossRef]
  95. Fertala, J.; Kostas, J.; Hou, C.; Steplewski, A.; Beredjiklian, P.; Abboud, J.; Arnold, W.V.; Williams, G.; Fertala, A. Testing the anti-fibrotic potential of the single-chain Fv antibody against the alpha2 C-terminal telopeptide of collagen I. Connect. Tissue Res. 2014, 55, 115–122. [Google Scholar] [CrossRef]
  96. Fertala, J.; Romero, F.; Summer, R.; Fertala, A. Target-Specific Delivery of an Antibody That Blocks the Formation of Collagen Deposits in Skin and Lung. Monoclon. Antib. Immunodiagn. Immunother. 2017, 36, 199–207. [Google Scholar] [CrossRef] [PubMed]
  97. Steplewski, A.; Fertala, J.; Beredjiklian, P.K.; Abboud, J.A.; Wang, M.L.Y.; Namdari, S.; Barlow, J.; Rivlin, M.; Arnold, W.V.; Kostas, J.; et al. Blocking collagen fibril formation in injured knees reduces flexion contracture in a rabbit model. J. Orthop. Res. 2017, 35, 1038–1046. [Google Scholar] [CrossRef]
  98. Steplewski, A.; Fertala, J.; Tomlinson, R.E.; Wang, M.L.; Donahue, A.; Arnold, W.V.; Rivlin, M.; Beredjiklian, P.K.; Abboud, J.A.; Namdari, S.; et al. Mechanisms of reducing joint stiffness by blocking collagen fibrillogenesis in a rabbit model of posttraumatic arthrofibrosis. PLoS ONE 2021, 16, e0257147. [Google Scholar] [CrossRef] [PubMed]
  99. Font, B.; Eichenberger, D.; Goldschmidt, D.; Boutillon, M.M.; Hulmes, D.J. Structural requirements for fibromodulin binding to collagen and the control of type I collagen fibrillogenesis—Critical roles for disulphide bonding and the C-terminal region. Eur. J. Biochem. 1998, 254, 580–587. [Google Scholar] [CrossRef]
  100. Flynn, L.A.; Blissett, A.R.; Calomeni, E.P.; Agarwal, G. Inhibition of collagen fibrillogenesis by cells expressing soluble extracellular domains of DDR1 and DDR2. J. Mol. Biol. 2010, 395, 533–543. [Google Scholar] [CrossRef]
  101. Perumal, S.; Dubey, K.; Badhwar, R.; George, K.J.; Sharma, R.K.; Bagler, G.; Madhan, B.; Kar, K. Capsaicin inhibits collagen fibril formation and increases the stability of collagen fibers. Eur. Biophys. J. 2015, 44, 69–76. [Google Scholar] [CrossRef] [PubMed]
  102. Bharathy, H.; Fathima, N.N. Exploiting oleuropein for inhibiting collagen fibril formation. Int. J. Biol. Macromol. 2017, 101, 179–186. [Google Scholar] [CrossRef]
  103. Rasheeda, K.; Fathima, N.N. Trigonelline hydrochloride: A promising inhibitor for type I collagen fibrillation. Colloids Surf. B Biointerfaces 2018, 170, 273–279. [Google Scholar] [CrossRef] [PubMed]
  104. Rasheeda, K.; Muvva, C.; Fathima, N.N. Governing the Inhibition of Reconstituted Collagen Type I Assemblies Mediated Through Noncovalent Forces of (+/-)-alpha Lipoic Acid. Langmuir 2019, 35, 980–989. [Google Scholar] [CrossRef]
  105. Simoes, F.C.; Cahill, T.J.; Kenyon, A.; Gavriouchkina, D.; Vieira, J.M.; Sun, X.; Pezzolla, D.; Ravaud, C.; Masmanian, E.; Weinberger, M.; et al. Macrophages directly contribute collagen to scar formation during zebrafish heart regeneration and mouse heart repair. Nat. Commun. 2020, 11, 600. [Google Scholar] [CrossRef] [PubMed]
  106. Sutherland, T.E.; Dyer, D.P.; Allen, J.E. The extracellular matrix and the immune system: A mutually dependent relationship. Science 2023, 379, eabp8964. [Google Scholar] [CrossRef]
  107. Solis, A.G.; Bielecki, P.; Steach, H.R.; Sharma, L.; Harman, C.C.D.; Yun, S.; de Zoete, M.R.; Warnock, J.N.; To, S.D.F.; York, A.G.; et al. Mechanosensation of cyclical force by PIEZO1 is essential for innate immunity. Nature 2019, 573, 69–74. [Google Scholar] [CrossRef] [PubMed]
  108. Aciksari, K.; Yanar, H.T.; Hepgul, G.; Ozucelik, D.N.; Yanar, F.; Agcaoglu, O.; Eser, M.; Tanriverdi, G.; Topacoglu, H.; Ayvaci, B.M.; et al. The Effect of Beta-Aminopropionitrile and Prednisolone on the Prevention of Fibrosis in Alkali Esophageal Burns: An Experimental Study. Gastroenterol. Res. Pract. 2013, 2013, 574260. [Google Scholar] [CrossRef]
  109. Riley, D.J.; Kerr, J.S.; Berg, R.A.; Ianni, B.D.; Pietra, G.G.; Edelman, N.H.; Prockop, D.J. beta-Aminopropionitrile prevents bleomycin-induced pulmonary fibrosis in the hamster. Am. Rev. Respir. Dis. 1982, 125, 67–73. [Google Scholar]
  110. Mohseni, R.; Arab Sadeghabadi, Z.; Goodarzi, M.T.; Karimi, J. Co-administration of resveratrol and beta-aminopropionitrile attenuates liver fibrosis development via targeting lysyl oxidase in CCl4-induced liver fibrosis in rats. Immunopharmacol. Immunotoxicol. 2019, 41, 644–651. [Google Scholar] [CrossRef]
  111. Erasmus, M.; Samodien, E.; Lecour, S.; Cour, M.; Lorenzo, O.; Dludla, P.; Pheiffer, C.; Johnson, R. Linking LOXL2 to Cardiac Interstitial Fibrosis. Int. J. Mol. Sci. 2020, 21, 5913. [Google Scholar] [CrossRef]
  112. Barry-Hamilton, V.; Spangler, R.; Marshall, D.; McCauley, S.; Rodriguez, H.M.; Oyasu, M.; Mikels, A.; Vaysberg, M.; Ghermazien, H.; Wai, C.; et al. Allosteric inhibition of lysyl oxidase-like-2 impedes the development of a pathologic microenvironment. Nat. Med. 2010, 16, 1009–1017. [Google Scholar] [CrossRef]
  113. Raghu, G.; Brown, K.K.; Collard, H.R.; Cottin, V.; Gibson, K.F.; Kaner, R.J.; Lederer, D.J.; Martinez, F.J.; Noble, P.W.; Song, J.W.; et al. Efficacy of simtuzumab versus placebo in patients with idiopathic pulmonary fibrosis: A randomised, double-blind, controlled, phase 2 trial. Lancet Respir. Med. 2017, 5, 22–32. [Google Scholar] [CrossRef] [PubMed]
  114. Hecht, J.R.; Benson, A.B., III; Vyushkov, D.; Yang, Y.; Bendell, J.; Verma, U. A Phase II, Randomized, Double-Blind, Placebo-Controlled Study of Simtuzumab in Combination with FOLFIRI for the Second-Line Treatment of Metastatic KRAS Mutant Colorectal Adenocarcinoma. Oncologist 2017, 22, e223–e243. [Google Scholar] [CrossRef]
  115. Benson, A.B., III; Wainberg, Z.A.; Hecht, J.R.; Vyushkov, D.; Dong, H.; Bendell, J.; Kudrik, F. A Phase II Randomized, Double-Blind, Placebo-Controlled Study of Simtuzumab or Placebo in Combination with Gemcitabine for the First-Line Treatment of Pancreatic Adenocarcinoma. Oncologist 2017, 22, e215–e241. [Google Scholar] [CrossRef] [PubMed]
  116. Harrison, S.A.; Abdelmalek, M.F.; Caldwell, S.; Shiffman, M.L.; Diehl, A.M.; Ghalib, R.; Lawitz, E.J.; Rockey, D.C.; Schall, R.A.; Jia, C.; et al. Simtuzumab Is Ineffective for Patients With Bridging Fibrosis or Compensated Cirrhosis Caused by Nonalcoholic Steatohepatitis. Gastroenterology 2018, 155, 1140–1153. [Google Scholar] [CrossRef]
  117. Puente, A.; Fortea, J.I.; Cabezas, J.; Arias Loste, M.T.; Iruzubieta, P.; Llerena, S.; Huelin, P.; Fabrega, E.; Crespo, J. LOXL2-A New Target in Antifibrogenic Therapy? Int. J. Mol. Sci. 2019, 20, 1634. [Google Scholar] [CrossRef] [PubMed]
  118. Oh, K.; Park, H.-B.; Byoun, O.-J.; Shin, D.-M.; Jeong, E.M.; Kim, Y.W.; Kim, Y.S.; Melino, G.; Kim, I.-G.; Lee, D.-S. Epithelial transglutaminase 2 is needed for T cell interleukin-17 production and subsequent pulmonary inflammation and fibrosis in bleomycin-treated mice. J. Exp. Med. 2011, 208, 1707–1719. [Google Scholar] [CrossRef]
  119. Johnson, T.S.; Fisher, M.; Haylor, J.L.; Hau, Z.; Skill, N.J.; Jones, R.; Saint, R.; Coutts, I.; Vickers, M.E.; El Nahas, A.M. Transglutaminase inhibition reduces fibrosis and preserves function in experimental chronic kidney disease. J. Am. Soc. Nephrol. 2007, 18, 3078–3088. [Google Scholar] [CrossRef]
  120. Badarau, E.; Wang, Z.; Rathbone, D.L.; Costanzi, A.; Thibault, T.; Murdoch, C.E.; El Alaoui, S.; Bartkeviciute, M.; Griffin, M. Development of potent and selective tissue transglutaminase inhibitors: Their effect on TG2 function and application in pathological conditions. Chem. Biol. 2015, 22, 1347–1361. [Google Scholar] [CrossRef]
  121. Wang, Z.; Stuckey, D.J.; Murdoch, C.E.; Camelliti, P.; Lip, G.Y.; Griffin, M. Cardiac fibrosis can be attenuated by blocking the activity of transglutaminase 2 using a selective small-molecule inhibitor. Cell Death Dis. 2018, 9, 613. [Google Scholar] [CrossRef]
  122. Kunoki, S.; Tatsukawa, H.; Sakai, Y.; Kinashi, H.; Kariya, T.; Suzuki, Y.; Mizuno, M.; Yamaguchi, M.; Sasakura, H.; Ikeno, M. Inhibition of transglutaminase 2 reduces peritoneal injury in a chlorhexidine-induced peritoneal fibrosis model. Lab. Investig. 2023, 103, 100050. [Google Scholar] [CrossRef]
  123. Troilo, H.; Steer, R.; Collins, R.F.; Kielty, C.M.; Baldock, C. Independent multimerization of Latent TGFβ Binding Protein-1 stabilized by cross-linking and enhanced by heparan sulfate. Sci. Rep. 2016, 6, 34347. [Google Scholar] [CrossRef] [PubMed]
  124. Harrison, C.A.; Layton, C.M.; Hau, Z.; Bullock, A.J.; Johnson, T.S.; MacNeil, S. Transglutaminase inhibitors induce hyperproliferation and parakeratosis in tissue-engineered skin. Br. J. Dermatol. 2007, 156, 247–257. [Google Scholar] [CrossRef]
  125. Keillor, J.W.; Apperley, K.Y.P.; Akbar, A. Inhibitors of tissue transglutaminase. Trends Pharmacol. Sci. 2015, 36, 32–40. [Google Scholar] [CrossRef] [PubMed]
  126. Freund, K.F.; Doshi, K.P.; Gaul, S.L.; Claremon, D.A.; Remy, D.C.; Baldwin, J.J.; Pitzenberger, S.M.; Stern, A.M. Transglutaminase inhibition by 2-[(2-oxopropyl) thio] imidazolium derivatives: Mechanism of factor XIIIa inactivation. Biochemistry 1994, 33, 10109–10119. [Google Scholar] [CrossRef] [PubMed]
  127. Maamra, M.; Benayad, O.M.; Matthews, D.; Kettleborough, C.; Atkinson, J.; Cain, K.; Bon, H.; Brand, H.; Parkinson, M.; Watson, P.F. Transglutaminase 2: Development of therapeutic antibodies reveals four inhibitory epitopes and confirms extracellular function in fibrotic remodelling. Br. J. Pharmacol. 2022, 179, 2697–2712. [Google Scholar] [CrossRef]
  128. Rahman, S.R.; Roper, J.A.; Grove, J.I.; Aithal, G.P.; Pun, K.T.; Bennett, A.J. Integrins as a drug target in liver fibrosis. Liver Int. 2022, 42, 507–521. [Google Scholar] [CrossRef]
  129. Lawrence, D.A.; Pircher, R.; Krycève-Martinerie, C.; Jullien, P. Normal embryo fibroblasts release transforming growth factors in a latent form. J. Cell. Physiol. 1984, 121, 184–188. [Google Scholar] [CrossRef]
  130. Annes, J.P.; Munger, J.S.; Rifkin, D.B. Making sense of latent TGFβ activation. J. Cell Sci. 2003, 116, 217–224. [Google Scholar] [CrossRef]
  131. Dubois, C.M.; Laprise, M.-H.; Blanchette, F.; Gentry, L.E.; Leduc, R. Processing of transforming growth factor β1 precursor by human furin convertase. J. Biol. Chem. 1995, 270, 10618–10624. [Google Scholar] [CrossRef]
  132. Dallas, S.; Park-Snyder, S.; Miyazono, K.; Twardzik, D.; Mundy, G.; Bonewald, L. Characterization and autoregulation of latent transforming growth factor beta (TGF beta) complexes in osteoblast-like cell lines. Production of a latent complex lacking the latent TGF beta-binding protein. J. Biol. Chem. 1994, 269, 6815–6821. [Google Scholar] [CrossRef]
  133. Saharinen, J.; Keski-Oja, J. Specific sequence motif of 8-Cys repeats of TGF-β binding proteins, LTBPs, creates a hydrophobic interaction surface for binding of small latent TGF-β. Mol. Biol. Cell 2000, 11, 2691–2704. [Google Scholar] [CrossRef] [PubMed]
  134. Ten Dijke, P.; Arthur, H.M. Extracellular control of TGFβ signalling in vascular development and disease. Nat. Rev. Mol. Cell Biol. 2007, 8, 857–869. [Google Scholar] [CrossRef] [PubMed]
  135. Taipale, J.; Miyazono, K.; Heldin, C.-H.; Keski-Oja, J. Latent transforming growth factor-beta 1 associates to fibroblast extracellular matrix via latent TGF-beta binding protein. J. Cell Biol. 1994, 124, 171–181. [Google Scholar] [CrossRef] [PubMed]
  136. Schoppet, M.; Chavakis, T.; Al-Fakhri, N.; Kanse, S.M.; Preissner, K.T. Molecular interactions and functional interference between vitronectin and transforming growth factor-β. Lab. Investig. 2002, 82, 37–46. [Google Scholar] [CrossRef]
  137. Ge, G.; Greenspan, D.S. BMP1 controls TGFβ1 activation via cleavage of latent TGFβ-binding protein. J. Cell Biol. 2006, 175, 111–120. [Google Scholar] [CrossRef]
  138. D’Angelo, M.; Billings, P.C.; Pacifici, M.; Leboy, P.S.; Kirsch, T. Authentic matrix vesicles contain active metalloproteases (MMP): A role for matrix vesicle-associated MMP-13 in activation of transforming growth factor-β. J. Biol. Chem. 2001, 276, 11347–11353. [Google Scholar] [CrossRef]
  139. Maeda, S.; Dean, D.; Gomez, R.; Schwartz, Z.; Boyan, B. The first stage of transforming growth factor β1 activation is release of the large latent complex from the extracellular matrix of growth plate chondrocytes by matrix vesicle stromelysin-1 (MMP-3). Calcif. Tissue Int. 2002, 70, 54–65. [Google Scholar] [CrossRef]
  140. Wipff, P.-J.; Hinz, B. Integrins and the activation of latent transforming growth factor β1–an intimate relationship. Eur. J. Cell Biol. 2008, 87, 601–615. [Google Scholar] [CrossRef]
  141. Scaffidi, A.K.; Petrovic, N.; Moodley, Y.P.; Fogel-Petrovic, M.; Kroeger, K.M.; Seeber, R.M.; Eidne, K.A.; Thompson, P.J.; Knight, D.A. αvβ3 Integrin interacts with the transforming growth factor β (TGFβ) type II receptor to potentiate the proliferative effects of TGFβ1 in living human lung fibroblasts. J. Biol. Chem. 2004, 279, 37726–37733. [Google Scholar] [CrossRef]
  142. Munger, J.S.; Harpel, J.G.; Giancotti, F.G.; Rifkin, D.B. Interactions between growth factors and integrins: Latent forms of transforming growth factor-β are ligands for the integrin αvβ1. Mol. Biol. Cell 1998, 9, 2627–2638. [Google Scholar] [CrossRef]
  143. Sarrazy, V.; Koehler, A.; Chow, M.L.; Zimina, E.; Li, C.X.; Kato, H.; Caldarone, C.A.; Hinz, B. Integrins αvβ5 and αvβ3 promote latent TGF-β1 activation by human cardiac fibroblast contraction. Cardiovasc. Res. 2014, 102, 407–417. [Google Scholar] [CrossRef]
  144. Peng, D.; Fu, M.; Wang, M.; Wei, Y.; Wei, X. Targeting TGF-β signal transduction for fibrosis and cancer therapy. Mol. Cancer 2022, 21, 104. [Google Scholar] [CrossRef]
  145. Zhang, M.; Zhang, Y.Y.; Chen, Y.; Wang, J.; Wang, Q.; Lu, H. TGF-β signaling and resistance to cancer therapy. Front. Cell Dev. Biol. 2021, 9, 3310. [Google Scholar] [CrossRef]
  146. Ihn, H. Pathogenesis of fibrosis: Role of TGF-beta and CTGF. Curr. Opin. Rheumatol. 2002, 14, 681–685. [Google Scholar] [CrossRef]
  147. Kok, H.M.; Falke, L.L.; Goldschmeding, R.; Nguyen, T.Q. Targeting CTGF, EGF and PDGF pathways to prevent progression of kidney disease. Nat. Rev. Nephrol. 2014, 10, 700–711. [Google Scholar] [CrossRef]
  148. Henderson, N.C.; Arnold, T.D.; Katamura, Y.; Giacomini, M.M.; Rodriguez, J.D.; McCarty, J.H.; Pellicoro, A.; Raschperger, E.; Betsholtz, C.; Ruminski, P.G.; et al. Targeting of αv integrin identifies a core molecular pathway that regulates fibrosis in several organs. Nat. Med. 2013, 19, 1617–1624. [Google Scholar] [CrossRef]
  149. Zhou, X.; Zhang, J.; Haimbach, R.; Zhu, W.; Mayer-Ezell, R.; Garcia-Calvo, M.; Smith, E.; Price, O.; Kan, Y.; Zycband, E.; et al. An integrin antagonist (MK-0429) decreases proteinuria and renal fibrosis in the ZSF1 rat diabetic nephropathy model. Pharm. Res. Perspect. 2017, 5, e00354. [Google Scholar] [CrossRef]
  150. Reed, N.I.; Jo, H.; Chen, C.; Tsujino, K.; Arnold, T.D.; DeGrado, W.F.; Sheppard, D. The αvβ1 integrin plays a critical in vivo role in tissue fibrosis. Sci. Transl. Med. 2015, 7, 288ra279. [Google Scholar] [CrossRef]
  151. Chang, Y.; Lau, W.L.; Jo, H.; Tsujino, K.; Gewin, L.; Reed, N.I.; Atakilit, A.; Nunes, A.C.F.; DeGrado, W.F.; Sheppard, D. Pharmacologic blockade of αvβ1 integrin ameliorates renal failure and fibrosis in vivo. J. Am. Soc. Nephrol. 2017, 28, 1998–2005. [Google Scholar] [CrossRef]
  152. Asano, Y.; Ihn, H.; Jinnin, M.; Mimura, Y.; Tamaki, K. Involvement of αvβ5 integrin in the establishment of autocrine TGF-β signaling in dermal fibroblasts derived from localized scleroderma. J. Investig. Dermatol. 2006, 126, 1761–1769. [Google Scholar] [CrossRef]
  153. Patsenker, E.; Popov, Y.; Stickel, F.; Jonczyk, A.; Goodman, S.L.; Schuppan, D. Inhibition of integrin αvβ6 on cholangiocytes blocks transforming growth factor-β activation and retards biliary fibrosis progression. Gastroenterology 2008, 135, 660–670. [Google Scholar] [CrossRef] [PubMed]
  154. Hahm, K.; Lukashev, M.E.; Luo, Y.; Yang, W.J.; Dolinski, B.M.; Weinreb, P.H.; Simon, K.J.; Wang, L.C.; Leone, D.R.; Lobb, R.R. αvβ6 integrin regulates renal fibrosis and inflammation in Alport mouse. Am. J. Pathol. 2007, 170, 110–125. [Google Scholar] [CrossRef] [PubMed]
  155. Ruiz-Ortega, M.; Lamas, S.; Ortiz, A. Antifibrotic Agents for the Management of CKD: A Review. Am. J. Kidney Dis. 2022, 80, 251–263. [Google Scholar] [CrossRef] [PubMed]
  156. Raghu, G.; Mouded, M.; Chambers, D.C.; Martinez, F.J.; Richeldi, L.; Lancaster, L.H.; Hamblin, M.J.; Gibson, K.F.; Rosas, I.O.; Prasse, A.; et al. A Phase IIb Randomized Clinical Study of an Anti-alpha(v)beta(6) Monoclonal Antibody in Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2022, 206, 1128–1139. [Google Scholar] [CrossRef]
  157. Yokosaki, Y.; Nishimichi, N. New Therapeutic Targets for Hepatic Fibrosis in the Integrin Family, α8β1 and α11β1, Induced Specifically on Activated Stellate Cells. Int. J. Mol. Sci. 2021, 22, 12794. [Google Scholar] [CrossRef]
  158. Nishimichi, N.; Tsujino, K.; Kanno, K.; Sentani, K.; Kobayashi, T.; Chayama, K.; Sheppard, D.; Yokosaki, Y. Induced hepatic stellate cell integrin, α8β1, enhances cellular contractility and TGFβ activity in liver fibrosis. J. Pathol. 2021, 253, 366–373. [Google Scholar] [CrossRef]
  159. Nishimichi, N.; Kanno, K.; Sentani, K.; Yasui, W.; Yokosaki, Y. A possible therapeutic agent for pulmonary fibrosis: Antibody against integrin α8β1. Eur. Respir. J. 2013, 42, P4885. [Google Scholar]
  160. Slack, R.; Macdonald, S.; Roper, J.; Jenkins, R.; Hatley, R. Emerging therapeutic opportunities for integrin inhibitors. Nat. Rev. Drug Discov. 2022, 21, 60–78. [Google Scholar] [CrossRef]
  161. Leitinger, B. Transmembrane collagen receptors. Annu. Rev. Cell Dev. Biol. 2011, 27, 265–290. [Google Scholar] [CrossRef]
  162. Leitinger, B. Chapter Two—Discoidin Domain Receptor Functions in Physiological and Pathological Conditions. In International Review of Cell and Molecular Biology; Jeon, K.W., Ed.; Academic Press: Cambridge, MA, USA, 2014; Volume 310, pp. 39–87. [Google Scholar]
  163. Moll, S.; Desmoulière, A.; Moeller, M.J.; Pache, J.C.; Badi, L.; Arcadu, F.; Richter, H.; Satz, A.; Uhles, S.; Cavalli, A.; et al. DDR1 role in fibrosis and its pharmacological targeting. Biochim. Biophys. Acta Mol. Cell Res. 2019, 1866, 118474. [Google Scholar] [CrossRef]
  164. Leitinger, B. Discoidin domain receptor functions in physiological and pathological conditions. Int. Rev. Cell Mol. Biol. 2014, 310, 39–87. [Google Scholar]
  165. Tao, J.; Zhang, M.; Wen, Z.; Wang, B.; Zhang, L.; Ou, Y.; Tang, X.; Yu, X.; Jiang, Q. Inhibition of EP300 and DDR1 synergistically alleviates pulmonary fibrosis in vitro and in vivo. Biomed. Pharmacother. 2018, 106, 1727–1733. [Google Scholar] [CrossRef]
  166. Lino, M.; Ngai, D.; Liu, A.; Mohabeer, A.; Harper, C.; Caruso, L.-L.; Schroer, S.A.; Fu, F.; McKee, T.; Giacca, A.; et al. Discoidin domain receptor 1-deletion ameliorates fibrosis and promotes adipose tissue beiging, brown fat activity, and increased metabolic rate in a mouse model of cardiometabolic disease. Mol. Metab. 2020, 39, 101006. [Google Scholar] [CrossRef]
  167. Li, X.A.; Bu, X.; Yan, F.; Wang, F.; Wei, D.; Yuan, J.; Zheng, W.; Su, J.; Yuan, J. Deletion of discoidin domain receptor 2 attenuates renal interstitial fibrosis in a murine unilateral ureteral obstruction model. Ren. Fail. 2019, 41, 481–488. [Google Scholar] [CrossRef]
  168. Kim, H.-G.; Tan, L.; Weisberg, E.L.; Liu, F.; Canning, P.; Choi, H.G.; Ezell, S.A.; Wu, H.; Zhao, Z.; Wang, J. Discovery of a potent and selective DDR1 receptor tyrosine kinase inhibitor. ACS Chem. Biol. 2013, 8, 2145–2150. [Google Scholar] [CrossRef]
  169. Liu, L.; Hussain, M.; Luo, J.; Duan, A.; Chen, C.; Tu, Z.; Zhang, J. Synthesis and biological evaluation of novel dasatinib analogues as potent DDR 1 and DDR 2 kinase inhibitors. Chem. Biol. Drug Des. 2017, 89, 420–427. [Google Scholar] [CrossRef]
  170. Murray, C.W.; Berdini, V.; Buck, I.M.; Carr, M.E.; Cleasby, A.; Coyle, J.E.; Curry, J.E.; Day, J.E.; Day, P.J.; Hearn, K. Fragment-based discovery of potent and selective DDR1/2 inhibitors. ACS Med. Chem. Lett. 2015, 6, 798–803. [Google Scholar] [CrossRef]
  171. Gumireddy, K.; Baker, S.J.; Cosenza, S.C.; John, P.; Kang, A.D.; Robell, K.A.; Reddy, M.R.; Reddy, E.P. A non-ATP-competitive inhibitor of BCR-ABL overrides imatinib resistance. Proc. Natl. Acad. Sci. USA 2005, 102, 1992–1997. [Google Scholar] [CrossRef]
  172. Lu, X.; Smaill, J.B.; Ding, K. New promise and opportunities for allosteric kinase inhibitors. Angew. Chem. Int. Ed. 2020, 59, 13764–13776. [Google Scholar] [CrossRef]
  173. Patten, J.; Wang, K. Fibronectin in development and wound healing. Adv. Drug Deliv. Rev. 2021, 170, 353–368. [Google Scholar] [CrossRef]
  174. Shinde, A.V.; Kelsh, R.; Peters, J.H.; Sekiguchi, K.; Van De Water, L.; McKeown-Longo, P.J. The α4β1 integrin and the EDA domain of fibronectin regulate a profibrotic phenotype in dermal fibroblasts. Matrix Biol. 2015, 41, 26–35. [Google Scholar] [CrossRef] [PubMed]
  175. Bhattacharyya, S.; Tamaki, Z.; Wang, W.; Hinchcliff, M.; Hoover, P.; Getsios, S.; White, E.S.; Varga, J. FibronectinEDA promotes chronic cutaneous fibrosis through Toll-like receptor signaling. Sci. Transl. Med. 2014, 6, 232ra250. [Google Scholar] [CrossRef] [PubMed]
  176. Muro, A.F.; Moretti, F.A.; Moore, B.B.; Yan, M.; Atrasz, R.G.; Wilke, C.A.; Flaherty, K.R.; Martinez, F.J.; Tsui, J.L.; Sheppard, D. An essential role for fibronectin extra type III domain A in pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 2008, 177, 638–645. [Google Scholar] [CrossRef] [PubMed]
  177. Klingberg, F.; Chau, G.; Walraven, M.; Boo, S.; Koehler, A.; Chow, M.L.; Olsen, A.L.; Im, M.; Lodyga, M.; Wells, R.G. The fibronectin ED-A domain enhances recruitment of latent TGF-β-binding protein-1 to the fibroblast matrix. J. Cell Sci. 2018, 131, jcs201293. [Google Scholar] [CrossRef] [PubMed]
  178. Zhang, L.; Yan, H.; Tai, Y.; Xue, Y.; Wei, Y.; Wang, K.; Zhao, Q.; Wang, S.; Kong, D.; Midgley, A.C. Design and Evaluation of a Polypeptide that Mimics the Integrin Binding Site for EDA Fibronectin to Block Profibrotic Cell Activity. Int. J. Mol. Sci. 2021, 22, 1575. [Google Scholar] [CrossRef]
  179. Bornstein, P. Matricellular proteins: An overview. J. Cell Commun. Signal. 2009, 3, 163–165. [Google Scholar] [CrossRef]
  180. Pinto, A.R. Matricellular Proteins As Critical Regulators of Fibrosis. Am. Heart Assoc. 2021, 129, 1036–1038. [Google Scholar] [CrossRef]
  181. Feng, D.; Gerarduzzi, C. Emerging roles of matricellular proteins in systemic sclerosis. Int. J. Mol. Sci. 2020, 21, 4776. [Google Scholar] [CrossRef]
  182. Kurundkar, A.R.; Kurundkar, D.; Rangarajan, S.; Locy, M.L.; Zhou, Y.; Liu, R.M.; Zmijewski, J.; Thannickal, V.J. The matricellular protein CCN1 enhances TGF-beta1/SMAD3-dependent profibrotic signaling in fibroblasts and contributes to fibrogenic responses to lung injury. FASEB J. 2016, 30, 2135–2150. [Google Scholar] [CrossRef]
  183. Rotstein, B.; Post, Y.; Reinhardt, M.; Lammers, K.; Buhr, A.; Heinisch, J.J.; Meyer, H.; Paululat, A. Distinct domains in the matricellular protein Lonely heart are crucial for cardiac extracellular matrix formation and heart function in Drosophila. J. Biol. Chem. 2018, 293, 7864–7879. [Google Scholar] [CrossRef]
  184. van Caam, A.; Vonk, M.; van den Hoogen, F.; van Lent, P.; van der Kraan, P. Unraveling SSc pathophysiology; the myofibroblast. Front. Immunol. 2018, 9, 2452. [Google Scholar] [CrossRef]
  185. Effendi, W.I.; Nagano, T. Connective tissue growth factor in idiopathic pulmonary fibrosis: Breaking the bridge. Int. J. Mol. Sci. 2022, 23, 6064. [Google Scholar] [CrossRef]
  186. Richeldi, L.; Fernández Pérez, E.R.; Costabel, U.; Albera, C.; Lederer, D.J.; Flaherty, K.R.; Ettinger, N.; Perez, R.; Scholand, M.B.; Goldin, J.; et al. Pamrevlumab, an anti-connective tissue growth factor therapy, for idiopathic pulmonary fibrosis (PRAISE): A phase 2, randomised, double-blind, placebo-controlled trial. Lancet Respir. Med. 2020, 8, 25–33. [Google Scholar] [CrossRef]
  187. Barbe, M.F.; Hilliard, B.A.; Amin, M.; Harris, M.Y.; Hobson, L.J.; Cruz, G.E.; Popoff, S.N. Blocking CTGF/CCN2 reduces established skeletal muscle fibrosis in a rat model of overuse injury. FASEB J. 2020, 34, 6554. [Google Scholar] [CrossRef]
  188. Vainio, L.E.; Szabó, Z.; Lin, R.; Ulvila, J.; Yrjölä, R.; Alakoski, T.; Piuhola, J.; Koch, W.J.; Ruskoaho, H.; Fouse, S.D. Connective tissue growth factor inhibition enhances cardiac repair and limits fibrosis after myocardial infarction. JACC Basic Transl. Sci. 2019, 4, 83–94. [Google Scholar] [CrossRef]
  189. Mohamed, I.A.; Gadeau, A.-P.; Fliegel, L.; Lopaschuk, G.; Mlih, M.; Abdulrahman, N.; Fillmore, N.; Mraiche, F. Na+/H+ exchanger isoform 1-induced osteopontin expression facilitates cardiomyocyte hypertrophy. PLoS ONE 2015, 10, e0123318. [Google Scholar] [CrossRef]
  190. Singh, M.; Foster, C.R.; Dalal, S.; Singh, K. Osteopontin: Role in extracellular matrix deposition and myocardial remodeling post-MI. J. Mol. Cell. Cardiol. 2010, 48, 538–543. [Google Scholar] [CrossRef]
  191. Lenga, Y.; Koh, A.; Perera, A.S.; McCulloch, C.A.; Sodek, J.; Zohar, R. Osteopontin expression is required for myofibroblast differentiation. Circ. Res. 2008, 102, 319–327. [Google Scholar] [CrossRef]
  192. Cabiati, M.; Svezia, B.; Matteucci, M.; Botta, L.; Pucci, A.; Rinaldi, M.; Caselli, C.; Lionetti, V.; Del Ry, S. Myocardial expression analysis of osteopontin and its splice variants in patients affected by end-stage idiopathic or ischemic dilated cardiomyopathy. PLoS ONE 2016, 11, e0160110. [Google Scholar] [CrossRef]
  193. Lopez, B.; Gonzalez, A.; Lindner, D.; Westermann, D.; Ravassa, S.; Beaumont, J.; Gallego, I.; Zudaire, A.; Brugnolaro, C.; Querejeta, R. Osteopontin-mediated myocardial fibrosis in heart failure: A role for lysyl oxidase? Cardiovasc. Res. 2013, 99, 111–120. [Google Scholar] [CrossRef]
  194. Frangogiannis, N.G. Matricellular proteins in cardiac adaptation and disease. Physiol. Rev. 2012, 92, 635–688. [Google Scholar] [CrossRef] [PubMed]
  195. Mamazhakypov, A.; Sartmyrzaeva, M.; Sarybaev, A.S.; Schermuly, R.; Sydykov, A. Clinical and Molecular Implications of Osteopontin in Heart Failure. Curr. Issues Mol. Biol. 2022, 44, 3573–3597. [Google Scholar] [CrossRef] [PubMed]
  196. Shirakawa, K.; Sano, M. Osteopontin in Cardiovascular Diseases. Biomolecules 2021, 11, 1047. [Google Scholar] [CrossRef] [PubMed]
  197. Collins, A.R.; Schnee, J.; Wang, W.; Kim, S.; Fishbein, M.C.; Bruemmer, D.; Law, R.E.; Nicholas, S.; Ross, R.S.; Hsueh, W.A. Osteopontin modulates angiotensin II-induced fibrosis in the intact murine heart. J. Am. Coll. Cardiol. 2004, 43, 1698–1705. [Google Scholar] [CrossRef] [PubMed]
  198. Sam, F.; Xie, Z.; Ooi, H.; Kerstetter, D.L.; Colucci, W.S.; Singh, M.; Singh, K. Mice lacking osteopontin exhibit increased left ventricular dilation and reduced fibrosis after aldosterone infusion. Am. J. Hypertens. 2004, 17, 188–193. [Google Scholar] [CrossRef]
  199. Xie, Z.; Singh, M.; Singh, K. Osteopontin modulates myocardial hypertrophy in response to chronic pressure overload in mice. Hypertension 2004, 44, 826–831. [Google Scholar] [CrossRef]
  200. Dai, J.; Matsui, T.; Abel, E.D.; Dedhar, S.; Gerszten, R.E.; Seidman, C.E.; Seidman, J.; Rosenzweig, A. DSAGE identifies osteopontin as a downstream effector of integrin-linked kinase (ILK) in cardiac-specific ILK knockout mice. Circ. Heart Fail. 2014, 7, 184. [Google Scholar] [CrossRef]
  201. Brigstock, D.R. Extracellular Vesicles in Organ Fibrosis: Mechanisms, Therapies, and Diagnostics. Cells 2021, 10, 1596. [Google Scholar] [CrossRef]
  202. Rilla, K.; Mustonen, A.-M.; Arasu, U.T.; Härkönen, K.; Matilainen, J.; Nieminen, P. Extracellular vesicles are integral and functional components of the extracellular matrix. Matrix Biol. 2019, 75, 201–219. [Google Scholar] [CrossRef]
  203. Lenzini, S.; Bargi, R.; Chung, G.; Shin, J.-W. Matrix mechanics and water permeation regulate extracellular vesicle transport. Nat. Nanotechnol. 2020, 15, 217–223. [Google Scholar] [CrossRef]
  204. Martin-Medina, A.; Lehmann, M.; Burgy, O.; Hermann, S.; Baarsma, H.A.; Wagner, D.E.; De Santis, M.M.; Ciolek, F.; Hofer, T.P.; Frankenberger, M. Increased extracellular vesicles mediate WNT5A signaling in idiopathic pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 2018, 198, 1527–1538. [Google Scholar] [CrossRef]
  205. Liu, X.; Miao, J.; Wang, C.; Zhou, S.; Chen, S.; Ren, Q.; Hong, X.; Wang, Y.; Hou, F.F.; Zhou, L. Tubule-derived exosomes play a central role in fibroblast activation and kidney fibrosis. Kidney Int. 2020, 97, 1181–1195. [Google Scholar] [CrossRef]
  206. Yang, J.; Yu, X.; Xue, F.; Li, Y.; Liu, W.; Zhang, S. Exosomes derived from cardiomyocytes promote cardiac fibrosis via myocyte-fibroblast cross-talk. Am. J. Transl. Res. 2018, 10, 4350. [Google Scholar]
  207. Charrier, A.; Chen, R.; Chen, L.; Kemper, S.; Hattori, T.; Takigawa, M.; Brigstock, D.R. Connective tissue growth factor (CCN2) and microRNA-21 are components of a positive feedback loop in pancreatic stellate cells (PSC) during chronic pancreatitis and are exported in PSC-derived exosomes. J. Cell Commun. Signal. 2014, 8, 147–156. [Google Scholar] [CrossRef]
  208. Iversen, L.V.; Ullman, S.; Østergaard, O.; Nielsen, C.T.; Halberg, P.; Karlsmark, T.; Heegaard, N.H.; Jacobsen, S. Cross-sectional study of soluble selectins, fractions of circulating microparticles and their relationship to lung and skin involvement in systemic sclerosis. BMC Musculoskelet. Disord. 2015, 16, 191. [Google Scholar] [CrossRef]
  209. Mansouri, N.; Willis, G.R.; Fernandez-Gonzalez, A.; Reis, M.; Nassiri, S.; Mitsialis, S.A.; Kourembanas, S. Mesenchymal stromal cell exosomes prevent and revert experimental pulmonary fibrosis through modulation of monocyte phenotypes. JCI Insight 2019, 4, e128060. [Google Scholar] [CrossRef]
  210. Willis, G.R.; Fernandez-Gonzalez, A.; Anastas, J.; Vitali, S.H.; Liu, X.; Ericsson, M.; Kwong, A.; Mitsialis, S.A.; Kourembanas, S. Mesenchymal stromal cell exosomes ameliorate experimental bronchopulmonary dysplasia and restore lung function through macrophage immunomodulation. Am. J. Respir. Crit. Care Med. 2018, 197, 104–116. [Google Scholar] [CrossRef]
  211. Royce, S.G.; Patel, K.P.; Mao, W.; Zhu, D.; Lim, R.; Samuel, C.S. Serelaxin enhances the therapeutic effects of human amnion epithelial cell-derived exosomes in experimental models of lung disease. Br. J. Pharmacol. 2019, 176, 2195–2208. [Google Scholar] [CrossRef]
  212. Tan, J.L.; Lau, S.N.; Leaw, B.; Nguyen, H.P.; Salamonsen, L.A.; Saad, M.I.; Chan, S.T.; Zhu, D.; Krause, M.; Kim, C. Amnion epithelial cell-derived exosomes restrict lung injury and enhance endogenous lung repair. Stem Cells Transl. Med. 2018, 7, 180–196. [Google Scholar] [CrossRef]
  213. Guiot, J.; Cambier, M.; Boeckx, A.; Henket, M.; Nivelles, O.; Gester, F.; Louis, E.; Malaise, M.; Dequiedt, F.; Louis, R. Macrophage-derived exosomes attenuate fibrosis in airway epithelial cells through delivery of antifibrotic miR-142-3p. Thorax 2020, 75, 870–881. [Google Scholar] [CrossRef]
  214. Wang, J.; Lee, C.J.; Deci, M.B.; Jasiewicz, N.; Verma, A.; Canty, J.M.; Nguyen, J. MiR-101a loaded extracellular nanovesicles as bioactive carriers for cardiac repair. Nanomedicine 2020, 27, 102201. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Crucial steps in fibroblast activation. A blood vessel (Bv), neutrophils (Ne), macrophages (Mc), and mast cells (Ma) are indicated.
Figure 1. Crucial steps in fibroblast activation. A blood vessel (Bv), neutrophils (Ne), macrophages (Mc), and mast cells (Ma) are indicated.
Biomolecules 13 00758 g001
Figure 2. A schematic showing activation of mast cells (Ma) and macrophages (Mc) by forces generated by myofibroblasts (Mf) embedded in the ECM and acting via integrins. Collagen fibrils (CF) and proteoglycans (PG) are indicated.
Figure 2. A schematic showing activation of mast cells (Ma) and macrophages (Mc) by forces generated by myofibroblasts (Mf) embedded in the ECM and acting via integrins. Collagen fibrils (CF) and proteoglycans (PG) are indicated.
Biomolecules 13 00758 g002
Figure 3. A schematic depicting crucial elements of mechanisms that contribute to the formation of collagen-rich, stiff matrices that promote excessive scarring. The STOP signs indicate the elements whose blocking is associated with anti-fibrotic effects. (A) A fibroblastic cell that includes discoidin domain receptors (DDR) and integrins (INT). Matricellular proteins (MCPs) and extracellular vesicles (Ev) are also indicated. (B) Intracellular biosynthesis of procollagen molecules formed by trimerization of individual procollagen chains. Crucial collagen-modifying enzymes include prolyl 4-hydroxylase (P4H), lysyl hydroxylase (LH), and glycosylases. Moreover, heat-shock protein 47 (HSP47) and protein disulfide isomerase (PDI) exemplify protein chaperones participating in procollagen formation. (C) Extracellular processing of procollagen propeptides with procollagen N proteinase (PNP) and procollagen C proteinase (PCP), whose activity is accelerated by PCP enhancer (PCPE). The N-terminal (Np) and the C-terminal (Cp) propeptides are also indicated. The N-terminal (Nt) and the C-terminal (Ct) telopeptides are also indicated. (D) Site-specific self-assembly of collagen molecules into a fibril; the assembly is driven by the interaction of collagen telopeptides with an interacting partner’s telopeptide-binding region (TBR). (E) A depiction of a collagen microfibril, in which collagen molecules undergo cross-linking catalyzed by lysyl oxidases (LOX) and transglutaminases (TG). (F) A mature collagen fibril associated with other structural macromolecules, e.g., proteoglycans (PG). (G) An example of a fibrotic site that affects an injured organ. In addition to muscle cells (M), a magnified insert shows fibroblasts (F) embedded in collagen-rich fibrotic tissue.
Figure 3. A schematic depicting crucial elements of mechanisms that contribute to the formation of collagen-rich, stiff matrices that promote excessive scarring. The STOP signs indicate the elements whose blocking is associated with anti-fibrotic effects. (A) A fibroblastic cell that includes discoidin domain receptors (DDR) and integrins (INT). Matricellular proteins (MCPs) and extracellular vesicles (Ev) are also indicated. (B) Intracellular biosynthesis of procollagen molecules formed by trimerization of individual procollagen chains. Crucial collagen-modifying enzymes include prolyl 4-hydroxylase (P4H), lysyl hydroxylase (LH), and glycosylases. Moreover, heat-shock protein 47 (HSP47) and protein disulfide isomerase (PDI) exemplify protein chaperones participating in procollagen formation. (C) Extracellular processing of procollagen propeptides with procollagen N proteinase (PNP) and procollagen C proteinase (PCP), whose activity is accelerated by PCP enhancer (PCPE). The N-terminal (Np) and the C-terminal (Cp) propeptides are also indicated. The N-terminal (Nt) and the C-terminal (Ct) telopeptides are also indicated. (D) Site-specific self-assembly of collagen molecules into a fibril; the assembly is driven by the interaction of collagen telopeptides with an interacting partner’s telopeptide-binding region (TBR). (E) A depiction of a collagen microfibril, in which collagen molecules undergo cross-linking catalyzed by lysyl oxidases (LOX) and transglutaminases (TG). (F) A mature collagen fibril associated with other structural macromolecules, e.g., proteoglycans (PG). (G) An example of a fibrotic site that affects an injured organ. In addition to muscle cells (M), a magnified insert shows fibroblasts (F) embedded in collagen-rich fibrotic tissue.
Biomolecules 13 00758 g003
Figure 4. A diagram showing crucial players involved in mechanical signal transduction in fibrotic scars. Collagen fibril-activated and phosphorylated discoidin domain receptor 1 (DDR1), in association with myosin II, is indicated. A matrix-bound integrin with its cytoplasmic interactants, comprising talin, vinculin, paxillin, focal adhesion kinase (FAK), and integrin-linked kinase (ILK), is also presented. In addition, a TGFBR bound to its TGF-β ligand is demonstrated. The diagram also illustrates a nuclear location of yes-associated protein (YAP), transcriptional co-activator with PDZ-binding motif (TAZ), and myocardin-related transcription factor (MRTF).
Figure 4. A diagram showing crucial players involved in mechanical signal transduction in fibrotic scars. Collagen fibril-activated and phosphorylated discoidin domain receptor 1 (DDR1), in association with myosin II, is indicated. A matrix-bound integrin with its cytoplasmic interactants, comprising talin, vinculin, paxillin, focal adhesion kinase (FAK), and integrin-linked kinase (ILK), is also presented. In addition, a TGFBR bound to its TGF-β ligand is demonstrated. The diagram also illustrates a nuclear location of yes-associated protein (YAP), transcriptional co-activator with PDZ-binding motif (TAZ), and myocardin-related transcription factor (MRTF).
Biomolecules 13 00758 g004
Figure 5. A depiction of TGF-β1 activation by the mechanical forces (arrow) generated by myofibroblasts (Mf) embedded in soft ECM (A) or stiff ECM (B). TGF-β1 and latency-associated protein (LAP) are indicated.
Figure 5. A depiction of TGF-β1 activation by the mechanical forces (arrow) generated by myofibroblasts (Mf) embedded in soft ECM (A) or stiff ECM (B). TGF-β1 and latency-associated protein (LAP) are indicated.
Biomolecules 13 00758 g005
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

Fertala, J.; Wang, M.L.; Rivlin, M.; Beredjiklian, P.K.; Abboud, J.; Arnold, W.V.; Fertala, A. Extracellular Targets to Reduce Excessive Scarring in Response to Tissue Injury. Biomolecules 2023, 13, 758. https://doi.org/10.3390/biom13050758

AMA Style

Fertala J, Wang ML, Rivlin M, Beredjiklian PK, Abboud J, Arnold WV, Fertala A. Extracellular Targets to Reduce Excessive Scarring in Response to Tissue Injury. Biomolecules. 2023; 13(5):758. https://doi.org/10.3390/biom13050758

Chicago/Turabian Style

Fertala, Jolanta, Mark L. Wang, Michael Rivlin, Pedro K. Beredjiklian, Joseph Abboud, William V. Arnold, and Andrzej Fertala. 2023. "Extracellular Targets to Reduce Excessive Scarring in Response to Tissue Injury" Biomolecules 13, no. 5: 758. https://doi.org/10.3390/biom13050758

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