Skip to main content

Collagenases and their inhibitors: a review

Abstract

Hide and skin are complex tissue where the most abundant component is collagen. Matrix metalloproteinases and bacterial collagenases are two kinds of collagenases that can cleave the triple-helical domain of native fibrillar collagens. In this paper, the family members and domain composition of matrix metalloproteinases and bacterial collagenases are summarized. The catalytic mechanism of collagen hydrolysis by collagenases is described, and the methods adopted to date for investigating and regulating collagenases and their inhibitors are reviewed. Furthermore, the applications of collagenases and their inhibitors in biomedicine, food processing and the enzymatic unhairing process in the leather-making industry are presented.

Graphical Abstract

1 Introduction

Collagens are important components of the extracellular matrix (ECM) in mammalian tissues such as skin, ligaments and tendons [1]. To date, 40 vertebrate collagen genes have been identified. These genes form at least 28 different collagen molecules [2,3,4], fibrillar or non-fibrillar molecules, which are sequentially numbered using Roman numerals (I–XXVIII). Fibrillar collagens have high thermal stability and strong mechanical strength [5,6,7,8,9].

The hide and skin of mammalians consist of various type of collagens, such as types I, III, IV and VII, among which type I and type III collagens are the major types of fibrillar collagen [10, 11]. The most abundant collagen is type I collagen, whose molecule consists of three α polypeptides (two α1(I)-chains and one α2(I)-chain) that are wound around each other to form a triple helix, and at either end of the triple helix are non-helical moieties, called telopeptides (Fig. 1) [12,13,14]. Native type I collagen is highly resistant to common proteases, such as trypsin and chymotrypsin, while enzymes with collagenolytic activity, such as matrix metalloproteinase-1 (MMP-1; also known as interstitial collagenase), show the ability to cleave type I collagen at the triple-helical domain [15].

Fig. 1
figure 1

Triple-helical structures of type I and type III collagens. There are three α-chains in each collagen molecule. Type I collagen includes two α1(I)-chains and one α2(I)-chain and type III collagen includes three α1(III)-chains. The triple-helical domain can be degraded by collagenases into peptides

The most widely studied collagenases are matrix metalloproteinases (MMPs) from animals and bacterial collagenases from micro-organisms. MMPs are a family of multi-domain proteolytic enzymes containing zinc ions [16]. They have the potential to decompose polypeptides in the ECM and play an important role in physiological and pathological processes [17,18,19,20,21]. Bacterial collagenases are mainly secreted by micro-organisms such as Clostridium and Vibrio. In the Merops peptidase database, bacterial collagenases and MMPs belong to the M9 family and the M10 family, respectively.

Unhairing is an essential process in leather manufacturing. However, the conventional hair-burning unhairing process will cause seriously pollution problem. As a reliable alternative to the conventional lime-sulfide process, unhairing by proteases (enzymatic unhairing) in the leather-making process has been employed over decades [22, 23]. A wide range of enzymes have been investigated for their potential applications in enzymatic unhairing, such as alkaline proteinases from Bacillus [24], keratinases from Actinomadura keratinilytica [25], and serine proteases from Caldicoprobacter algeriensis [26]. It has been noted that the main risk of grain damage during enzymatic unhairing is due to the over-hydrolysis of collagens by collagenases. To fully exploit the potential of enzymatic unhairing, it is crucial to understand the action and effect of components with collagenase activity on different collagen types in the hide and skin. Herein, the family classification and structural characteristics of collagenases (MMPs and bacterial collagenases) are summarized. The catalytic mechanism of collagen hydrolysis by collagenases is described. Furthermore, the methods so far employed for investigating and regulating collagenases and their inhibitors are reviewed.

2 Progress of research on collagenases

Proteases are a group of hydrolytic enzymes that can cleave the peptide bonds of protein molecules and degrade them into small peptides and amino acids. Collagenases are proteases possessing the ability to degrade various types of collagens at the specific site [24] and are predominantly found in animal and micro-organisms, which differ in substrate specificity [21].

2.1 MMPs

2.1.1 Classification of MMPs

MMPs are a family of proteolytic enzymes that have different substrate preferences, while share the same zinc-dependent active site and have similar structural features [27]. They are grouped into collagenases, gelatinases, stromelysins, matrilysins and membrane-type (MT)MMPs and others, which are characterized by their domain organization and substrate preference (Fig. 2 and Table 1) [28,29,30,31,32,33].

Fig. 2
figure 2

Schematic representation of the various domains of different MMPs

Table 1 Classification of MMP family with their domain organization and substrates preference

Collagenases cleave ECM proteins and other soluble proteins, however, the most important role of this type of MMPs is that they can digest fibrillar collagens of types I, II, III, IV and XI into characteristic 3/4 and 1/4 fragments [34,35,36]. In addition to MMP-1, MMP-8, MMP-13 and MMP-18, other specific MMPs may also cleave fibrillar collagens such as MMP-2 and MMP-14. They can digest type I, II and III collagens in a manner similar to that of collagenases but are divided into other subgroups due to their domain organization [37].

2.1.2 Domains of MMPs

Although different subgroups of MMPs have different domain organizations, several important domains are common to almost all MMPs. They are (from N-terminal to C-terminal positions) a propeptide (Prop), a catalytic domain (Cat), a linker of variable length, and a hemopexin-like domain (Hpx), as shown in Fig. 3A [38].

Fig. 3
figure 3

Diagram of the structure of MMP-1 and the sequence of MMP-1 between different organism species. A Ribbon diagram of proMMP-1 (human) and active MMP-1 (pig). From N-terminal to the C-terminalpositions: a propeptide, a catalytic domain, a linking region and a hemopexin-like domain. The catalytic zinc(II) ion is represented by a blue sphere. B Ribbon diagram of active site of MMP-1. The propeptide is shown in red; the catalytic domain, in purple; the linker region, in yellow; the hemopexin domain, in lime; zinc ions, in gray; calcium ions, in green; the sodium ion, in cyan; and the chloride ion, in orange. The propeptide is removed when the MMPs are activated. The images were prepared with Bookhaven Protein Data bank entries 1SU3 (proMMP-1) and 1FBL(MMP-1) using Pymol. C. The sequence of MMP-1 between different organism species. The conserved sequences are marked in red

2.1.3 Propeptide

The propeptide consists of about 80 amino acid residues forming three α-helices [39]. The α(3)-helices are followed by a “cysteine switch”, a very conserved region (PRCGXPD) [40, 41], where the sulfhydryl group is coordinated with the catalytic Zn(II) ion, forming a tetrahedral coordination sphere. This keeps the MMPs in a latent state until the cysteine-Zn2+ interaction is disrupted; then, the Zn(II) ion can combine with water molecules which are necessary for the hydrolysis of polypeptides.

2.1.4 Catalytic domain

The catalytic domain consists of about 160 amino acid residues, including a five-stranded β-sheet (four parallel(β2–β1–β3–β5) and one anti-parallel(β4) components), three α-helices (α1, α2 and α3) (Fig. 3A), two Zn(II) ions (one catalytic ion and one structural ion, about 80–100 nm apart), and at least one Ca(II) ion (usually three) [28, 36]. These structures are assembled into a sphere with a diameter of about 400 nm [39,40,41]. The catalytic Zn(II) ion is essential for the catalytic process of MMPs. It coordinates with three the His residues in the HEXGHXXGXXH sequence (Fig. 3C) [34, 41,42,43,44], which is conserved among all the MMPs. In the activated free enzyme, the catalytic Zn(II) ion will also be coordinated by a water molecule hydrogen-bonded to the Glu residue flanking the active site [32]. Moreover, the catalytic Zn(II) ion flanked by a hydrophobic cavity (S1′ pocket) contains mainly hydrophobic residues and is comprised of a Ω-loop, which includes the loop linking helices α2 and α3. The loop holds a hydrophobic Met residue which is part of the highly conserved 1,4-β-turn referred to as the “Met-turn” (Fig. 3B).The structural Zn(II) ion and the Ca(II) ions play crucial roles in maintaining the conformation of proteins [43].

2.1.5 Linker region

The catalytic domain and hemopexin-like domain are connected by a proline-rich linker region (also called hinge region) (Fig. 3A). This region is flexible to a certain extent due to its variable length, which gives inter-domain flexibility to the structure of MMPs, retains enzyme stability, and participates in the hydrolysis of some complex substrates [45]. Therefore, the linker region is important for MMPs to express collagenolytic activity.

2.1.6 Hemopexin-like domain

The hemopexin-like domain has the shape of an ellipsoidal disc (Fig. 3A). It consists of approximately 210 amino acid residues and contains a four-bladed β-propeller structure with a single disulfide bond between the 1st and 4th blade, [46]. Each blade is composed of four antiparallel β-strands and has a Cys residue at each end. A calcium ion and a chloride ion are usually present in the center of the four-bladed β-propeller [47, 48]. Although it has been found that the catalytic domain is sufficient for degradation of non-collagenous substrates, the hemopexin-like domain is inevitable for recognition and hydrolyzation of fibrillar collagen [49,50,51].

2.1.7 Catalytic mechanism of MMPs with collagen-hydrolyzing activity

The catalytic action of MMPs on collagen has been explored for decades. To understand the catalytic mechanism, triple-helical peptides (THPs) have been used as models of native collagens [51,52,53]. Moreover, computer simulation using molecular dynamics has been applied for exploring the interaction between MMPs and collagens [54,55,56,57,58]. Both approaches can help us to better discover the interaction mode between collagenases and their substrates.

MMP-1 serves as the prototype for all interstitial collagenases. MMP-1(activated form) cleaves type I collagen molecules at a site three-quarters from the N-terminus, after which collagen becomes a poor substrate of collagenase due to denaturation, and it continues to be degraded by gelatinase or common cathepsin [59].

2.1.8 Interdomain communications modulate the activity of MMP-1

The effective hydrolysis of collagens by MMP-1 requires synergy between the substrates and multiple domains. Although the catalytic domain is essential for substrate hydrolysis, it alone can only degrade non-collagen substrate (gelatin or heat-denatured collagen). The binding process between enzyme and substrate is initiated by the hemopexin-like domain, which can interact with triple-helical substrates with residues at specific positions [60,61,62,63,64,65,66], including residues Ile290, Arg291, Phe301, Val319, and Asp338. Moreover, through the allosteric communication between the catalytic domain and the hemopexin-like domain, the catalytic domain is located at the position of the scissile bond, thus initiating the degradation of collagen [53, 56, 67].

2.1.9 Collagen hydrolysis mechanism induced by catalytic domain

It is suggested with the three-dimensional structure of MMP-1 that its substrate binding site is too small to accommodate the triple-helical structure of collagen. When the MMP locates the scissile bond between collagen and active site with interdomain flexibility, the catalytic domain begins to hydrolyze the substrate. However, it was found that the closely intertwined triple-helical structure of collagen (about 150 nm in diameter) could not match the gap in the active site of the catalytic domain of MMP-1 (about 50 nm in diameter). Chung et al. [68] found that it is necessary to locally unwind the collagen triple-helical structure before the enzyme cleaves collagen. After unwinding, the single-strand collagen polypeptide chain can be surrounded by the catalytic domain of MMP-1; thus, hydrolysis occurs. During this process, the conformation changes for both collagen and MMP-1.

MMP-1 requires the catalytic Zn(II) ion, three His residues, a Glu residue and a water molecule to exhibit its catalytic ability. Futhermore, there is a Met residue as a hydrophobic base, supporting the structure around the catalytic Zn(II) ion [69]. The breaking process of the scissile bond is shown in Fig. 4. When the scissile bond of the substrate is located at the active site, the catalytic Zn(II) ion carries out penta-coordination with the carbonyl oxygen atom of the scissile bond, an oxygen atom from Glu-bound water and three His residues. At the beginning of hydrolysis, the water molecule polarized by the Glu residue carries out nucleophilic attacks on the carbonyl carbon of the scissile bond [45]. Subsequently, protons are transferred to amino nitrogen atoms with Glu residues, promoting the formation of an intermediate with tetrahedral geometry [35, 70], resulting in the cleavage of the scissile bond. The whole catalytic process can be divided into four steps: (a) the substrate enters the active site and the water molecules between the Zn(II) ion and the Glu residue are polarized; (b) the polarized water molecules carry out nucleophilic attacks on the carbonyl group of the scissile bond to form a tetrahedral structure (which is unstable); (c) the tetrahedral structure collapses, and the scissile bond breaks; (d) the substrate broken by the scissile bond leaves the active site, and the MMP begins to hydrolyze the next substrate.

Fig. 4
figure 4

Catalytic mechanism of hydrolysis of scissile bond of collagen at the active site of the MMP. A The substrate enters the active site, and the Zn(II) ion becomes penta-coordinated. B The glutamate-bound water molecule is polarized and attacks the carbonyl carbon of the scissile bond to form an unstable tetrahedral intermediate. C The structure of the intermediate collapse and the scissile bond breaks. D The cleaved substrate leaves the active site and the MMP starts the next cycle. The catalytic Zn(II) ion is represented by a blue sphere

2.2 Bacterial collagenases

2.2.1 Members of bacterial collagenase

Collagenases produced by Clostridium histolyticum are the most studied bacterial collagenases, which can be categorized into class I (encoded by gene ColG) and class II (encoded by gene ColH) [71]. Class I collagenases show high activity with collagen and moderate activity with FALGPA (a synthetic peptide used for studying the kinetics of collagenases), while class II collagenases are highly active with synthetic peptides and moderately active with collagen [72]. These two types of collagenases belong to the M9B subfamily in the Merops peptidase database. Proteases produced by Vibrio are divided into three classes (I, II and III) characterized by their function and structure. Proteases in class I possess no collagenolytic activity, while classes II and III members are able to degrade collagenous substrates.

2.2.2 Domain of bacterial collagenase

The main literature source on bacterial collagenase is the extensive research on the enzymes produced by Clostridium histolyticum. Collagenases produced by Clostridium are multi-domain zinc metalloproteinases, possessing a signal peptide, the N-terminal collagenase unit and the C-terminal recruitment domain [17], as shown in Fig. 5. The N-terminal collagenase unit contains an activator domain (AD) and a peptidase domain (PD). The C-terminal recruitment domain usually contains one or two collagen-binding domains (CBDs) and one or two polycystic kidney disease-like domains (PKDs).

Fig. 5
figure 5

Diagram of structure of bacteria collagenases. A Schematic diagram of the structure of two kinds of collagenase from Clostridium histolyticum. Class I collagenase(left) and class II collagenase(right). The catalytic Zn(II) ion is represented by a blue sphere; B Ribbon diagram of collagenase unit, collagen-binding domain (orange) and polycystic kidney disease-like domain (cyan) of ColG. The collagen unit contains an activator domain (pink) and a peptidase domain (yellow). The zinc ion is shown in gray; calcium ions, in green; and the chloride ion, in violet. The images were prepared with Bookhaven Protein Data bank entries 4ARE (collagenase unit), 2O8O (CBD) and 4TN9 (PKD) using Pymol

2.2.3 Collagenase unit

The molecular weight of a collagenase unit is about 78 kDa. The activator domain and peptidase domain show a saddle-shaped architecture [17]. There is a conserved Zn(II) ion in the peptidase domain, which coordinates with two His residues in a conserved motif (-HEXXH), a Glu residue downstream of the motif and a water molecule [73, 74]. In addition to the Zn(II) ion, there is also a Ca(II) ion in the active site. These two ions are indispensable for enzyme activity [75].

2.2.4 Recruitment domains

The molecular weight of both the CBD and PKD is about 10 kDa. Based on the studies of the CBD of two kinds of Clostridium collagenase, it was found that the CBD can recognize and bind to the triple-helical structure of collagen, which is necessary for the hydrolysis of insoluble collagen fibers by Clostridium collagenase, and the binding can be enhanced in the presence of calcium ions [76, 77]. The PKD, which consists of 80–90 amino acid residues, is the alignment platform between the collagenase unit and the CBD, but its specific role in collagen hydrolysis is still unclear [78]. It was found that the PKD of some special collagenases not only can bind to collagen but can also swell microfibrils to improve the collagenolytic efficiency [79]. The Ca(II) ions in the PKD can stabilize the conformation of the domain, thus stabilizing the overall stability of collagenase [80].

2.2.5 Catalytic mechanism of bacteria collagenase

MMPs show strict substrate specificity for collagen hydrolysis, cleaving collagen only at well-defined recognition sites. However, Clostridium collagenases have no obvious preference for any type of collagen and can hydrolyze collagen in multiple locations and completely decompose it into small peptides [81].

The information regarding the structure and hydrolysis mechanism of bacterial collagenases is scarce, partly owing to their multi-domain organization. Attempts have been made to sovle the three-dimensional structure of the ColG collagenase from Clostridium histolyticum [72]. In ColG, the recruitment domain locates and anchors the collagen by specifically recognizing their triple-helical conformation, while the collagenase unit hydrolyzes the prepared collagen. Similar to MMPs, collagen changes its configuration before being hydrolyzed by bacterial collagenases. Eckhard called this a “chew-and-digest mechanism” [74]. The hydrolysis process of bacterial collagenases can be divided into two steps: At the beginning of binding with collagen, the collagenase gradually changes its configuration from open to closed; then, the collagenase changes its configuration to semi-open. Meanwhile, the triple helices of collagen are unlocked, and the scissile bond is exposed to the active site to complete hydrolysis. Once this part of collagen is hydrolyzed, the collagenase returns to the open configuration so that another collagen molecule can be degraded.

2.3 Applications of collagenases

The structure and catalytic mechanism of collagenases have been clearly identified, therefore, a wide range for applications of collagenases have been investigated in fields such as biomedicine, food processing and leather manufacturing. The following section focuses on collagenases applications.

2.3.1 Biomedical field

Collagenases can be directly applied to clinical treatment [82,83,84]. Collagens are the most abundant proteins in the ECM and have numerous functions. However, the anomalous expression or degradation of collagens might cause some diseases.

One of the applications of collagenases is related to the enzymatic debridement of wounds and other injuries [85]. Debridement is applied to the treatment of chronic wounds, which is stalled in the inflammatory phase [86]. The use of collagenases in wound debridement is considered as a safe and effective choice [87, 88]. Karagol et al. [89] reported a successful case of enzymatic debridement, showing that collagenases can be used not only for the removal of eschar but also for the avoidance of the progression of necrotic tissue. The Clostridium collagenase ability to minimize burn progression was evaluated using the classic burn comb model in pigs. It was found that collagenase treatment had multiple effects including providing an early and improved inflammatory response and preventing the destruction of dermal collagen [90]. Recently, Francesco and coworkers performed enzymatic debridement on 70 patients with chronic wounds of different etiologies using ointment (Bionect Start®) based on hyaluronic acid and collagenase. Most of the patients (62 of 70) healed completely within 8 weeks [91]. It has been confirmed that with the aid of collagenase, the complications of surgery can be partly avoided, and the progress and enlargement of necrotic tissue can be limited [92].

Another application of collagenases is related to the treatment of Dupuytren’s disease (DD), a disorder caused by the immoderate deposit of collagen (mainly types I and III) [93]. Before the production of collagenase, the only treatment for DD was surgery [94, 95]. In the early 1990s, Badalamente et al. began to investigate therapeutic treatment for DD with collagenases, and they finally obtained the approval of the Food and Drug Administration (FDA) in 2010 [96, 97]. Reports have indicated that collagenase treatment is effective and simple and promotes quick recovery in comparison with surgical treatment [98]. However, collagenase treatment has shown a certain recurrence rate and might be accompanied by some complications. Nayar and coworkers investigated 34 patients for the rates of contracture resolution and recurrence who went through collagenase treatment. It was observed that 42 of 44 metacarpophalangeal joints and 14 of 33 proximal interphalangeal joints performed immediate contracture resolution (improving from 50° to 1.5° with p < 0.001 and from 44° to 16° with p = 0.182, respectively). Moreover, they re-evaluated the patients at different time (up to 2 years) and found that reoccurrence may occur within 2 years after treatment [99]. Coleman et al. summarized the efficiency of collagenase treatment and adverse events on post-injection day 30. The results indicated that complications following collagenase treatment are common, among which the most (> 75% of patients) were contusion, local peripheral edema, and pain in the extremities. But the severity of most adverse events is mild to moderate [100]. Nonetheless, studies of patient-reported outcome measures suggested that more than 85% of patients were satisfied following collagenase treatment for DD [101].

Collagenases are applied in genetic therapy, which has been viewed as a promising approach to treat cancers [102]. Cemazar and coworkers treated the tumors with enzymes and evaluated the transfection efficiency in 3, 9, and 15 days post-transfection. It was shown that a tenfold increase in the percentage of transfected area of GFP and a tenfold increase in functional luciferase within the tumor occurred after treatment with both collagenase and hyaluronidase, which suggested that the administration of two enzymes active against collagen and hyaluronan may be required to substantially increase gene titers within the tumor [103]. Increased interstitial fluid pressure (IFP) in tumors is an obstacle to the accumulation of systemically delivered nanocarriers [104]. The intravenous injection of MMP-1 was utilized to investigate whether this collagenase could reduce IFP in tumors. The results showed that IFP and the amount of collagen in the tumor are significantly reduced by MMP-1 at 1 h post-injection, supporting the potential use of collagenase has the ability to improve systemic gene delivery into tumors [105]. The applications of collagenases in the biomedical field are shown in Table 2.

Table 2 Applications of collagenases in the biomedical field

2.3.2 Food processing

Collagenases have been applied to meat tenderization. The toughness of meats is partly due to the presence of collagen. Accordingly, its digestion may result in meat tenderness. The tenderization effect and the action of a cold-adapted collagenolytic enzyme MCP-01 on beef meat were investigated. The shear force of meat was reduced by 23%, and the relative myofibrillar fragmentation index of meat was increased by 91.7% at 4 ℃, while the fresh color and the moisture content of meat remained unchanged [106]. Ekram et al. isolated strains producing collagenases from slaughterhouse waste and applied them to the beef tenderization process. The results showed that the enzyme produced by the isolated strains was able to tenderize meat [107]. It is worth noting that safety concerns about pathogenicity and other unfavorable effects have limited the industrial use of microbial collagenases in the process of meat tenderization. One possible approach is the production of recombinant collagenases in non-pathogenic micro-organisms, which may avoid the existence of virulence factors. A recombinant metallopeptidase of Aeromonas salmonicida was produced using Pichia pastoris transformation for further meat tenderization. The pure peptide “Pro-Leu-Gly-Met-Trp-Ser-Arg” was used to determine the activity of recombinant collagenase. The concentration of the substrate (peptide) after 180 min was two times lower than that of the control. Meanwhile, the result of histological studies of beef shank samples revealed a marked detachment of the perimetry from the muscle bundles and the disintegration of collagen fibrils, while the muscle fibers remained unchanged [108].

Collagen hydrolysates are bioactive and have been approved by the FDA as Generally Recognized as Safe (GRAS), which makes them widely exploited in the pharmaceutical, food and cosmetic industries [109,110,111,112]. Collagenases can be used for the preparation of collagen hydrolysates due to the fact that the hydrolysates obtained using enzymatic hydrolysis are environmental-friend and safe. A multitude of works have focused on the isolation of collagenases produced by various micro-organisms and their use for the hydrolysis of collagen tissues. Lima et al. described a simplified strategy to hydrolyze type I collagen using a collagenolytic protease produced by P. aurantiogriseum. The maximum value of the degree of collagen hydrolysis was achieved at 7.5 mg/mL collagen concentration, pH 8.0 and 25 °C [113]. Furthermore, some byproducts in the food industry have been used for the preparation of collagen hydrolysates. Recombinant bacterial collagenase mining using Bacillus cereus was successfully performed and applied to hydrolyze bovine bone to obtain collagen-soluble peptides. The collagenase exhibited optimal collagen hydrolytic capacity under the conditions of 110.0 μg/mL collagenase concentration, 35 ℃, pH 8.0 and 6.0 h. [114]. Recently, marine bacteria have become an important source for the identification of novel collagenases. Collagenases produced by marine bacteria are usually more efficient in the catalysis of marine collagen in fish skin and bone than collagenases from land animals [108]. Yang and coworkers used Pseudoalteromonas sp. SJN2 to optimize the method for producing marine collagenases. The experimental collagenase production was 322.58 U/mL under the optimal fermentation conditions [115].

3 Regulatory approaches using collagenase inhibitors

MMPs are involved in the degradation of ECM proteins. Ever since their first description, MMPs and their inhibitors have been in the focus of pharmaceutical research. To regulate the enzyme activity of MMPs, researchers have screened and developed MMP inhibitors (MMPis) in recent decades [116]. Over time, the overuse of antibiotics led to a high number of bacterial populations resistant to multiple antimicrobials. A promising way of overcoming the problem is to design compounds that target virulence factors rather than bacteria. As a prominent virulence factor, it is necessary to design inhibitors against bacterial collagenases. Inhibitors can be divided into two categories: zinc-binding inhibitors and non zinc-binding inhibitors.

3.1 Zinc-binding inhibitors

The first zinc-binding inhibitors were small molecular mimics of native peptide substrates; the design idea was to bind the small molecular mimics of these peptides with a zinc-binding group (ZBG) represented by hydroxamic acid and chelate the catalytic Zn(II) ion. The ZBGs in these inhibitors serve to replace the zinc-bound water molecule, lock the inhibitor in the active site and direct the backbone of the inhibitor to the substrate binding pocket to inactivate the enzyme [117, 118]. These inhibitors have a wide range of inhibition and show a strong inhibitory effect because the catalytic Zn(II) ion is conserved in MMPs and bacteria collagenases. Hydroxamic acid (–CONHOH) emerged as a preferred zinc-binding group, as it can effectively interact with the catalytic Zn(II) ion in a bidentate fashion (Fig. 6) and is relatively easy to synthesize, which makes hydroxamic acid-based compounds account for a large part of the inhibitors [119,120,121,122]. Compounds 1 (Marimastat), 2 (Batimastat) and 3 (Ilomastat) are typical examples of such inhibitors (Table 3) [123,124,125].

Fig. 6
figure 6

Diagram of chelation of ZBGs (a hydroxamate; b carboxylate; c phosphonate; d thiol; e phosphinate) with the catalytic Zn(II) ion (represented by a blue sphere)

Table 3 List of inhibitors with different inhibitory principles

However, there has been some concern about the use of strong chelation of Zn(II) ions with hydroxamic acid as the ZBG, such as low selectivity for different enzymes and a possible off-target effect [126]. Accordingly, researchers have been looking for other relatively weak ZBGs to replace hydroxamic acid. Many functional groups have been investigated, such as carboxylic acid, phosphonate, phosphinate, and thiol (compounds 4, 5, 6 and 7; Table 3) [127,128,129,130]. Figure 5 shows the chelation of these groups with the zinc ion. The application of these weaker ZBGs reduces the effect of metal chelation on binding affinity and enhances the supramolecular interaction between ligands and proteins [131, 132].

Due to the occupancy of the catlytic Zn(II) ion by the zinc-binding inhibitors, the substrates cannot be hydrolyzed by collagenases. As a result, these inhibitors are all competitive inhibitors.

3.2 Non zinc-binding inhibitors

3.2.1 Catalytic domain (non zinc-binding) inhibitors

In addition to the catalytic Zn(II) ion binding site, there are several other binding sites (recognition pockets) in the catalytic domain of MMPs. Some of them are on the right side of the catalytic Zn(II) ion (called primed side, with pockets being named S1′, S2′ and S3′); others are on the left side of the catalytic Zn(II) ion (called unprimed side, with pockets being named S1, S2 and S3) [133]. The interaction with these pockets has been found to be able to guide the development of new species of inhibitors. Among them, it has been proved that the S1' subsite (pocket) presents the greatest difference in amino acid sequence and depth among different MMPs, which is very important for substrate selectivity [134]. Based on the depth of the S1' pocket, MMPs can be divided into shallow (such as MMP-1), intermediate (such as MMP-8) and deep (such as MMP-13) [135]. Learning from the limitations of the former MMPis, researchers began to shift their attentions to the catalytic domain of MMPs and have developed several non zinc-binding inhibitors targeting the catalytic domain (compounds 8, 9 and 10; Table 3) [134,135,136,137,138,139,140]. Most of these inhibitors have long molecular structures and have aromatic or planar connecting ring structures [141].

It is notable that different MMP inhibitors can be designed according to the depth of the pocket, but it is not the only issue to consider. For example, the X-ray structure of MMP-1 binding to an inhibitor (compound 11; Table 3) with a diphenyl ether sulfone substituent indicates that the group can penetrate the S1′ pocket with an induced matching mechanism. This indicates that the consideration of enzyme kinetics is equally crucial in the process of collagenase binding with the substrate [142].

3.2.2 Inhibitors binding to the hemopexin-like domain

Apart from the catalytic domain, MMPs also contain other distal structures, such as a propeptide domain and a hemopexin-like domain. Previous works have not devoted much attention to them, because these domains are far from the catalytic domain. It is now recognized that the hemopexin-like domain can allosterically manipulate enzyme activity during collagen hydrolysis, which can be applied for the development of novel inhibitors (compound 12 and 13, Table 3) [143,144,145,146,147,148,149]. With increasing studies being conducted on the overall structure of MMPs, these non-catalytic domains may be further utilized for designing selective MMPis. Unlike inhibitors that employ Zn(II) chelation, these compounds generally display a non-competitive mechanism of inhibition [141]. Although, some inhibitors still have a competitive inhibition mechanism [150, 151]. As a consequence, the absence of a ZBG is not an obvious guarantee of a non-competitive/uncompetitive mechanism, and a detailed kinetic characterization is required.

3.3 Applications of collagenase inhibitors

3.3.1 Biomedical field

MMPs have the potential to degrade different components of the ECM, including collagen, and play a role in cell proliferation and apoptosis, as well as physiological processes such as immune function and tissue healing [152]. The over-expression of MMPs may give rise to quite a few diseases, such as osteoarthritis (OA) and cancer [153]. Consequently, the use of MMPis becomes a promising therapeutic tool for these diseases.

The inhibition of MMP-13 is one of the most promising approaches for the treatment of cartilage degradation in OA, which is an age-related chronic disease characterized by the destruction of articular cartilage and progressive degradation due to mechanical stress [154]. The fact that type II collagen (the major structure protein of the cartilage matrix) can be degraded by MMP-13 makes the enzyme a pivotal role in OA formation [155, 156]. Several strategies, such as high-throughput screening (HTS) and natural-product-derived fragments (NPDFs) [140], have been used for discovering of MMP-13 inhibitors and have yielded a number of MMP-13 inhibitors, including zinc-binding inhibitors and non zinc-binding inhibitors. A variety of ZBGs have been used in zinc-binding inhibitors, such as hydroxamic acids, reverse hydroxamic acids, pyrimidinetriones and carboxylic acids [127, 157,158,159].

A series of compounds that possess a 1,2,4-triazol-3-yl group as a ZBG were designed and synthesized as MMP-13 inhibitors. Among these evaluated compounds, compound 14 (Table 4) exhibited outstanding potency against MMP-13 (IC50 = 0.036 nM) and selectivity (> 1500-fold) for other MMPs and tumor necrosis factor-α converting enzyme (TACE). Further evaluation demonstrated that compound 12 was effective in preventing in vitro degradation (70.8% inhibition of cartilage degradation at 1 µM) [160]. Recently, a series of novel inhibitors based on a previously developed MMP-13 inhibitor that interacts with the S1′ subsite and reaches over the catalytically active Zn(II) ion within the active site were also designed and synthesized, among which oxetane-containing compound 15 (Table 4) was found to exhibit favorable inhibition of MMP-13 (IC50 = 42 nM) in in vitro studies with an excellent selectivity among other MMPs. Furthermore, in vivo pharmacokinetic studies within the synovial fluid of the rat knee joint were performed and demonstrated that the compound is a promising lead compound for osteoarthritis [161].

Table 4 List of MMP-13 inhibitors for the treatment of OA

Apart from binding to the Zn(II) ion, it is found that the introduction of fully hydrophobic interaction with its S1′ pocket is another common strategy for obtaining inhibitory potency and selectivity against MMP-13 [162]. The synthesis and evaluation of a non zinc-binding MMP-13 selective inhibitor (compound 16; Table 4) were reported. The compound possessed outstanding potency against the catalytic domain of recombinant human MMP-13 (IC50 = 0.03 nM) with great selectivity for other MMPs, TACE and Aggrecanase 1 (> 20,000-fold). Furthermore, it could penetrate and attain high, sustained concentrations (≥ 2 μM for 8 weeks) in cartilage and demonstrated 100% inhibition for 3 weeks [163]. Bendele reported a potent pyrimidine dicarboxamide derivative, compound 17 (Table 4), as a potential disease-modifying OA drug (DMOAD). The potency of compound 17 against the catalytic domain of recombinant human MMP-13 was determined to be 4.8 nM, and it showed high selectivity, with little inhibition for other MMPs (IC50 > 100,000 nM for MMP-1, -2, -3, -7, -12, and -14) [164].

In addition to the treatment of arthritis, MMPis also play a role in other fields of medicine, such as cancer [165, 166], idiopathic pulmonary fibrosis [167], cardiac disease [168] and acute lung injury [169].

Alternative non-antibiotic therapy is urgently needed with the gradual rising in bacterial drug resistance and the slow discovery of new antibiotics [170]. The pharmacological inhibition of bacterial collagenase is a promising strategy to block bacterial virulence without exerting selective pressure [171]. Several compounds have been designed that bind closely to the active site of bacterial collagenase using different ZBGs to obtain the expected inhibitory potency [172,173,174,175,176,177,178], some of which, while being reasonably effective inhibitors of clostridium collagenase, demonstrate negligible activity against human MMPs. Schönauer and coworkers discovered an N-aryl mercaptoacetamide-based inhibitor scaffold using surface plasmon resonance-based screening complemented with enzyme inhibition assays. Compound 18 (Table 5), containing a thiocarbamate unit, showed submicromolar affinity (IC50 = 0.01 μM) for ColH from the human pathogen Clostridium histolyticum with more than 1000-fold selectivity for MMPs [177]. Recently, they substituted the thiol moiety with phosphonate, which is not prone to oxidation or degradation, while keeping the N-arylacetamide core structure intact. This novel compound, compound 19 (Table 5), retained the binding mode of compound 18 with reasonably potent inhibition (IC50 = 7 μM) against the clostridial collagenase ColH remained. Therefore, it still showed remarkable selectivity for other MMPs and even better selectivity for other human off-targets (histone deacetylases (HDACs) and TACE) [178]. These studies may pave the way for the development of selective bacterial collagenase inhibitors with potential therapeutic applications in humans.

Table 5 List of bacterial collagenases (ColH) inhibitors

3.3.2 Leather manufacturing

Collagenase inhibitors are applied in leather manufacturing. The conventional unhairing process generates and discharges a wide range of pollutants [179], urging people to look for novel materials and processes. Compared with traditional methods, enzymatic unhairing shows advantages in reducing the pollution load [180]. However, the collagenase component in enzyme preparations hydrolyzes the collagens of the hide and skin, which may cause grain damage. Bivalent metal ions Mn(II), Mg(II), and Zn(II) were introduced to inhibit the activity of AS1.398 protease of hydrolyzing collagen during the enzymatic unhairing of bovine hide. The presence of Mn(II), Mg(II), and Zn(II) ions at 10 mM led to the AS1.398 protease hydrolysis of collagen being inhibited by 10%, 35% and 55%, respectively. Compared with metal ion-free enzymatic unhairing, the addition of certain bivalent metal ions resulted in the same unhairing rate with less damage to the grain and hair pores [181, 191]. Furthermore, Li et al. [182] investigated the mixture inhibitors system based on Cu(II) ions for the enzymatic unhairing of bovine hide. The results inferred that the addition of 5 mmol/L Cu(II) and 20 mmol/L sodium succinate could maintain the unhairing rate while performing efficient inhibition of collagenolytic activity; thus, the leather was subjected to less grain damage and maintained an intact pore structure.

Currently, the study of collagenases and their inhibitors mainly focuses on the medical field, and only some studies investigated enzymatic unhairing in the leather-making process. The inhibitory effect of different inhibitors on enzyme preparations for unhairing also needs to be urgently solved. Insights into collagenases and their inhibitors could bring new impetus to the innovation and promotion of the enzymatic unhairing process.

4 Conclusion

MMPs and bacterial collagenases are two kinds of collagenases that can cleave the triple-helical domain of native fibrillar collagens. They have different domain compositions and catalytic mechanisms on hydrolysis of collagens. This review presents approaches on the applications of collagenases and their inhibitors in biomedicine, food processing, and the enzymatic unhairing process in leather-making industry. MMPs and bacterial collagenases are recognized as targets for the treatment of a variety of diseases, therefore their inhibitors are desperately needed and the achievements about the collagenases and their inhibitors derived from previous studies are helpful for the future research.

Availability of data and materials

Not applicable.

Abbreviations

ECM:

Extracellular matrix

MMP:

Matrix metalloproteinase

MT-MMPs:

Membranetype-MMPs

Prop:

Propeptide

Cat:

Catalytic domain

Hpx:

Hemopexin-like domain

AD:

Activator domain

PD:

Peptidase domain

CBD:

Collagen binding domain

PKD:

Polycystic kidney disease-like domain

FDA:

Food and Drug Administration

DD:

Dupuytren’s disease

GFP:

Green fluorescent protein

IFP:

Interstitial fluid pressure

GRAS:

Generally recognized as safe

ATPS:

Aqueous two-phase system

ABTS:

2,2′-Azino-bis (3-ethylbenzothiazoline-6-sulfonic acid)

DPPH:

1,1-Diphenyl-2-picryl-hydrazyl

ORAC:

Oxygen radical absorbance capacity

MMPi:

MMP inhibitor

ZBG:

Zinc binding group

OA:

Osteoarthritis

HTS:

High-throughput screening

NPDFs:

Natural-product-derived fragments

TACE:

Tumor necrosis factor-α converting enzyme

DMOAD:

Disease modifying OA drug

HDAC:

Histone deacetylases

References

  1. Gordon MK, Hahn RA. Collagens. Cell Tissue Res. 2010;339(1):247–57.

    Article  CAS  Google Scholar 

  2. Sorushanova A, Delgado LM, Wu Z, et al. The collagen suprafamily: from biosynthesis to advanced biomaterial development. Adv Mater. 2019;31(1):1801651.

    Article  Google Scholar 

  3. Zeugolis DI, Raghunath M. The physiological relevance of wet versus dry differential scanning calorimetry for biomaterial evaluation: a technical note. Polym Int. 2010;59(10):1403–7.

    Article  CAS  Google Scholar 

  4. Bella J. Collagen structure: new tricks from a very old dog. Biochem J. 2016;473(8):1001–25.

    Article  CAS  Google Scholar 

  5. Boryskina OP, Bolbukh TV, Semenov MA, et al. Energies of peptide–peptide and peptide–water hydrogen bonds in collagen: evidences from infrared spectroscopy, quartz piezogravimetry and differential scanning calorimetry. J Mol Struct. 2007;827(1–3):1–10.

    Article  CAS  Google Scholar 

  6. Hyde TJ, Bryan MA, Brodsky B, et al. Sequence dependence of renucleation after a Gly mutation in model collagen peptides. J Biol Chem. 2006;281(48):36937–43.

    Article  CAS  Google Scholar 

  7. Khoshnoodi J, Cartailler JP, Alvares K, et al. Molecular recognition in the assembly of collagens: terminal noncollagenous domains are key recognition modules in the formation of triple helical protomers. J Biol Chem. 2006;281(50):38117–21.

    Article  CAS  Google Scholar 

  8. Buehler MJ. Nature designs tough collagen: explaining the nanostructure of collagen fibrils. Proc Natl Acad Sci. 2006;103(33):12285–90.

    Article  CAS  Google Scholar 

  9. Birk DE, Bruckner P. Collagen suprastructures. Collagen. 2005;185–205.

  10. Ramshaw JAM. Distribution of type III collagen in bovine skin of various ages. Connect Tissue Res. 1986;14(4):307–14.

    Article  CAS  Google Scholar 

  11. Ricard-Blum S. The collagen family. Cold Spring Harbor Perspect Biol. 2011;3(1):1–19.

    Article  CAS  Google Scholar 

  12. Naomi R, Ridzuan PM, Bahari H. Current insights into collagen type I. Polymers. 2021;13:2642.

    Article  CAS  Google Scholar 

  13. Holmes DF, Lu Y, Starborg T, et al. Collagen fibril assembly and function. Curr Top Dev Biol. 2018;130:107–42.

    Article  CAS  Google Scholar 

  14. Naomi R, Fauzi MB. Cellulose/collagen dressings for diabetic foot ulcer: a review. Pharmaceutics. 2020;12(9):881.

    Article  CAS  Google Scholar 

  15. Pardo A, Selman M. MMP-1: the elder of the family. Int J Biochem Cell Biol. 2005;37(2):283–8.

    Article  CAS  Google Scholar 

  16. Lauer-Fields JL, Juska D, Fields GB. Matrix metalloproteinases and collagen catabolism. Peptide Sci. 2002;66(1):19–32.

    Article  CAS  Google Scholar 

  17. Zhang YZ, Ran LY, Li CY, et al. Diversity, structures, and collagen-degrading mechanisms of bacterial collagenolytic proteases. Appl Environ Microbiol. 2015;81(18):6098–107.

    Article  CAS  Google Scholar 

  18. Lepetit J. Collagen contribution to meat toughness: theoretical aspects. Meat Sci. 2008;80(4):960–7.

    Article  CAS  Google Scholar 

  19. Guo L, Harnedy PA, O’Keeffe MB, et al. Fractionation and identification of Alaska pollock skin collagen-derived mineral chelating peptides. Food Chem. 2015;173:536–42.

    Article  CAS  Google Scholar 

  20. Naim A, Pan Q, Baig MS. Matrix metalloproteinases (MMPs) in liver diseases. J Clin Exp hepatol. 2017;7(4):367–72.

    Article  Google Scholar 

  21. Bhagwat PK, Dandge PB. Collagen and collagenolytic proteases: a review. Biocatal Agric Biotechnol. 2018;15:43–55.

    Article  Google Scholar 

  22. Anzani C, Prandi B, Buhler S, et al. Towards environmentally friendly skin unhairing process: a comparison between enzymatic and oxidative methods and analysis of the protein fraction of the related wastewaters. J Clean Prod. 2017;164:1446–54.

    Article  CAS  Google Scholar 

  23. Sujitha P, Kavitha S, Shakilanishi S, et al. Enzymatic dehairing: a comprehensive review on the mechanistic aspects with emphasis on enzyme specificity. Int J Biol Macromol. 2018;118:168–79.

    Article  CAS  Google Scholar 

  24. Jaouadi B, Ellouz-Chaabouni S, Ali MB, et al. Excellent laundry detergent compatibility and high dehairing ability of the Bacillus pumilus CBS alkaline proteinase (SAPB). Biotechnol Bioprocess Eng. 2009;14(4):503.

    Article  CAS  Google Scholar 

  25. Kerouaz B, Jaouadi B, Brans A, et al. Purification and biochemical characterization of two novel extracellular keratinases with feather-degradation and hide-dehairing potential. Process Biochem. 2021;106:137–48.

    Article  CAS  Google Scholar 

  26. Bouacem K, Bouanane-Darenfed A, Jaouadi NZ, et al. Novel serine keratinase from Caldicoprobacter algeriensis exhibiting outstanding hide dehairing abilities. Int J Biol Macromo. 2016;86:321–8.

    Article  CAS  Google Scholar 

  27. Murphy G, Nagase H. Progress in matrix metalloproteinase research. Mol Aspects Med. 2008;29(5):290–308.

    Article  CAS  Google Scholar 

  28. Nagase H, Visse R, Murphy G. Structure and function of matrix metalloproteinases and TIMPs. Cardiovasc Res. 2006;69(3):562–73.

    Article  CAS  Google Scholar 

  29. Lemaitre V, D’Armiento J. Matrix metalloproteinases in development and disease. Birth Defects Res Part C. 2006;78(1):1–10.

    Article  CAS  Google Scholar 

  30. Kapoor C, Vaidya S, Wadhwan V, et al. Seesaw of matrix metalloproteinases (MMPs). J Cancer Res Ther. 2016;12(1):28–35.

    Article  CAS  Google Scholar 

  31. Jabłońska-Trypuć A, Matejczyk M, Rosochacki S. Matrix metalloproteinases (MMPs), the main extracellular matrix (ECM) enzymes in collagen degradation, as a target for anticancer drugs. J Enzym Inhib Med Chem. 2016;31:177–83.

    Article  Google Scholar 

  32. Laronha H, Caldeira J. Structure and function of human matrix metalloproteinases. Cells. 2020;9(5):1076.

    Article  CAS  Google Scholar 

  33. Kahari VM, Saarialho-Kere U. Matrix metalloproteinases in skin. Exp Dermatol. 1997;6(5):199–213.

    Article  CAS  Google Scholar 

  34. Cui N, Hu M, Khalil RA. Biochemical and biological attributes of matrix metalloproteinases. Prog Mol Biol Transl Sci. 2017;147:1–73.

    Article  CAS  Google Scholar 

  35. Amalinei C, Caruntu ID, Balan RA. Biology of metalloproteinases. Rom J Morphol Embryol. 2007;48(4):323–34.

    Google Scholar 

  36. Tallant C, Marrero A, Gomis-Rüth FX. Matrix metalloproteinases: fold and function of their catalytic domains. Biochim Biophys Acta Mol Cell Res. 2010;1803(1):20–8.

    Article  CAS  Google Scholar 

  37. Coussens LM, Fingleton B, Matrisian LM. Matrix metalloproteinase inhibitors and cancer—trials and tribulations. Sci. 2002;295(5564):2387–92.

    Article  CAS  Google Scholar 

  38. Klein T, Bischoff R. Physiology and pathophysiology of matrix metalloproteases. Amino Acids. 2011;41(2):271–90.

    Article  CAS  Google Scholar 

  39. Visse R, Nagase H. Matrix metalloproteinases and tissue inhibitors of metalloproteinases: structure, function, and biochemistry. Cirs Res. 2003;92(8):827–39.

    Article  CAS  Google Scholar 

  40. Mannello F, Medda V. Nuclear localization of matrix metalloproteinases. Prog Histochem Cytochem. 2012;47(1):27–58.

    Article  Google Scholar 

  41. Rangasamy L, Di Geronimo B, Ortin I, et al. Molecular imaging probes based on matrix metalloproteinase inhibitors (MMPIs). Molecules. 2019;24(16):2982.

    Article  CAS  Google Scholar 

  42. Lockhart AC, Braun RD, Yu D, et al. Reduction of wound angiogenesis in patients treated with BMS-275291, a broad spectrum matrix metalloproteinase inhibitor. Clin Cancer Res. 2003;9(2):586–93.

    CAS  Google Scholar 

  43. Cerofolini L, Fragai M, Luchinat C. Mechanism and inhibition of matrix metalloproteinases. Curr Med Chem. 2019;26(15):2609–33.

    Article  CAS  Google Scholar 

  44. Spurlino JC, Smallwood AM, Carlton DD, et al. 156 Å structure of mature truncated human fibroblast collagenase. Proteins Struct Funct Bioinf. 1994;19(2):98–109.

    Article  CAS  Google Scholar 

  45. Maskos K. Crystal structures of MMPs in complex with physiological and pharmacological inhibitors. Biochime. 2005;87(3–4):249–63.

    Article  CAS  Google Scholar 

  46. Piccard H, Van den Steen PE, Opdenakker G. Hemopexin domains as multifunctional liganding modules in matrix metalloproteinases and other proteins. J Leukoc Biol. 2007;81(4):870–92.

    Article  CAS  Google Scholar 

  47. Li J, Brick P, O’hare MC, et al. Structure of full-length porcine synovial collagenase reveals a C-terminal domain containing a calcium-linked, four-bladed β-propeller. Structrue. 1995;3(6):541–9.

    Article  CAS  Google Scholar 

  48. Gomis-Ruth FX, Gohlke U, Betz M, et al. The helping hand of collagenase-3 (MMP-13): 2.7 Å crystal structure of its C-terminal haemopexin-like domain. J Mol Biol. 1996;264(3):556–66.

    Article  CAS  Google Scholar 

  49. Tsukada H, Pourmotabbed T. Unexpected crucial role of residue 272 in substrate specificity of fibroblast collagenase. J Biol Chem. 2002;277(30):27378–84.

    Article  CAS  Google Scholar 

  50. Andreini C, Banci L, Bertini I, et al. Bioinformatic comparison of structures and homology-models of matrix metalloproteinases. J Phys Chem Lett. 2004;3(1):21–31.

    CAS  Google Scholar 

  51. Bertini I, Fragai M, Luchinat C, et al. Interdomain flexibility in full-length matrix metalloproteinase-1 (MMP-1). J Biol Chem. 2009;284(19):12821–8.

    Article  CAS  Google Scholar 

  52. Fields GB. Interstitial collagen catabolism. J Biol Chem. 2013;288(13):8785–93.

    Article  CAS  Google Scholar 

  53. Varghese A, Chaturvedi SS, Fields GB, et al. A synergy between the catalytic and structural Zn (II) ions and the enzyme and substrate dynamics underlies the structure–function relationships of matrix metalloproteinase collagenolysis. J Biol Inorg Chem. 2021;26(5):583–97.

    Article  CAS  Google Scholar 

  54. Lauer-Fields JL, Chalmers MJ, Busby SA, et al. Identification of specific hemopexin-like domain residues that facilitate matrix metalloproteinase collagenolytic activity. J Biol Chem. 2009;284(36):24017–24.

    Article  CAS  Google Scholar 

  55. Cerofolini L, Fields GB, Fragai M, et al. Examination of matrix metalloproteinase-1 in solution: a preference for the pre-collagenolysis state. J Biol Chem. 2013;288(42):30659–71.

    Article  CAS  Google Scholar 

  56. Manka SW, Brew K. Thermodynamic and mechanistic insights into coupled binding and unwinding of collagen by matrix metalloproteinase 1. J Mol Biol. 2020;432(22):5985–93.

    Article  CAS  Google Scholar 

  57. Karabencheva-Christova TG, Christov CZ, Fields GB. Collagenolytic matrix metalloproteinase structure-function relationships: insights from molecular dynamics studies. Adv Protein Chem Str. 2017;109:1–24.

    CAS  Google Scholar 

  58. Kumar L, Nash A, Harms C, et al. Allosteric communications between domains modulate the activity of matrix metalloprotease-1. Biophys J. 2020;119(2):360–74.

    Article  CAS  Google Scholar 

  59. Van Wart HE, Birkedal-Hansen H. The cysteine switch: a principle of regulation of metalloproteinase activity with potential applicability to the entire matrix metalloproteinase gene family. Proc Natl Acad Sci. 1990;87(14):5578–82.

    Article  Google Scholar 

  60. Morrison CJ, Butler GS, Rodríguez D, et al. Matrix metalloproteinase proteomics: substrates, targets, and therapy. Curr Opin Cell Biol. 2009;21(5):645–53.

    Article  CAS  Google Scholar 

  61. Overall CM. Molecular determinants of metalloproteinase substrate specificity. Mol Biotechnol. 2002;22(1):51–86.

    Article  CAS  Google Scholar 

  62. Bigg HF, Rowan AD, Barker MD, et al. Activity of matrix metalloproteinase-9 against native collagen types I and III. FEBS J. 2007;274(5):1246–55.

    Article  CAS  Google Scholar 

  63. Springman EB, Angleton EL, Birkedal-Hansen H, et al. Multiple modes of activation of latent human fibroblast collagenase: evidence for the role of a Cys73 active-site zinc complex in latency and a" cysteine switch" mechanism for activation. Proc Natl Acad Sci. 1990;87(1):364–8.

    Article  CAS  Google Scholar 

  64. Manka SW, Carafoli F, Visse R, et al. Structural insights into triple-helical collagen cleavage by matrix metalloproteinase 1. Proc Natl Acad Sci. 2012;109(31):12461–6.

    Article  CAS  Google Scholar 

  65. Chung L, Shimokawa K, Dinakarpandian D, et al. Identification of the 183RWTNNFREY191 region as a critical segment of matrix metalloproteinase 1 for the expression of collagenolytic activity. J Biol Chem. 2000;275(38):29610–7.

    Article  CAS  Google Scholar 

  66. Arnold LH, Butt LE, Prior SH, et al. The interface between catalytic and hemopexin domains in matrix metalloproteinase-1 conceals a collagen binding exosite. J Biol Chem. 2011;286(52):45073–82.

    Article  CAS  Google Scholar 

  67. Bertini I, Fragai M, Luchinat C, et al. Structural basis for matrix metalloproteinase 1-catalyzed collagenolysis. J Am Chem Soc. 2012;134(4):2100–10.

    Article  CAS  Google Scholar 

  68. Chung L, Dinakarpandian D, Yoshida N, et al. Collagenase unwinds triple-helical collagen prior to peptide bond hydrolysis. EMBO J. 2004;23(15):3020–30.

    Article  CAS  Google Scholar 

  69. Liu J, Khalil RA. Matrix metalloproteinase inhibitors as investigational and therapeutic tools in unrestrained tissue remodeling and pathological disorders. Prog Mol Biol Transl Sci. 2017;148:355–420.

    Article  CAS  Google Scholar 

  70. Whittaker M, Floyd CD, Brown P, et al. Design and therapeutic application of matrix metalloproteinase inhibitors. Chem Rev. 1999;99(9):2735–76.

    Article  CAS  Google Scholar 

  71. Matsushita O, Jung CM, Katayama S, et al. Gene duplication and multiplicity of collagenases in Clostridium histolyticum. J Bacteriol. 1999;181(3):923–33.

    Article  CAS  Google Scholar 

  72. Duarte AS, Correia A, Esteves AC. Bacterial collagenases–a review. Crit Rev Microbiol. 2016;42(1):106–26.

    Article  Google Scholar 

  73. Shinoda S, Miyoshi SI. Proteases produced by vibrios. Biocontrol Sci. 2011;16(1):1–11.

    Article  CAS  Google Scholar 

  74. Eckhard U, Schönauer E, Nüss D, et al. Structure of collagenase G reveals a chew-and-digest mechanism of bacterial collagenolysis. Nat Struct Mol Biol. 2011;18(10):1109–14.

    Article  CAS  Google Scholar 

  75. López-Pelegrín M, Cerdà-Costa N, Martínez-Jiménez F, et al. A novel family of soluble minimal scaffolds provides structural insight into the catalytic domains of integral membrane metallopeptidases. J Biol Chem. 2013;288(29):21279–94.

    Article  Google Scholar 

  76. Eckhard U, Schonauer E, Brandstetter H. Structural basis for activity regulation and substrate preference of clostridial collagenases G, H, and T. J Biol Chem. 2013;288(28):20184–94.

    Article  CAS  Google Scholar 

  77. Matsushita O, Koide T, Kobayashi R, et al. Substrate recognition by the collagen-binding domain of Clostridium histolyticum class I collagenase. J Biol Chem. 2001;276(12):8761–70.

    Article  CAS  Google Scholar 

  78. Eckhard U, Brandstetter H. Polycystic kidney disease-like domains of clostridial collagenases and their role in collagen recruitment. Biol Chem. 2011;392(11):1039–45.

    Article  CAS  Google Scholar 

  79. Wang YK, Zhao GY, Li Y, et al. Mechanistic insight into the function of the C-terminal PKD domain of the collagenolytic serine protease deseasin MCP-01 from deep sea Pseudoalteromonas sp. SM9913: binding of the PKD domain to collagen results in collagen swelling but does not unwind the collagen triple helix. J Biol Chem. 2010;285(19):14285–91.

    Article  CAS  Google Scholar 

  80. Ohbayashi N, Yamagata N, Goto M, et al. Enhancement of the structural stability of full-length clostridial collagenase by calcium ions. Appl Environ Microb. 2012;78(16):5839–44.

    Article  CAS  Google Scholar 

  81. French MF, Bhown A, Van Wart HE. Identification of Clostridium histolyticum collagenase hyperreactive sites in type I, II, and III collagens: lack of correlation with local triple helical stability. J Protein Chem. 1992;11(1):83–97.

    Article  CAS  Google Scholar 

  82. Wolfe HR, Rosenberg E, Ciftci K, et al. Evaluation of alternative diluents for clinical use of collagenase Clostridium histolyticum (CCH-aaes). J Cosmet Dermatol. 2021;20(6):1643–7.

    Article  Google Scholar 

  83. Song M, Lee S, Choe D, et al. Clinical and biological evaluations of biodegradable collagen matrices for glaucoma drainage device implantation. Invest Ophthalmol Vis Sci. 2017;58(12):5329–35.

    Article  CAS  Google Scholar 

  84. Gur S, Limin M, Hellstrom WJG. Current status and new developments in Peyronie’s disease: medical, minimally invasive and surgical treatment options. Expert Opin Pharmacother. 2011;12(6):931–44.

    Article  Google Scholar 

  85. Ramundo J, Gray M. Enzymatic wound debridement. J Wound Ostomy Cont. 2008;35(3):273–80.

    Article  Google Scholar 

  86. Kirshen C, Woo K, Ayello EA, et al. Debridement: a vital component of wound bed preparation. Adv Skin Wound Care. 2006;19(9):506–17.

    Article  Google Scholar 

  87. Hoppe IC, Granick MS. Debridement of chronic wounds: a qualitative systematic review of randomized controlled trials. Clin Plast Surg. 2012;39(3):221–8.

    Article  Google Scholar 

  88. Tallis A, Motley TA, Wunderlich RP, et al. Clinical and economic assessment of diabetic foot ulcer debridement with collagenase: results of a randomized controlled study. Clin Ther. 2013;35(11):1805–20.

    Article  Google Scholar 

  89. Karagol BS, Okumus N, Dursun A, Karadag N, Zencıroglu A. Early and successful enzymatıc debridement via collagenase application to pinna in a preterm neonate. Pediatr Dermatol. 2011;28(5):600–1.

    Article  Google Scholar 

  90. Frederick RE, Bearden R, Jovanovic A, et al. Clostridium collagenase impact on zone of stasis stabilization and transition to healthy tissue in burns. Int J Mol Sci. 2021;22(16):8643.

    Article  CAS  Google Scholar 

  91. De Francesco F, De Francesco M, Riccio M. Hyaluronic acid/collagenase ointment in the treatment of chronic hard-to-heal wounds: an observational and retrospective study. J Clin Med. 2022;11(3):537.

    Article  Google Scholar 

  92. McCallon SK, Weir D, Lantis JC 2nd. Optimizing wound bed preparation with collagenase enzymatic debridement. J Am Coll Clin Wound Spec. 2015;6(1–2):14–23.

    Google Scholar 

  93. Hurst LC, Badalamente MA, Hentz VR, et al. Injectable collagenase Clostridium histolyticum for Dupuytren’s contracture. N Engl J Med. 2009;361(10):968–79.

    Article  CAS  Google Scholar 

  94. Kaplan FT. Collagenase Clostridium histolyticum injection for the treatment of Dupuytren’s contracture. Drugs Today. 2011;47(9):653–67.

    Article  CAS  Google Scholar 

  95. Carr L, Michelotti B, Brgoch M, et al. Dupuytren disease management trends: a survey of hand surgeons. Hand. 2020;15(1):97–102.

    Article  Google Scholar 

  96. Badalamente MA, Hurst LC. Development of collagenase treatment for Dupuytren disease. Hand Clin. 2018;34(3):345–9.

    Article  Google Scholar 

  97. Badalamente MA, Hurst LC, Benhaim P, et al. Efficacy and safety of collagenase Clostridium histolyticum in the treatment of proximal interphalangeal joints in dupuytren contracture: combined analysis of 4 phase 3 clinical trials. J Hand Surg. 2015;40(5):975–83.

    Article  Google Scholar 

  98. Winberg M, Turesson C. Patients’ perspectives of collagenase injection or needle fasciotomy and rehabilitation for Dupuytren disease, including hand function and occupational performance. Disabil Rehabil. 2022;45(6):986–96.

    Article  Google Scholar 

  99. Nayar SK, Pfisterer D, Ingari JV. Collagenase Clostridium histolyticum injection for Dupuytren contracture: 2-year follow-up. Clin Orthop Surg. 2019;11(3):332–6.

    Article  Google Scholar 

  100. Coleman S, Gilpin D, Kaplan FT, et al. Efficacy and safety of concurrent collagenase Clostridium histolyticum injections for multiple Dupuytren contractures. J Hand Surg Am. 2014;39(1):57–64.

    Article  Google Scholar 

  101. Boe C, Blazar P, Iannuzzi N. Dupuytren contractures: an update of recent literature. J Hand Surg. 2021;46(10):896–906.

    Article  Google Scholar 

  102. Dolor A, Szoka FC Jr. Digesting a path forward: the utility of collagenase tumor treatment for improved drug delivery. Mol Pharm. 2018;15(6):2069–83.

    Article  CAS  Google Scholar 

  103. Cemazar M, Golzio M, Sersa G, et al. Hyaluronidase and collagenase increase the transfection efficiency of gene electrotransfer in various murine tumors. Hum Gene Ther. 2012;23(1):128–37.

    Article  CAS  Google Scholar 

  104. Hu J, Yuan X, Wang F, et al. The progress and perspective of strategies to improve tumor penetration of nanomedicines. Chin Chem Lett. 2021;32(4):1341–7.

    Article  Google Scholar 

  105. Kato M, Hattori Y, Kubo M, et al. Collagenase-1 injection improved tumor distribution and gene expression of cationic lipoplex. Int J Pharm. 2012;423(2):428–34.

    Article  CAS  Google Scholar 

  106. Zhao GY, Zhou MY, Zhao HL, et al. Tenderization effect of cold-adapted collagenolytic protease MCP-01 on beef meat at low temperature and its mechanism. Food Chem. 2012;134(4):1738–44.

    Article  CAS  Google Scholar 

  107. Ekram HA, Adzim H, Prasetyo EN. Beef tenderization using bacterial collagenase isolated from slaughterhouse. In: 3rd international biology conference. 2016.

  108. Minaev MY, Makhova AA. Recombinant metalloprotease as a perspective enzyme for meat tenderization. Potravinarstvo. 2019;13(1):628–33.

    Article  Google Scholar 

  109. Bilek SE, Bayram SK. Fruit juice drink production containing hydrolyzed collagen. J Funct Foods. 2015;14:562–9.

    Article  CAS  Google Scholar 

  110. Al-Nimry S, Dayah AA, Hasan I, et al. Cosmetic, biomedical and pharmaceutical applications of fish gelatin/hydrolysates. Mar drugs. 2021;19(3):145.

    Article  CAS  Google Scholar 

  111. Halim NRA, Yusof HM, Sarbon NM. Functional and bioactive properties of fish protein hydolysates and peptides: a comprehensive review. Trends Food Sci Technol. 2016;51:24–33.

    Article  CAS  Google Scholar 

  112. Prokopová A, Pavlačková J, Mokrejš P, et al. Collagen hydrolysate prepared from chicken by-product as a functional polymer in cosmetic formulation. Molecules. 2021;26(7):2021.

    Article  Google Scholar 

  113. Lima CA, Campos JF, Filho JLL, et al. Antimicrobial and radical scavenging properties of bovine collagen hydrolysates produced by Penicillium aurantiogriseum URM 4622 collagenase. J Food Sci Technol. 2015;52(7):4459–66.

    Article  CAS  Google Scholar 

  114. Song Y, Fu Y, Huang S, et al. Identification and antioxidant activity of bovine bone collagen-derived novel peptides prepared by recombinant collagenase from Bacillus cereus. Food Chem. 2021;349: 129143.

    Article  CAS  Google Scholar 

  115. Yang X, Xiao X, Liu D, et al. Optimization of collagenase production by Pseudoalteromonas sp. SJN2 and application of collagenases in the preparation of antioxidative hydrolysates. Mar Drugs. 2017;15(12):377.

    Article  Google Scholar 

  116. Brew K, Nagase H. The tissue inhibitors of metalloproteinases (TIMPs): an ancient family with structural and functional diversity. BBA Mol Cell Res. 2010;1803(1):55–71.

    CAS  Google Scholar 

  117. Jacobsen FE, Lewis JA, Cohen SM. The design of inhibitors for medicinally relevant metalloproteins. ChemMedChem. 2007;2(2):152–71.

    Article  CAS  Google Scholar 

  118. Farkas E, Enyedy EA, Micera G, et al. Coordination modes of hydroxamic acids in copper (II), nickel (II) and zinc (II) mixed-ligand complexes in aqueous solution. Polyhedron. 2000;19(14):1727–36.

    Article  CAS  Google Scholar 

  119. Rao BG. Recent developments in the design of specific matrix metalloproteinase inhibitors aided by structural and computational studies. Curr Pharm Design. 2005;11(3):295–322.

    Article  CAS  Google Scholar 

  120. Brown S, Meroueh SO, Fridman R, et al. Quest for selectivity in inhibition of matrix metalloproteinases. Curr Top Med Chem. 2004;4(12):1227–38.

    Article  CAS  Google Scholar 

  121. Bramhall SR, Rosemurgy A, Brown PD, et al. Marimastat as first-line therapy for patients with unresectable pancreatic cancer: a randomized trial. J Clin Oncol. 2001;19(15):3447–55.

    Article  CAS  Google Scholar 

  122. Botos I, Scapozza L, Zhang D, et al. Batimastat, a potent matrix mealloproteinase inhibitor, exhibits an unexpected mode of binding. Proc Natl Acad Sci. 1996;93(7):2749–54.

    Article  CAS  Google Scholar 

  123. Brown PD. Clinical studies with matrix metalloproteinase inhibitors. APMIS. 1999;107(1–6):174–80.

    Article  CAS  Google Scholar 

  124. Watson SA, Morris TM, Robinson G, et al. Inhibition of organ invasion by the matrix metalloproteinase inhibitor batimastat (BB-94) in two human colon carcinoma metastasis models. Cancer Res. 1995;55(16):3629–33.

    CAS  Google Scholar 

  125. Jin Z, Dridi N, Palui G, et al. Evaluating the catalytic efficiency of the human membrane-type 1 matrix metalloproteinase (MMP-14) Using AuNP–Peptide conjugates. J Am Chem Soc. 2023;145(8):4570–82.

    Article  CAS  Google Scholar 

  126. Gupta SP. Quantitative structure-activity relationship studies on zinc-containing metalloproteinase inhibitors. Chem Rev. 2007;107(7):3042–87.

    Article  CAS  Google Scholar 

  127. Monovich LG, Tommasi RA, Fujimoto RA, et al. Discovery of potent, selective, and orally active carboxylic acid based inhibitors of matrix metalloproteinase-13. J Med Chem. 2009;52(11):3523–38.

    Article  CAS  Google Scholar 

  128. Pochetti G, Gavuzzo E, Campestre C, et al. Structural insight into the stereoselective inhibition of MMP-8 by enantiomeric sulfonamide phosphonates. J Med Chem. 2006;49(3):923–31.

    Article  CAS  Google Scholar 

  129. Gall AL, Ruff M, Kannan R, et al. Crystal structure of the stromelysin-3 (MMP-11) catalytic domain complexed with a phosphinic inhibitor mimicking the transition-state. J Mol Biol. 2001;307(2):577–86.

    Article  CAS  Google Scholar 

  130. Grams F, Reinemer P, Powers JC, et al. X-ray structures of human neutrophil collagenase complexed with peptide hydroxamate and peptide thiol inhibitors: implications for substrate binding and rational drug design. Eur J Biochem. 1995;228(3):830–41.

    Article  CAS  Google Scholar 

  131. Agrawal A, Romero-Perez D, Jacobsen JA, et al. Zinc-binding groups modulate selective inhibition of MMPs. ChemMedChem. 2008;3(5):812–20.

    Article  CAS  Google Scholar 

  132. Leighl NB, Paz-Ares L, Douillard JY, et al. Randomized phase III study of matrix metalloproteinase inhibitor BMS-275291 in combination with paclitaxel and carboplatin in advanced non-small-cell lung cancer: National Cancer Institute of Canada-Clinical Trials Group Study BR. 18. Clin Oncol. 2005;23(12):2831–9.

    CAS  Google Scholar 

  133. Li K, Tay FR, Yiu CKY. The past, present and future perspectives of matrix metalloproteinase inhibitors. Pharmacol Therapeut. 2020;207: 107465.

    Article  CAS  Google Scholar 

  134. Song J, Tan H, Boyd SE, et al. Bioinformatic approaches for predicting substrates of proteases. J Bioinf Comput Biol. 2011;9(1):149–78.

    Article  CAS  Google Scholar 

  135. Aureli L, Gioia M, Cerbara I, et al. Structural bases for substrate and inhibitor recognition by matrix metalloproteinases. Curr Med Chem. 2008;15(22):2192–222.

    Article  CAS  Google Scholar 

  136. Fischer T, Riedl R. Targeted fluoro positioning for the discovery of a potent and highly selective matrix metalloproteinase inhibitor. ChemistryOpen. 2017;6(2):192–5.

    Article  CAS  Google Scholar 

  137. Johnson AR, Pavlovsky AG, Ortwine DF, et al. Discovery and characterization of a novel inhibitor of matrix metalloprotease-13 that reduces cartilage damage in vivo without joint fibroplasia side effects. J Biol Chem. 2007;282(38):27781–91.

    Article  CAS  Google Scholar 

  138. Zapico JM, Acosta L, Pastor M, et al. Design and synthesis of water-soluble and potent MMP-13 inhibitors with activity in human osteosarcoma cells. Int J Mol Sci. 2021;22(18):9976.

    Article  CAS  Google Scholar 

  139. Awad LF, Teleb M, Ibrahim NA, et al. Structure-based design and optimization of pyrimidine-and 1, 2, 4-triazolo [4,3-a] pyrimidine-based matrix metalloproteinase-10/13 inhibitors via Dimroth rearrangement towards targeted polypharmacology. Bioorgan Chem. 2020;96: 103616.

    Article  Google Scholar 

  140. Shunmuga Priya V, Pradiba D, Aarthy M, et al. In-silico strategies for identification of potent inhibitor for MMP-1 to prevent metastasis of breast cancer. J Biomol Struct Dyn. 2021;39(18):7274–93.

    Article  CAS  Google Scholar 

  141. Jacobsen JA, Jourden JLM, Miller MT, et al. To bind zinc or not to bind zinc: an examination of innovative approaches to improved metalloproteinase inhibition. BBA Mol Cell Res. 2010;1803(1):72–94.

    CAS  Google Scholar 

  142. Lovejoy B, Welch AR, Carr S, et al. Crystal structures of MMP-1 and-13 reveal the structural basis for selectivity of collagenase inhibitors. Nat Struct Biol. 1999;6(3):217–21.

    Article  CAS  Google Scholar 

  143. Remacle AG, Golubkov VS, Shiryaev SA, et al. Novel MT1-MMP small-molecule inhibitors based on insights into hemopexin domain function in tumor growth MT1-MMP hemopexin domain as a drug target. Cancer Res. 2012;72(9):2339–49.

    Article  CAS  Google Scholar 

  144. Morales R, Perrier S, Florent JM, et al. Crystal structures of novel non-peptidic, non-zinc chelating inhibitors bound to MMP-12. J Mol Biol. 2004;341(4):1063–76.

    Article  CAS  Google Scholar 

  145. Sarkar SK. Identification of substrate-specific allosteric fingerprints in matrix metalloprotease-1 at the single molecule level for precision control of function. Biophys J. 2023;122(3):51a.

    Article  Google Scholar 

  146. Overall CM, Butler GS. Protease yoga: extreme flexibility of a matrix metalloproteinase. Strcuture. 2007;15(10):1159–61.

    Article  CAS  Google Scholar 

  147. Sela-Passwell N, Rosenblum G, Shoham T, et al. Structural and functional bases for allosteric control of MMP activities: can it pave the path for selective inhibition? BBA Mol Cell Res. 2010;1803(1):29–38.

    CAS  Google Scholar 

  148. Dufour A, Sampson NS, Li J, et al. Small-molecule anticancer compounds selectively target the hemopexin domain of matrix metalloproteinase-9 novel synthetic inhibitors targeting the PEX domain of MMP-9. Cancer Res. 2011;71(14):4977–88.

    Article  CAS  Google Scholar 

  149. Xu X, Chen Z, Wang Y, et al. Inhibition of MMP-2 gelatinolysis by targeting exodomain–substrate interactions. Biochem J. 2007;406(1):147–55.

    Article  CAS  Google Scholar 

  150. Toth M, Bernardo MM, Gervasi DC, et al. Tissue inhibitor of metalloproteinase (TIMP)-2 acts synergistically with synthetic matrix metalloproteinase (MMP) inhibitors but not with TIMP-4 to enhance the (Membrane type 1)-MMP-dependent activation of pro-MMP-2. J Biol Chem. 2000;275(52):41415–23.

    Article  CAS  Google Scholar 

  151. Santamaria S, Nuti E, Cercignani G, et al. Kinetic characterization of 4,4 ’-biphenyl sulfonamides as selective non-zinc binding MMP inhibitors. J Enzyme Inhib Med Chem. 2015;30(6):947–54.

    Article  CAS  Google Scholar 

  152. Egeblad M, Werb Z. New functions for the matrix metalloproteinases in cancer progression. Nat Rev Cancer. 2002;2(3):161–74.

    Article  CAS  Google Scholar 

  153. Fischer T, Riedl R. Challenges with matrix metalloproteinase inhibition and future drug discovery avenues. Expert Opin Drug Discov. 2021;16(1):75–88.

    Article  CAS  Google Scholar 

  154. Arden N, Nevitt MC. Osteoarthritis: epidemiology. Best Pract Res Clin Rheumatol. 2006;20(1):3–25.

    Article  Google Scholar 

  155. Xie XW, Wan RZ, Liu ZP. Recent research advances in selective matrix metalloproteinase-13 inhibitors as anti-osteoarthritis agents. ChemMedChem. 2017;12(15):1157–68.

    Article  CAS  Google Scholar 

  156. Troeberg L, Nagase H. Proteases involved in cartilage matrix degradation in osteoarthritis. BBA Proteins Proteom. 2012;1824(1):133–45.

    Article  CAS  Google Scholar 

  157. Nuti E, Casalini F, Avramova SI, et al. NO-isopropyl sulfonamido-based hydroxamates: design, synthesis and biological evaluation of selective matrix metalloproteinase-13 inhibitors as potential therapeutic agents for osteoarthritis. J Med Chem. 2009;52(15):4757–73.

    Article  CAS  Google Scholar 

  158. De Savi C, Pape A, Cumming JG, et al. The design and synthesis of novel N-hydroxyformamide inhibitors of ADAM-TS4 for the treatment of osteoarthritis. Bioorgan Med Chem Lett. 2011;21(5):1376–81.

    Article  Google Scholar 

  159. Nara H, Sato K, Kaieda A, et al. Design, synthesis, and biological activity of novel, potent, and highly selective fused pyrimidine-2-carboxamide-4-one-based matrix metalloproteinase (MMP)-13 zinc-binding inhibitors. Bioorgan Med Chem. 2016;24(23):6149–65.

    Article  CAS  Google Scholar 

  160. Nara H, Kaieda A, Sato K, et al. Discovery of novel, highly potent, and selective matrix metalloproteinase (MMP)-13 inhibitors with a 1, 2, 4-triazol-3-yl moiety as a zinc binding group using a structure-based design approach. J Med Chem. 2017;60(2):608–26.

    Article  CAS  Google Scholar 

  161. Fuerst R, Choi JY, Knapinska AM, et al. Development of a putative Zn2+-chelating but highly selective MMP-13 inhibitor. Bioorgan Med Chem Lett. 2022;76: 129014.

    Article  CAS  Google Scholar 

  162. Fabre B, Ramos A, de Pascual-Teresa B. Targeting matrix metalloproteinases: exploring the dynamics of the S1′ pocket in the design of selective, small molecule inhibitors: miniperspective. J Med Chem. 2014;57(24):10205–19.

    Article  CAS  Google Scholar 

  163. Gege C, Bao B, Bluhm H, et al. Discovery and evaluation of a non-Zn chelating, selective matrix metalloproteinase 13 (MMP-13) inhibitor for potential intra-articular treatment of osteoarthritis. J Med Chem. 2012;55(2):709–16.

    Article  CAS  Google Scholar 

  164. Bendele AM, Neelagiri M, Neelagiri V, et al. Development of a selective matrix metalloproteinase 13 (MMP-13) inhibitor for the treatment of Osteoarthritis. Eur J Med Chem. 2021;224: 113666.

    Article  CAS  Google Scholar 

  165. Chin LT, Liu KW, Chen YH, et al. Cell-based assays and molecular simulation reveal that the anti-cancer harmine is a specific matrix metalloproteinase-3 (MMP-3) inhibitor. Comput Biol Chem. 2021;94: 107556.

    Article  CAS  Google Scholar 

  166. Peng SY, Hsiao CC, Lan TH, et al. Pomegranate extract inhibits migration and invasion of oral cancer cells by downregulating matrix metalloproteinase-2/9 and epithelial-mesenchymal transition. Environ Toxicol. 2020;35(6):673–82.

    Article  CAS  Google Scholar 

  167. Yue L, Shi Y, Su X, et al. Matrix metalloproteinases inhibitors in idiopathic pulmonary fibrosis: medicinal chemistry perspectives. Eur J Med Chem. 2021;224: 113714.

    Article  CAS  Google Scholar 

  168. Gömöri K, Szabados T, Kenyeres É, et al. Cardioprotective effect of novel matrix metalloproteinase inhibitors. Int J Mol Sci. 2020;21(19):6990.

    Article  Google Scholar 

  169. Tong Y, Yu Z, Chen Z, et al. The HIV protease inhibitor Saquinavir attenuates sepsis-induced acute lung injury and promotes M2 macrophage polarization via targeting matrix metalloproteinase-9. Cell Death Dis. 2021;12(1):67.

    Article  CAS  Google Scholar 

  170. Heim-Riether A, Taylor SJ, Liang S, et al. Improving potency and selectivity of a new class of non-Zn-chelating MMP-13 inhibitors. Bioorg Med Chem Lett. 2009;19(18):5321–4.

    Article  CAS  Google Scholar 

  171. Frieri M, Kumar K, Boutin A. Antibiotic resistance. J Infect Public Health. 2017;10(4):369–78.

    Article  Google Scholar 

  172. Nitulescu G, Nitulescu GM, Zanfirescu A, et al. Candidates for repurposing as anti-virulence agents based on the structural profile analysis of microbial collagenase inhibitors. Pharmaceutics. 2021;14(1):62.

    Article  Google Scholar 

  173. Ilies M, Banciu MD, Scozzafava A, et al. Protease inhibitors: synthesis of bacterial collagenase and matrix metalloproteinase inhibitors incorporating arylsulfonylureido and 5-dibenzo-suberenyl/suberyl moieties. Bioorgan Med Chem Lett. 2003;11(10):2227–39.

    Article  CAS  Google Scholar 

  174. Amélia Santos M, Marques S, Gil M, et al. Protease inhibitors: synthesis of bacterial collagenase and matrix metalloproteinase inhibitors incorporating succinyl hydroxamate and iminodiacetic acid hydroxamate moieties. J Enzyme Inhib Med Chem. 2003;18(3):233–42.

    Article  Google Scholar 

  175. Alhayek A, Khan ES, Schönauer E, et al. Inhibition of collagenase Q1 of Bacillus cereus as a novel antivirulence strategy for the treatment of skin-wound infections. Adv ther. 2022;5(3):2100222.

    Article  CAS  Google Scholar 

  176. Alhayek A, Abdelsamie AS, Schönauer E, et al. Discovery and characterization of synthesized and FDA-approved inhibitors of clostridial and bacillary collagenases. J Med Chem. 2022;65(19):12933–55.

    Article  CAS  Google Scholar 

  177. Schönauer E, Kany AM, Haupenthal J, et al. Discovery of a potent inhibitor class with high selectivity toward clostridial collagenases. J Am Chem Soc. 2017;139(36):12696–703.

    Article  Google Scholar 

  178. Voos K, Schönauer E, Alhayek A, et al. Phosphonate as a stable zinc-binding group for “Pathoblocker” inhibitors of clostridial collagenase H (ColH). ChemMedChem. 2021;16(8):1257–67.

    Article  CAS  Google Scholar 

  179. Senthilvelan T, Kanagaraj J, Mandal AB. Application of enzymes for dehairing of skins: cleaner leather processing. Clean Technol Environ Policy. 2012;14(5):889–97.

    Article  CAS  Google Scholar 

  180. Aravindhan R, Saravanabhavan S, Thanikaivelan P, et al. A chemo-enzymatic pathway leads towards zero discharge tanning. J Clean Prod. 2007;15(13–14):1217–27.

    Article  Google Scholar 

  181. Chen M, Jiang M, Chen M, et al. Approach towards safe and efficient enzymatic unhairing of bovine hides. J Am Leather Chem Assoc. 2018;113(02):59–64.

    CAS  Google Scholar 

  182. Li H, Zhou XW, Cheng HM. Compounding inhibitors based on Cu(II) Ions reduce the hydrolytic efficiency of collagen fibers on enzymatic unhairing process. J Leather Sci Eng. 2022;32(3):54–8.

    Google Scholar 

  183. DeLay K, Diao L, Nguyen HMT, et al. Successful treatment of residual curvature in Peyronie disease in men previously treated with intralesional collagenase Clostridium histolyticum. Urology. 2017;110:110–3.

    Article  Google Scholar 

  184. Honkanen RA, Kaplowitz K, Yung E, et al. Utility of purified collagenase (Xiaflex®) as a possible aid in glaucoma surgery: a pilot study. Invest Ophthalmol Vis Sci. 2016;57(12):2926–2926.

    Google Scholar 

  185. Zhang D, Zhang Y, Wang Z, et al. Target radiofrequency combined with collagenase chemonucleolysis in the treatment of lumbar intervertebral disc herniation. Int J Clin Exp Med. 2015;8(1):526–32.

    Google Scholar 

  186. Bae-Harboe YSC, Harboe-Schmidt JE, Graber E, et al. Collagenase followed by compression for the treatment of earlobe keloids. Dermatol Surg. 2014;40(5):519–24.

    Article  CAS  Google Scholar 

  187. Graham JJ, Bagai A, Wijeysundera H, et al. Collagenase to facilitate guidewire crossing in chronic total occlusion PCI-The Total Occlusion Study in Coronary Arteries-5 (TOSCA-5) trial. Catheter Cardiovasc Interv. 2022;99(4):1065–73.

    Article  Google Scholar 

  188. Brunengraber LN, Jayes FL, Leppert PC. Injectable Clostridium histolyticum collagenase as a potential treatment for uterine fibroids. Reprod Sci. 2014;21:1452–9.

    Article  CAS  Google Scholar 

  189. McCallon SK, Weir D, Lantis JC II. Optimizing wound bed preparation with collagenase enzymatic debridement. J Am Col Certif Wound Spec. 2014;6(1):14–23.

    Google Scholar 

  190. Engel CK, Pirard B, Schimanski S, et al. Structural basis for the highly selective inhibition of MMP-13. Chem Biol. 2005;12(2):181–9.

    Article  CAS  Google Scholar 

  191. Chen M, Jiang MF, Li H, et al. Screening of additives to reduce grain damage risk on unhairing by proteinase K. J Leather Sci Eng. 2020;2:18. https://doi.org/10.1186/s42825-020-00032-1.

    Article  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

The financial support of this study was from the National Natural Science Foundation of China (No. 22178231).

Author information

Authors and Affiliations

Authors

Contributions

Conceived and designed the review, and critical revised the manuscript: HMC; Wrote and revised the manuscript: SJW, XWZ and ZCJ. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Haiming Cheng.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

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

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wu, S., Zhou, X., Jin, Z. et al. Collagenases and their inhibitors: a review. Collagen & Leather 5, 19 (2023). https://doi.org/10.1186/s42825-023-00126-6

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s42825-023-00126-6

Keywords