Open access peer-reviewed chapter

Tendon Adhesion and Novel Solutions

Written By

Shen Liu, Qinglin Kang, Rui Zhang, Yanhao Li and Rong Bao

Submitted: 22 August 2022 Reviewed: 12 September 2022 Published: 03 November 2022

DOI: 10.5772/intechopen.108019

From the Edited Volume

Tendons - Trauma, Inflammation, Degeneration, and Treatment

Edited by Nahum Rosenberg

Chapter metrics overview

189 Chapter Downloads

View Full Metrics

Abstract

Tendon adhesion refers to the development of fibrotic tissue accumulation between injured tendon and the surrounding tissue, which usually happens as complications after surgical intervention for tendinopathies or traumatic rupture of tendon, resulting in undesired outcomes in the aspects of mechanical properties and functionality. Researches and understanding of tendon adhesion indicate that the process is related to the dominance of extrinsic tendon healing, with important factors such as inflammatory response, cell transference, certain growth factors, mistakenly stimulated signaling pathways and infection, and overdriving tendon remodeling. Taken the advantage of advanced material science and biochemistry, novel biomimetic materials have gradually emerged and been revealed to obtain satisfying antiadhesion capabilities. Taken the advantage of advanced material science and biochemistry, novel strategies, including hydrogels, nanoparticles, nanofibrous membranes, and substitutions for tendon and peritendinous apparatus, have gradually emerged and been revealed to obtain satisfying anti-adhesion capability solely or as drug delivery platforms. Although most of these results are currently limited in vitro or in animal models, future modification of these biosynthetic materials will help gain better mechanical properties and biocompatibility for clinical application. The establishment of next-generation delivery platforms against tendon adhesion requires the crosstalk among multiple fields.

Keywords

  • tendon adhesion
  • mechanism
  • countermeasure
  • advanced material
  • drug delivery system
  • future direction

1. Introduction

Tendons are dense connective tissue extending from muscles, which travel across joints to transmit force and produce motion. Although tendons possess remarkable tensile strength that can tolerate large force generated from muscle contraction, they are susceptible to damages caused by chronic overuse tendinopathies and traumatic rupture [1, 2]. It is noted that lacerated tendons cannot undergo spontaneous healing and surgical procedures are often required [3, 4, 5]. Conventional reconstruction techniques, including suturing, grafting, and synthetic prothesis replacement, however, are unfortunately associated with postoperative adhesion formation between surrounding tissue and the injured site, which results in undesired outcomes in the aspects of mechanical properties and range of motions.

With a better understanding of peritendinous adhesion, it is widely accepted that the process is related to the dominance of extrinsic tendon healing. To illustrate, the early inflammatory response and external fibroblast invasion are mainly responsible for the promoted tendon adhesion [6, 7, 8, 9]. Intrinsic factors of tendons, including low cellularity and metabolic activity, along with limited blood supply, also lead to a slow rate of tendon healing and remodeling, which increases tendency of re-rupture and hampering outcomes [10, 11].

Diverse strategies have been applied to overcome this clinical challenge. When anti-inflammatory drugs, antiadhesion growth factors, and certain genes that inhibit peritendinous adhesion have shown satisfactory outcomes in experimental studies, solely using these degradable molecules still face many limitations including swift inactivity, uncontrolled release, and toxicity before clinical translation. Besides, although physical barriers such as silk, silica gel and gold foil promote tendon gliding when wrapping around the injured site, with little biodegradability, these materials may inhibit intrinsic healing and cause body rejection, eventually leading to tendon necrosis and reoperation [12, 13]. Hence, current researchers have shifted their attention towards combinatory approaches, combatting tendon adhesion through loading different pharmaceutics and biologics into three-dimensional scaffolds, which were fabricated via various techniques including cryo-drying, solvent casting, gas foaming, braiding, and tissue engineering [14]. These scaffolds enable critical functions such as cell adhesion, proliferation, differentiation, and response to extracellular signals, while important issues still need to be discussed and improved. The scaffold should generally allow sufficient vascularization and interchange of nutrients and wastes, which is essential for tendon healing, while abnormal immune reactions should not be risen by the artificial scaffold [15, 16]. Additionally, adequate mechanical strength of the scaffold is also of vital importance to support tendon repair, mechanical load and gliding [17]. Besides, cargoes of the scaffolds should have promising effects on preventing peritendinous adhesion through alleviating inflammatory response, restricting unusual fibrogenesis, accelerating intrinsic healing, and promoting lubrication [18, 19]. In recent years, biosynthetic materials, including hydrogels, nanoparticles, nanofibrous membranes have gained wide attention for tendon adhesion prevention and have shown marvelous effects ex and in vivo [14, 20, 21]. At the same time, the advancement of biological materials and their derivations such as amnion, pericardium, and tendon sheath also shed light on strategies of tendon reconstruction and adhesion prevention [22, 23, 24, 25, 26, 27]. However, low accessibility of the resources, low productivity of the advanced biochemical products, and unknown biocompatibility to human body curbed their application to clinical usage [14, 20].

This chapter provides a comprehensive discussion of the current understanding of the mechanisms through which tendon adhesion is supposed to form, and identifies the pearls and pitfalls of the advanced biomaterials in preventing tendon adhesion.

Advertisement

2. Tendon structure and mechanisms of peritendinous adhesion

2.1 Tendon structure and composition

Tendon is composed of water (55–70% of whole tendon) and collagen (60–85% of dry weight) [28]. Type I collagen is the primary collagen in tendon, which accounts for 90% and the rest are type III, V and XI [29, 30]. Tendon also contains glycoproteins, cells and so on [31, 32].

Tendon is arranged orderly in a hierarchical manner which includes six levels: collagen molecule, pentafibril, collagen fibril, collagen fiber, fascicle and whole tendon. The basic unit of tendon is collagen molecule, five of which are bound together to form pentafibrils (also called microfibrils). Pentafibrils pack together to form collagen fibrils [33, 34].

In tendon, collagen fibrils are the unit of collagen fibers that aggregate to form fascicle with diameters ranging from 50 to 300 μm. A connective tissue, interfascicular matrix (IFM), is bound around the fascicles which is also the elemental structure of tendon. Moreover, the tendon is covered by the epitenon which is connected with IFM.

2.2 Tendon regeneration and repair

Tendon healing is a long period including three phases: inflammatory (days 1 to 7), fibroblastic (days 3 to 14), and remodeling (beyond day 10). Once tendon is injured, both external and internal cells are recruited and proliferated surrounding the injury site such as macrophage outside and tenocytes inside. After 3 days, the collagen is deposited to form extracellular matrix at the injury site, especially collagen type III. Then collagen type III is turned to be type I during the remodeling phase to heal the wounded tendon. However, the biomechanical strength of the healed tendon cannot reach as good as the one of the uninjured (Figure 1) [12, 35].

Figure 1.

(A) Peritendinous adhesion formation due to the unbalance between the intrinsic and extrinsic tendon healing, where antiadhesion materials are placed and function locally. (B) the key cellular and matrix changes and duration of three tendon injury repair phases (inflammation, reparation, and remodeling).

2.3 Factors affecting tendon adhesion

2.3.1 Inflammation

An acute inflammatory response to tendon rupture site is initiated lasting for 3 to 7 days [7, 8, 36]. In the initiation stage after tendon injury, the gene expression of pro-inflammatory cytokines significantly ascends attributed to recruitment of neutrophils, macrophages, and monocytes [37, 38, 39]. The inflammatory storm response to defect site after tendon injury is extensively considered as contribution to tendon adhesion formation and confusing matrix degradation, both of which are attributed to substantial up-regulation of inflammatory factors simulated by activated fibroblasts and matrix degradation [40, 41]. Besides, inflammation-mediated increased exudation and aggravation of fibrin leakage also lead to promotion of tendon adhesion formation. Therefore, spatially and temporally further understanding and adjustment to inflammatory response to tendon adhesion during the whole process of tendon healing will be far more crucial in inhibition of tendon adhesion formation.

2.3.2 Cell transference

Tendon healing includes the effect of both internal and external cells [42]. During the phase of inflammation, external cells play an important role in the adult tendon [43]. Neutrophils are recruited into the injury site in the first 24 h. Few minutes later, monocytes and macrophages reach, and macrophages become the dominant cell population instead of neutrophils after 24 h. Macrophages can be categorized into two main types: classically activated (M1) or alternatively activated (M2). Generally, M1 refers to the function of proinflammation, while M2 is considered an element of antiinflammatory response [44]. There are also other cell populations contacted with the phase of initial tendon injury such as T cells and mast cells [45, 46].

When it comes to the proliferative phase, the obvious character is the deposit of collagen type III in the injury site released by cells, especially fibroblasts. To begin with, it is the epitenon cells that proliferate. Both canine and murine models can be seen that the layer of epitenon become thicker in the early phase of proliferation [6, 47]. Epitenon cells can release more fibronectin than tendon itself, especially in the scar of injury site [9]. Nearly 2 days after injury, fibroblasts were recruited and proliferate rapidly, while the origin of fibroblasts remains unclear. Recently, researchers have identified a few different populations of resident tendon stem/progenitor cells involved in tendon healing, including cells from tendon fascicles, epitenon, and perivascular cells [48, 49, 50, 51, 52, 53].

After 10 days, there are fewer cells in the scar which is called the remodeling phase. Generally, the scar is finally replaced by organ-specific cell populations such as bones. However, scar in the tendon is partly replaced. The decrease of scleraxis basic helix–loop–helix transcription factor positive tendon cells may attribute to this phenomenon [54]. Moreover, the external cells such as inflammatory cells and myofibroblasts also play an essential role, but the explicit mechanism remains to discover [55, 56].

2.3.3 Growth factors

Generally, it is considered that GFs mainly work on proliferation phase overlapping the inflammation phase [7]. GFs, including transforming growth factor-beta (TGF-β), vascular endothelial growth factor (VEGF), basic fibroblast growth factor (bFGF), platelet-derived growth factor (PDGF), bone morphogenetic proteins-12, -13, and -14 (BMPs) also known as growth and differentiation factors-5, -6, and -7 (GDFs) respectively and insulin-like growth factor-1 (IGF-1) has been extensively authenticated their existence during different phases of tendon healing and regeneration process, and their significant roles have been widely studied [21, 57, 58, 59, 60]. TGF-β1 is an isoform of TGF-β with multifunction, which is widely demonstrated as main cytokine for tendon adhesion formation [61, 62]. VEGF, bFGF, and PDGF have been demonstrated to promote tendon healing by intensifying mesenchymal stem cells (MSCs) proliferation and differentiation to tenogenic lineages, increasing vascularization and ascending biomechanical strength after tendon regeneration [63, 64, 65, 66]. BMPs have been shown great induction force to tenogenic differentiation of MSCs as well as necessity to the regeneration of tenocytes, especially BMP-14 has been utilized to initiate tenogenic lineage of adipose-derived stem cells (ASCs), and it is widely demonstrated its high efficiency of pro-differentiation when associated with ASCs [67, 68, 69, 70, 71, 72]. However, BMP-14 has also been reported its an additional benefit of tendon adhesion resistance during tendon healing [58, 73]. Furthermore, whether the final production of BMPs activation is tenogenic or osteogenic lineage relies not only on the isoform of BMPs but also the type of biomechanical stimulation [74, 75]. IGF-1 has been proved to stimulate regeneration and tenogenic lineage differentiation of ASCs, however, the recent question regarding the practical usage of IGF-1 in tendon healing is whether IGF-1 acts independently or cooperates under the guideline of growth factors for single use of IGF-1 presented abominable outcomes in tendinopathic human patellar tendon [76, 77, 78, 79]. Therefore, in-depth understanding of the multifarious roles of GFs not only by themselves but also in synergy with others will help us better understand tendon adhesion formation and anchor efficient therapeutic targets.

2.3.4 Signaling pathways

Inflammation, cell recruitment, and growth factors provide new insights of mechanism of tendon adhesion formation, and these factors may produce marked effects under the specific guidelines of signaling transduction process concatenating above three. After tendon injury, the tendon biomechanical changes including but not limited to loss of collagen fiber tension, primary cilium deformation and nuclear deformation initiate the pathology reaction by accumulation of specific molecular messengers such as ions or cytokines [80]. Inflammation is the primary pathology reaction raised up by a certain signaling pathway. The nuclear factor-kappa B (NF-κB) signaling pathway is involved at the initial stage of inflammation during the tendon healing. It is located in the cytoplasm and enters the nucleus after activation playing a role of transcription factor. In the early stage of tendon injury, accumulated cytokines, including tumor necrosis factor-α (TNF-α), interleukin-1β and -6 (ILs) act on the surface of tendon cells by Toll-like receptor, which activates NF-κB and then up regulates the expression of the above inflammatory factors to form positive feedback, amplifying inflammatory effect [81, 82]. Abraham et al. have found that inhibition of NF-κB signaling pathway by blocking I-kappaB kinase beta could mitigate tendinopathy development [83].

The TGF-β signaling pathway is widely studied in the pathological process of tendon adhesion formation. Smads proteins 2 and 3 (SMAD2/3) are molecules acting as a transcription factor of TGF-β as well as a signal transducer in the TGF-β pathway [84, 85]. The activated TGF-β acts on the fibroblasts and then phosphorylates the Smad2/3 protein in the cytoplasm to regulate the expression of target genes promoting fibroblasts proliferation and differentiation into myofibroblasts, so as to promote collagen secretion [86, 87, 88, 89]. Down-regulate expression of TGF-β may significantly inhibit tendon adhesion formation. Wu el al. designed a three-dimensional tendon scaffold loading with TGF-β small interfering RNA (siRNA) plasmid and proved its satisfactory efficiency on prevention of tendon adhesion formation as well as promotion of tendon function repair [90]. Interestingly, after activated TGF-β binding to its receptor, extracellular signal-regulated kinase 1 and 2 (ERK1/2) which belongs to mitogen-activated protein kinases (MAPKs) pathways are also phosphorylated simultaneously to act on the binding sites of SMAD2/3, indirectly enhancing TGF-β/SMAD2/3 signaling pathway to promote exogenous fibroblast proliferation and collagen synthesis [86, 91, 92]. It indicated that the TGF-β signaling pathway could work in conjunction with other pathways like the MAPK pathway or BMP pathway [93].

Matrix metalloproteinase (MMP) is a protease family with metal ions as cofactors. TGF-β can induce the expression of plasminogen activator inhibitor-1 (PAL-1) in tenocytes to accelerate the degradation of plasmin and its mediated MMP-2, leading to excessive deposition of extracellular matrix (ECM) and type I collagen [94, 95, 96]. Lu et al. [97] confirmed that MMP comes from bone marrow cells that migrate to the injury site significantly enhance regeneration of tendon as well as tendon adhesion formation. Cai et al. [98] constructed a macrophage reactive siMMP hydrogel for high efficient synergistic prevention of tendon adhesion formation.

Besides, cyclooxygenase-2 (COX-2)/prostaglandin E (PGE)/prostaglandin type 4 receptor (EP4) signal transduction pathway also promote the tendon adhesion formation. It is found that the content of COX2 increases during tendon healing process, which can catalyze the decomposition of arachidonic acid into PGE acting on EP4 located in cell membrane to contribute to synthesis and accumulation of ECM [99, 100, 101]. Therefore, inhibition of COX2/PGE/EP4 signaling pathway may reduce the formation of tendon adhesion. However, some researches have demonstrated that the inhibition of COX-2 by high-dose utilization of non-steroidal anti-inflammatory drugs (NSAIDs) can increase the apoptosis of tenocytes recruited to defect sites, which is not conducive to tendon healing [102, 103, 104]. Furthermore, systemic application of EP4 inhibitor can increase the infiltration of macrophages and the secretion of type 1 collagen aggravating the degree of tendon adhesion [105]. In conclusion, the role of COX-2/PGE/EP4 signaling pathway in tendon adhesion formation is complex indicating more requirements for further studies on mechanism and more cautious usage of NSAIDs.

Altogether, it should be noted that tendon adhesion is a complex pathology process under the guidelines of multiple signaling pathways that interact with each other participating with various cytokines, growth factors and cells. Although studies on single regulation or recently synergistic regulation of signaling pathways have shown satisfactory outcomes in tendon adhesion prevention and tendon healing promotion, fully understanding on how exactly these signaling transduction pathways interact is on great demand.

2.3.5 Infections

Infection is mostly caused by a significant degree of contamination during the initial tendon injury [2, 7, 106]. The infection rate and severity depend on where the injury happens and how it is caused [107, 108]. And a review paper in 2018 reported the most common bacterial populations causing tendon adhesion formation even some devastating effects like gangrenosis. Therefore, it should be highly noted that infection which can mediate tendon adhesion formation should be completely forbidden, and recent study shows that it is capable to integrate an antimicrobial biomaterial. Shalumon et al. [109] constructed a multifunctional electrospun nanofiber membrane to perform excellent anti-infection effect as well as prolonged prevention of inflammation and tendon adhesion formation.

Advertisement

3. Tranditional strategies against tendon adhesion

Surgical intervention is usually unavoidable because lacerated tendon cannot repair by themselves and may retract and have remarkable defects after injury. Meanwhile, simultaneous injuries of adjacent skin, nerves, vessels, and bones also require surgical repair.

3.1 Intraoperative repair methods and tendon adhesion

Modified Kessler suture technique remains globally accepted method for flexor tendon repair with reliable mechanical strength and less peritendinous adhesion [3]. However, even if different modifications for tendon suture have been studied for decades, peritendinous adhesion often occurs and reoperation is still required due to undesired scar tissue formation [4, 110]. When it comes to tendon defect reconstruction, adhesion mechanism of autologous tenograft is believed to involve both intrinsic tenocyte necrosis and extrinsic fibrogenetic and inflammatory cell invasion, while mild peritendinous adhesion was observed in decellularized tendon allograft transplantation since the intrinsic mechanism was forbidden [106, 111]. Rather than adhesion formation, xenograft transplantation for tendon reconstruction may raise another important issue, high postoperative infection rate [112].

Physical barriers made of non-degradable materials including silica gel and gold foil have been previously applied intraoperatively to reduce peritendinous adhesion by wrapping the injured site [113]. However, these barriers are outdated and clinical application are eliminating since their non-degradability and non-permiability may prevent substance exchange and eventual tendon necrosis [24].

3.2 Pharmaceutic intervention

The use of anti-inflammatory drugs against tendon adhesion dates back to the 1980s when NSAIDs were simply injected to the injured area to reduce expression of pro-inflammatory factors that might promote scar tissue formation [114]. However, detrimental side effects of these drugs on cardiovascular and genitourinary systems sometimes occur when they are excessively used [115]. The effect of local steroid administration has been documented by several earlier studies as a dose-related decreased fibrogenesis, collagenesis, adhesion and tensile strength of the repaired tendon [116]. However, a recent animal study has also reported increased risk of peritendinous adhesion [117]. In the contrast, the controlled release of steroid by entrapping them with synthetic polymers showed promising antiadhesion and antiinflammatory effects in recent years [118]. To date, there is no consensus on whether steroid application should be a standard therapy for tendon adhesion.

Advertisement

4. Advanced materials against tendon adhesion

4.1 Basic characteristics

Taking advantage of advanced knowledge of materials, several polymeric or biogenetic materials have been studied and exploited as alternatives for conventional tendon restore strategies [12, 20, 119, 120]. Except for antiadhesion functionality, a few characteristics should be acquired for such materials, which include biodegradability, biocompatibility, accessibility with proper architecture, and reliable mechanical properties [21].

Polymer implantation materials are not supposed to permanently stay in human body, indicating that these synthetic materials should be biodegradable to avoid certain side effects, such as rejection reaction. The degradation should not interfere with the mechanical properties of the materials until safe gliding can be performed by the injured tendon, and the by-product should not be toxic and can be eventually eliminated by the human body [121]. The degradability of these polymers has been proven to be related with molar mass, crystallinity, and mechanical loading [122, 123, 124]. On the other hand, biogenetic materials against tendon adhesion, such as amnion, autograft membranes and allograft tendon sheath and pericardium, as well as their derivations, once show ability to replace natural tendon sheath, should otherwise tolerate biodegradation in order to perform as eternal sheath instead.

Biocompatibility is described as nontoxic, noncarcinogenic, nonthrombogenic, nonimmunogenic characteristics, and proper response to the host of implantation materials [125]. For acceptable biocompatibility of polymeric materials, in vitro cell seeding test should announce a cell viability over 70% [126]. The structural architecture of the implantation materials should manipulate the process of antiadhesion by allowing exchange of information and material, including growth factors, drugs, hormones and degradation by-products. Also, sufficient porosity is required in order to promote neovascularization [127, 128, 129]. Synthetic strategies determine not only the processability of the material but also the mechanical properties. Thus, it is of importance that the processing technique should be easy and cost-efficient for clinical application, and mechanical strength of the construct should be able to tolerate certain impact, e.g. tensile tests, to identify its biomechanical stability [130].

4.2 Synthetic materials against tendon adhesion

4.2.1 Hydrogel

Hydrogels represent a bunch of polymers crosslinking hydrophobic groups and hydrophilic residues with large water content, high porosity and similarity to extracellular environment. The hydrophobic outer layer stands for a physical barrier to isolate exogenous inflammatory response and inhibit fibroblast and macrophage migration during tendon repair, while the hydrophilic inner layer mimics the inner side of tendon sheath to direct and lubricate tendon gliding [131]. Topographic structure of hydrogel allows for carrying certain therapeutic agents to promote tendon healing and inhibit tendon adhesion. By controlling raw material concentration and crosslinking levels, appropriate hydrogel can be obtained with desired mechanical properties and degradation rate (Figure 2) [119, 131].

Figure 2.

Basic design considerations of therapeutic platforms for prevention of peritendinous adhesion including fabrication technologies, scaffold materials, therapeutic structures, and drug-loading.

Hyaluronic acid (HA) is a kind of hydrophilic polysaccharide component of natural synovial fluid. With negative charge, its antiadhesion ability was confirmed by inhibiting fibroblast proliferation and migration [18]. Besides, HA exhibits an essential source of nutrition and lubrication for tendon gliding and repair, with strong potential in eliminating harmful inflammatory factors [132]. Hundreds of injectable hydrogel-based materials have been developed with the participation of HA in order to prevent peritendinous adhesion, among which Seprafilm, consisting of carboxymethylcellulose and HA, and xanthan gum/gellan gum/hyaluronan hydrogel showed both biocompatibility and antiadhesion efficacy in vitro and in vivo [133, 134]. Phospholipid-based hydrogels possess considerable biocompatibility in antiadhesion application as they do not induce foreign body reaction or cause protein conformational changes [135, 136]. Controllable degradation rate of such hydrogels can be achieved by altering the rate of coordinating compound [136]. Injectable hydrogels show great potential in inhibiting tendon adhesion because of its convenient, cost-effective, and minimally invasive manner in postsurgery rehabilitation, and the potential can be amplified when cross-linked with functional molecules. By integrating thermo-responsive material with chitosan and HA, liquid hydrogel will transfer to gel state and suit the tendon defect once body temperature is reached, and then chitosan plays its cytostatic role locally by suppressing fibroblast growth and infiltration [137, 138]. Solely applying chemosynthetic hydrogel remains less satisfactory due to nonspecific blocking adhesion as physical barrier. Therefore, developing drug-delivery platform with formulated hydrogel becomes an alternative method for inhibiting peritendinous adhesion. Cyclooxygenase-engineered miRNA plasmid-loaded polymer nanoparticles has been designed to be encapsulated into HA hydrogel to extend plasmid release against oxidative stress and adheasion formation during tendon repair [139]. Besides, appropriate 5-fluorouracil loading of hydrogel was proved to suppress fibroblast proliferation and migration. Despite the fact that these agent-hydrogel crosslinking materials have shown remarkable ability in the aspect of antiadhesion, a lack of proper mechanical properties still challenges their application in mimicking natural tendon sheath tissue, which promotes proliferation and tenogenic differentiation of stem cells by possessing suitable tensile loading and elasticity [17, 140].

4.2.2 Nanoparticles

Another important therapeutic agent delivery vehicle system is various nanoparticles (NPs). Compared with large molecules, NPs have shown better delivery efficacy and biocompatibility in delivering growth factors, genes and drugs by easier internalization of cells [141]. Quicker escaping from endosomes of the NPs also eliminates cargo biodegradation and thus prolongs bioactive cargo release [142].

Multiple growth factors have shown essential roles during tendon repair, as VEGF enhances neovascularization and accelerates healing process, PDGF promotes tendon gliding, and bFGF induces the differentiation of MSCs towards tenogenic linkage [143, 144]. NPs have been reported to serve as nonviral vectors for growth factors genes delivery, such as bFGF and VEGF, to induce overexpression in tenocytes of lacerated tendon, and downstream macromorphological effects, including enhanced tendon mechanical strength and tendon gliding, were observed in vivo, indicating satisfactory antiadhesion function of the transfected genes [66]. Although the NPs system has been proved to restrain adhesion by loading TGF-β miRNA plasmid for weeks, and simplify the procedure by minimally invasive injection, the absence of rigid mechanical strength of these particles forbids them from acting as physical barriers to separate ruptured tendon from surrounding tissue and migrating inflammatory cells [24]. Similar obstacle was also faced by antiadhesion nanoparticle-coated suture, as no proper timberland could separate the extrinsic healing and cell invasion [145]. New insights of employing self-healing HA hydrogel with NPs entrapping antiadhesion molecules may be reliable solutions to avoid unexpected artificial physical barrier rupture during tendon gliding. Cai et al. [98] reported successful inhibition of fibroblast proliferation and peritendinous adhesion in murine model with self-healing HA hydrogel loading Smad3-siRNA nanoparticles.

4.2.3 Nanofibrous membranes

Nanofibrous membranes (NFMs) have been approved as both reliable antiadhesion barriers and effective carriers of pharmaceutical agents due to their functional characteristics in the aspects of systematically and locally delivering medication, stem cells, components of ECM and genes, and being as physical goalkeepers inhibiting adhesion related to external tendon healing [127, 128, 129, 146, 147].

Among multiple methods to produce such NFMs, electrospinning remains one of the most popular ones for the fabrication of different nanofibers with diverse biomedical applications. Manufactured with this convenient and robust technique, the electrospun scaffolds share similar characteristics with natural ECM including topography and high porosity, and their mechanical properties can be easily modulated by altering the fiber alignment and diameter, and by manipulating the viscosity and volatility of the solution, applied voltage, flow rate of each polymer, along with the distance between the capillary and the collectors.

The electrospun NFMs should defend their payload against rapid degradation and permit release of the drugs or molecules in desired patterns for their antiadhesion applications. By controlling the material composition, drug-encapsulation technology and the architecture of the NFMs, this delivery system has been modified in order to optimizing its pharmacodynamics.

A variety of techniques, including surface modification, blending, coaxial and emulsion electrospinning, have been employed in order to encapsulate pharmaceutic agents into these nanofibers [148]. By physically and chemically altering the surface of the nanofibers with biomolecules, the surface modification technique is usually applied for vulnerable agents such as nucleic acids, proteins, growth factors and polysaccharides, which possess swift biodegradation rate and may lose their functionality during the process of electrospinning [149, 150, 151, 152, 153, 154]. The rest three techniques were then established to accomplish gradual release of the therapeutic molecules. The blending technique requires the drugs or molecules to be dissolved in a polymeric solution before electrospinning, in which process the compatibility of the polymer and the solvend depends on the wettability to provide appropriate drug solubility and distribution [148, 150]. Coaxial electrospinning refers to the modification of traditional electrospinning process by concentrically locating the therapeutic cargoes during nanofiber fabrication, which will protect the biomolecules from environmental risks and attenuate drug degradation thus extending release period [155, 156]. Another effective method to prolong the drug release period is emulsion electrospinning, through which the organic solvent evaporates faster than the aqueous phase in which dissolves the molecular cargoes, leading to central migration of the biomolecules [157, 158, 159, 160]. Besides, sequential electrospinning technique has been widely utilized to produce multi-layered NFMs which combine hydrophilicity or hydrophobia as well as mechanical strength and permeability of each layer of polymer, lubricate the tendons wrapped inside, and protect the longevity of the delivered molecules in order to release them during desired periods [161, 162, 163, 164].

A wide spectrum of electrospun NFMs with diverse properties has been employed as flexible delivery platforms for pharmaceutic agents as well as nanoparticles to inhibit peritendinous adhesion. Despite alleviating tendon adhesion, the rapid clearance of NSAIDs and side effects on tendon repair limit their application for tendon healing and adhesion prevention. By loading ibuprofen (IBU) to poly (l-lactic acid)- polyethylene glycol (PELA) NFMs, Liu et al. [128] revealed that the delivery system could decrease peritendinous adhesion and kinetics of drug release was proved to be mostly dependent on its diffusion and polymer matrix degradation. Besides, more effective blocking of cell adhesion/proliferation and inflammation could be achieved by incorporating low content of polyethylene glycol (PEG) with PELA nanofibers [128]. To reduce postsurgical peritendinous adhesion with long-lasting release of NSAIDs, modified mesoporous silica (MMS) nanoparticles loading IBU was prepared and encapsulated within poly (l-lactic acid) (PLLA) nanofibers with emulsion electrospinning technique [165, 166, 167]. This IBU-MMS-PLLA drug delivery scaffold sustained release of drug by entrapping IBU within the porous MMS particles and its antiadhesion and antiinflammation functionality was observed even 8 weeks postoperatively. HA represents a widely used agent in preventing postsurgical adhesion and permitting tendon gliding due to its features of synovial fluid. By coating poly (ε-caprolactone) (PCL) with HA from the inner side, a bi-layer PCL/HA-PCL NFM has been reported to mimic tendon sheath as the outer layer PCL reduced cell adhesion/invasion and pro-inflammatory cytokine penetration, while the inner layer of HA lubricated the tendon and promoted gliding and healing [146]. Chitosan is another well-known polymer that can enhance the mechanical strength of NFMs when hybridized with other polymers, such as PCL, but the effect of chitosan-based NFMs against peritendinous adhesion remains controversial [147]. Although the hydrophilic nature of chitosan may partially reduce surface adhesion, other factors, including electric charge and roughness of the NFM also influences fibroblast attachment and recent studies have confirmed that PCL/chitosan NFMs had little impact on decreasing peritendinous adhesion [19, 168, 169, 170, 171]. Meanwhile, the efficacy of loading certain particles to HA and then merging the HA-particle cargo to NFMs has also been well studied in terms of antiadhesion. Although mitomycin (MMC)-induced fibroblast apoptosis can inhibit collagen synthesis and is supposed to reduce tendon adhesion, misguided use of MMC may result in local and systematic toxicity [172]. By wrapping MMC in HA hydrogel and encapsulating the composite particle in PLLA nanofibers, notably controlled release of MMC was achieved and inhibition of tendon adhesion was observed in vivo [159]. Postsurgical infection also presents and important risk factor for peritendinous adhesion due to inflammation cascade and fibroblast infiltration [109]. Since silver (Ag) and its derived nanoparticles have shown strong antibacterial properties, NFMs with cored HA and Ag nanoparticles embedded in PCL shell have been investigated to possess adhesion prevention functionality [173]. Controlled release of the cored HA could promote tendon gliding and reduce fibroblast attachment, while the Ag nanoparticle-loaded NFM sheath with thinner diameter showed better capability of preventing fibroblast attachment in vitro and inhibiting peritendinous adhesion in vivo [174].

Incorporating biologics, including stem cells and growth factors, with NFMs scaffolds has recently gained significant attention in the field of tendon repair and antiadhesion [175]. Researchers have also attempted to load growth factors into nanofibers to exploit their antiinflammation and antiadhesion bioactivities, but the growth factors degrade easily when exposed to in vivo conditions and may be damaged during electrospinning preparation [65, 143]. Thus, similar to the afore-mentioned drug-nanoparticle-nanofiber delivery system, fabricating NFMs carrying growth factor-loaded nanoparticles may result in satisfying application of growth factors to prevent peritendinous adhesion. For instance, bFGF-loaded dextran glassy nanoparticles (DGNs) have been reported to contain higher proportion, less burst release and better-preserved bioactivity of bFGF when pre-formulated and encapsulated with PLLA nanofibers using the emulsion electrospinning technique, compared with bFGF-loaded PLLA, and in vivo investigations revealed that the bFGF-DGNs-PLLA NFM scaffold not only promoted tendon healing but also mediated collagen remodeling in ECM and reduced tendon adhesion [176]. However, few attempts were taken to eliminate tendon adhesion by delivering stem cells with NFM scaffold. Liao et al. [177] found no significant reduction in tendon adhesion, despite the delivery of MSCs with minimal burst release using bi-layered HA/copolyermized l-lactide and ε-caprolactone NFMs, indicating a plenty of scope and opportunities for this approach.

Given the fact that inhibiting crucial cellular signaling pathways would reduce formation of adhesion, NFMs loading exogenous genetic materials have been developed to alter expression of certain genes related to peritendinous adhesion [178]. Small interfering RNA downregulating expression of ERK2 and SMAD2/3 has been shown to prevent fibroblast proliferation as well as abnormal collagen accumulation [179]. To employ such functionality, a pyridinedicarboxaldehyde-polyethylenimine (PDA)-mediated exogenous ERK2-siRNA delivery system based on PLLA/hyaluronan (P/H) nanofiber scaffold has been developed in order to gradually inhibit ERK2 bioactivity and reduce peritendinous adhesion [178]. In vivo model demonstrated that ERK2-siRNA-PDA-P/H delivery system could reduce type III collagen density through down-regulating ERK2 and SMAD3 gene expression, thus inhibiting tendon adhesion (Table 1).

ResearchersType of materialSource of materialMolecule loadingManufactureMechanical strengthCell lineIn vitro resultsAnimal modelIn vivo resultsPublication yearPublication title
Liu et al.biomimetic bilayer sheath membranePCL, HAHAcombination of sequential and microgel electrospinning technologiesbilayer of
PCL-HA0%:2.13 ± 0.25 MPa
PCL-HA4%:1.91 ± 0.21 MPa
PCLHA8%:1.77 ± 0.18 MPa
PCL-HA12%:1.55 ± 0.21 MPa
multipotent C3H10T1/2cellsMore cells adhered to and were better distributed on the surface of the inner HA-loaded PCL layer and the tissue culture plate surfacesflexor tendon of Leghorn chickensThe average scores of adhesions and tendon healing in the biomimetic bilayer sheath membrane group were lower than the other groups2012Biomimetic sheath membrane via electrospinning for antiadhesion of the repaired tendon
Liu et al.Electrospun fibrous membranesPELAIbuprofenelectrospinningMaximum tensile strength (MPa) PELA-0% 3.72 ± 0.32, PELA-2% 3.42 ± 0.36, PELA-6% 3.13 ± 0.38, PELA-10% 2.89 ± 0.31L929 mouse fibroblastFewer cells adhered to the ibuprofen-loaded PELA fibersflexor digitorum profundus tendons of Leghorn chickensNo peritendinous adhesions were detected in most tendons treated with the ibuprofen-loaded PELA fibrous membrane2013Prevention of Peritendinous Adhesions with Electrospun Ibuprofen-Loaded Poly(L-Lactic Acid)-Polyethylene Glycol Fibrous Membranes
Liu et al.Electrospun fibrous membranesPLLAdextran glassy nanoparticles (DGNs) loaded with bFGFelectrospinningTensile strength(MPA) PLLA 4.38 ± 0.33, bFGF PLLA 4.38 ± 0.33 and bFGF/DGNs-PLLA fibrous membranes 3.54 ± 0.25multipotent C3H10T½ (C3) cellsThe PLLA membrane showed an anti-adhesion effect, more cells adhered to and were better distributed on the surface of the bFGF/DGNs-PLLA membraneAchilles tendon of male SD ratsIn the bFGF/DGNs-PLLA group, A significantly increased tendon thickness was observed, tendon healing and vascular prominence and angiogenesis were significantly better, and an increase in breaking force was detected but still lower than control group.2013Tendon healing and anti-adhesion properties of electrospun fibrous membranes containing bFGF loaded nanoparticles
Kuo et al.Hydrogelxanthan gum/gellan gum/hyaluronan
hydrogel
/Blending///Achilles tendon of SD ratsHelp reduce the incidence of postoperative tendon adhesion, and enable tendon healing and preserve the mechanical strength as effectively as Seprafilm.2014Evaluation of the ability of xanthan gum gellan gum hyaluronan hydrogel membranes to prevent the adhesion of postrepaired tendons
Chen et al.nanofibrous membranechitosan-grafted PCL/electrospinningUltimate tensile strength (MPa) PCL1.4 ± 0.1, PCL-g-CS 2.2 ± 0.5Human foreskin fibroblast (Hs68) cellsThe inhibition of cell migration by PCL-g-CS NFMs was evident, and the CS layer on the PCL-g-CS NFM can prevent more nonspecific cell adhesion.flexor digitorum profundus of New Zealand white rabbitsIn the tendons treated with the PCL-g-CS NFM, no adhesion was observed, There was a statistical improvement in the DIP joint flexion angle, PIP joint flexion angle and sliding excursion. And tendons treated with PCL-g-CS NFM required the least pull-out force and showed the lowest degree of stiffness.2014Prevention of peritendinous adhesions with electrospun chitosan-grafted polycaprolactone nanofibrous membranes
Chen et al.nanofibrous membraneHA, PCL, PEOAgelectrospinninghuman foreskin fibroblasts (Hs68)Compared to other NFMs, the lowest number of cells and the least cell spreading were found on the PCL/HA + Ag NFMs.flexor tendon of New Zealand white rabbitsThe surface of the repaired tendon was smooth, and no adhesion was observed between the repaired tendon and the peritendinous tissue. No adhesions were observed between the repaired tendon and the surrounding tissue in the HA/PCL + Ag NFM treatment
group
2015Dual functional core-sheath electrospun hyaluronic acid/polycaprolactone nanofibrous membranes embedded with silver nanoparticles for prevention of peritendinous adhesion
Tang et al.NanoparticlePEI-modified PLGAAAV2-TGF-β1-miRNAsolvent evaporation/Rabbit tenocytessignificantly improved the expression of the gene of type I collagenflexor digitorum profundus tendons of chickensGene therapy through AAV2 vectors is efficient to deliver growth factor genes to the healing tendon and reduce adhesion formations, but reduces in tendon healing strength2016Gene therapy strategies to improve strength and quality of flexor tendon healing
Liu et al.Electrospun nanofibers menbranePLAIbuprofenBlending/RAW264.7 macrophagesLess RAW264.7 adhered to the surface of IBU/PLA-M than to the surface of PLA-M.flexor digitorum profundus tendons of Leghorn chickensAlthough granuloma formation was investigated in the IBU/PLA-M group, a clear space around the tendon could usually be observed. The number of α-SMA positive vessels in the IBU/PLA-M group is significantly lower than that in the PLA-M group.2017Macrophage infiltration of electrospun polyester fibers
Zhou et al.Hydrogelthiol-modified hyaluronic acid/PEG-diacrylateCOX-1 and COX-2 miRNA plasmidsBlendingThe elastic modulus, yield strength, and elongation to yield for hydrogel are 0.121 ± 0.01 MPa, 0.014 ± 0.003 MPa, and 3.50% ± 0.5%, respectively.flexor digitorum profundus tendons of white Leghorn chickensThis hydrogel could effectively reduce the expression of COX-1 and COX-2 proteins in the tendons and subcutaneous tissues.2018Localized delivery of miRNAs targets cyclooxygenases and reduces flexor tendon adhesions
Chen et al.Micro-hydrogel-generated asymmetric scaffoldchitosantendon stem/progenitor cellsSelf-deposition technique641.61 ± 12.43 MPatendon stem/progenitor cells sided onto the scaffolddisplayed higher levels of tenogenic specific genes expression and protein production.Achilles tendon of SD ratssynergistic effect
on tendon regeneration and yielded better-aligned collagen fibers with elongated, spindle-shaped
cells.
2018An asymmetric chitosan scaffold for tendon tissue engineering
Jayasree et alElectrospun menbranePCL-Collagen-bFGFPCL, collagen bFGF nanofiberelectrospinning89.4 ± 5.3 MPaRabbit tenocyteUpon dynamic stimulation, mPCL-nCol-bFGF-DS scaffolds showed significantly higher expression of tenascin C, collagen I, biglycan, and fibronectin.Achilles tendon of New Zealand white rabbitsThe alignment of collagen was highly in comparison to native tendon which showed perfectly aligned fiber morphology, whereas, by 12 weeks, the implants showed more aligned nature of collagen fibers which was further confirmed by MT staining.2019Bioengineered Braided Micro-Nano (Multiscale) Fibrous Scaffolds for Tendon Reconstruction
Park et al.filmcross-linked electrospun cartilage acellular matrix (CAM)CAM, PLGAcross-linking25.06 ± 1.4 NL929 mouse fibroblast cellsCell migration
from serum-free medium area towards serum containing medium area was inhibited in the CX-CAM film group
Achilles tendon of New Zealand White rabbitsThe degree of adhesion in histology was highest in the repair group, followed by Seprafilm, CX-CAM
film, and sham group.
2020Cross-linked cartilage acellular matrix film decreases postsurgical peritendinous adhesions
Song et al.Electrospun menbranePCLmechano-growth factorElectrospinning/RAW264.7 mouse macrophagesAntiinflammatory macrophage phenotype polarizationAchilles tendon of SD ratsAlmost no adhesion can be detected in the MGF-modified group with a sheath space formed between tendon and scaffold.2021Surface modification of electrospun fibers with mechano-growth factor for mitigating the foreign-body reaction
Chen et al.Core-sheath nanofiber membranePLA, HA,Ag/Tn/TkBlending and core-shell/NIH/3 T3 mouse embryonic fibroblastsA synergistic effect was found by combining Ag NPs with HA as Tn+, with Ag NPs embedded in the thin sheath and the highest amount of released HA, which demonstrated the least staining of vinculin by inhibiting cell adhesion as well as suppressing cell spreading.flexor digitorum profundus tendons of New Zealand White rabbitsThe Tn + group showed a nearly complete lack of adhesion and demonstrated the most regular collagen arrangement.2021Functional Hyaluronic Acid-Polylactic Acid/Silver Nanoparticles Core-Sheath Nanofiber Membranes for Prevention of PostOperative Tendon Adhesion

Table 1.

Recent research on biomaterials for the prevention of peritendinous adhesion.

Source: [128, 139, 146, 173, 174, 176, 180, 181, 182, 183, 184, 185]

4.3 Biomimetic materials against tendon adhesion

4.3.1 Amnion and its deviations

Human amniotic membrane (HAM) tissue is inexpensive, easily stored, of minimal antigenicity and with low immunogenic rejection, and has shown advanced potential in preventing postsurgical adhesion [186]. Recent findings also revealed its reinforcement on mechanical strength after combined with modified Kessler suture in flexor tendon repair, probably due to the expression of multiple growth factors that promotes tenocyte proliferation [187, 188]. The decellularized HAM can physically inhibit cell infiltration and preserve tendon gliding through “tunnel effect”, and biochemically suppress ILs-induced immunologic cascade to alleviate systematic inflammatory response and local abnormal collagen synthesis [189, 190, 191, 192, 193].

Typical strategy to apply HAM and other films to the injured site of tendon generally depends on wrapping and suture. However, excessive suture burden may also result in consequent fibrogenesis and adhesion, stimulated by the suture as foreign body. To reduce such complications, several modification techniques has been employed for HAM modification. Photochemical tissue bonding (PTB) technique requires photo-active substance, e.g. Rose Bengal (RB), to conglutinate tissues through illumination [194]. Ding et al. [27] found that by immersing decellularized HAM in 0.1% (w/v) RB, attaching HAM to repaired chicken tendon became suture-less and significantly reduced inflammatory cell chemotaxis and better joint performance were observed. Wrapping freeze-dried HAM around the sutured site of lacerated tendon may also result in moderate peritendinous adhesion because of the uncontrolled expression and emission of TGF-β, which promotes exogenous fibroblast migration and collagen synthesis through ERK/SMAD pathway [24]. Coating PCL on both surfaces of the HAM by electrospinning technique has been identified in a rabbit model to reduce such adhesion by (1) prevent outer fibroblast and inflammatory infiltration, (2) gradual release of multiple growth factors, e.g. PDGF, VEGF, and TGF-β from the intermediate HAM layer, and (3) maintaining tendon gliding inside the membrane, thanks to the appropriate porous network of the PCL NFMs [195].

HAM and its deviations have been gradually utilized in preventing peritendinous adhesion during clinical practice. Compared with synthesized polymer membrane and control group, HAM has been reported to significantly attenuated complication rate of erythema, exudate and rupture in repaired Zone II human flexor tendons, but there was no difference in ultimate interphalangeal joint range of motion [196]. In addition, local administration of HAM wrapping around injured human flexor tendon was observed to reduce serum level of IL-6 and TGF-β1, indicating a systematic antiinflammation effect during tendon repair, thus preventing tendon adhesion [190]. On the other hand, effect of HAM allograft against tendon adhesion, tenolysis and joint complications after application in flexor tendon repair was questioned by Leppänen et al. [197] since half of the 10 patients enrolled developed these complications.

4.3.2 Tendon and sheath graft and reconstruction

Two-stage flexor tendon defect reconstruction remains gold standard [198]. The technique requires first-stage silicone rod insertion for tendon pseudo-sheath formation for at least 3 months and second-stage rod removal with tendon grafting. Functional results of this technique are not always predictable, which also depend a lot on patient’s compliance with prolonged duration of rehabilitation and time off work [106]. By employing tubular polyurethane nanocomposite graft surrounding autologous tenograft, single-staged flexor tendon reconstruction in sheep hind extremity model revealed mild histological adhesion and satisfactory tendon gliding [199]. In addition to artificial synovial graft, the in vivo experimental study of artificial tendon substitution assembled with platelet gel-collagen-polydioxanone has also shed light on promising Achilles tendon healing and adhesion prevention [26, 200]. It is believed that abundant PDGF level in such platelet gel may alter fibro-adiogenic progenitors from fibrotic differentiation towards tendon stem cells, thus relatively alleviate scar tissue formation [201].

Two-stage extensor tendon reconstruction is not a preferable approach in reported literature because of the absence of fibroosseous sheath guiding both extrinsic and intrinsic musculotendinous movement [202]. However, one-stage extensor tendon reconstruction may be accompanied with subsequent adhesion in cases with multiple soft tissue defect, severe contamination, and bone defect or fracture which requires long-term immobilization [203, 204]. In those cases, first-stage artificial tendon substitutions insertion, including silicone rods, can be an alternative method to maintain tendon route and form pseudo-synovial tunnel, to inhibit inflammation and infection, and thus to prevent adhesion formation for second-stage tendon grafting [205]. The use of such synthetic silicone rods has declined over years due to many complications, such as pyogenic tenosynovitis and high failure rate [206, 207].

Extra synovial tendon autografts are currently wide-accepted donor for tendon defect reconstruction. These tendons, such as palmaris longus, however, lack natural synovial cells mounted on loose connective tissue in tendons possessing sheaths, such as flexor digitorum superficialis. It has been observed in canine model that extra synovial tendon autograft underwent quicker cell death and ECM remodeling, and more excessive peritendinous adhesion, compared with intra synovial tendon autograft [208]. Explanation for such differences may lie in the different initial expression spectrum of these two kinds of tendons, and in their different response to the extracellular environment of the recipient site [209]. Rabbit model revealed that after 28 days of autograft transplantation using extra or intrasynovial tendon, donor segment showed different proteomic features in the expression of certain proteins, including heat shock protein 47, tenascin, periostin, etc., indicating the relationship among environmental stimuli around recipient site, oxidative stress, cell homeostasis and consequent peritendinous adhesion [23].

The availability of tendon allograft offers reconstruction options for patients without adequate tendon autograft reservation. Early studies have demonstrated triumphs in clinical flexor tendon reconstruction using fresh tendon allograft with composite sheath and volar plates, which, however, arouse concerns, including the harvesting and storing technique, transmission of diseases, and ethical issues [210]. Since programmed sterilization, decellularization and lyophilization have been shown to have little influence on biomechanical properties of composite tendon allograft, diverse modifications of such allograft have been studied to minimize postsurgery peritendinous adhesion [211, 212]. These modifications include synthetic polymer loading, tonogenic stem cell repopulation, and antiadhesion genes delivery, all of which were shown to reduce tendon graft gliding resistance in animal models [213, 214, 215]. Although safe and effective clinical use of lyophilized and sterilized tendon allograft has been reported in upper extremity reconstruction, no biologically modified acellular tendon allograft has been implanted in humans to date in the aim of preventing tenograft adhesion [216].

4.3.3 Pericardium as tendon sheath substitution

Due to the difficulties in tendon sheath suture with conventional surgical techniques, tendon sheath engineering has drawn wide attention in order to support tendon gliding and prevent peritendinous adhesion, apart from chemosynthetic materials. Typical biogenetic tendon sheath usually requires a scaffold, either from decellularized membraneous allograft or xenograft, or simple scaffold synthesized with collagen or lipoprotein, and techniques reseeding certain cells onto the membraneous scaffold. Bioengineered tendon sheath with collagen scaffold and harvested synoviocyte has been reported to inhibit Achilles tendon adhesion formation in rabbit model [217]. Porcine pericardium has also been applied as bioengineered tendon sheath scaffold since the pericardium functions similar as tendon sheath to lubricate heart beating, based on which decellularized bovine pericardium tendon xenograft and allograft have achieved successful adhesion prevention outcomes in chicken and donkey models [22]. Although Megerle et al. [25] did not test the tendon gliding resistance in vivo, their repopulationed human synoviocyte and adipose-derived stem cell on decellularized porcine pericardium showed shifted expression of downregulated collagen and upregulated hyaluronan synthase, suggesting potential antiadhesion mechanism during human cell line transplantation [218].

Advertisement

5. Conclusion

Taken together, the successful exploration and application of advanced antiadhesion materials highly depend on the collaboration among experts with diverse backgrounds, including engineering, biology, chemistry, surgeons, and sociologists. Intensive understanding of the pathophysiological procedure of tendon adhesion in combination with advanced material fabrication technologies will do a great favor to establishing next-generation of therapeutic platforms against tendon adhesion.

References

  1. 1. Apostolakos J et al. The enthesis: A review of the tendon-to-bone insertion. Muscle, Ligaments and Tendons Journal. 2014;4:333-342
  2. 2. Millar NL et al. Tendinopathy. Nature Reviews Disease Primers. 2021;7:1
  3. 3. Dy CJ et al. Complications after flexor tendon repair: A systematic review and meta-analysis. The Journal of Hand Surgery. 2012;37:543-551.e541
  4. 4. Yoneda S et al. The effect of modified locking methods and suture materials on zone II flexor tendon repair-an ex vivo study. PLoS One. 2018;13:e0205121
  5. 5. Titan AL, Foster DS, Chang J, Longaker MT. Flexor tendon: Development, healing, adhesion formation, and contributing growth factors. Plastic and Reconstructive Surgery. 2019;144:639e-647e
  6. 6. Gelberman RH, Steinberg D, Amiel D, Akeson W. Fibroblast chemotaxis after tendon repair. The Journal of Hand Surgery. 1991;16:686-693
  7. 7. Pearce O, Brown MT, Fraser K, Lancerotto L. Flexor tendon injuries: Repair & Rehabilitation. Injury. 2021;52:2053-2067
  8. 8. Thomopoulos S, Parks WC, Rifkin DB, Derwin KA. Mechanisms of tendon injury and repair. Journal of Orthopaedic Research. 2015;33:832-839
  9. 9. Taylor SH et al. Tendon is covered by a basement membrane epithelium that is required for cell retention and the prevention of adhesion formation. PLoS One. 2011;6:e16337
  10. 10. Meier Burgisser G et al. Rabbit Achilles tendon full transection model—Wound healing, adhesion formation and biomechanics at 3, 6 and 12 weeks post-surgery. Biology Open. 2016;5:1324-1333
  11. 11. Liu CF et al. What we should know before using tissue engineering techniques to repair injured tendons: A developmental biology perspective. Tissue Engineering Reviews (Part B). 2011;17:165-176
  12. 12. Zhou H, Lu H. Advances in the development of anti-adhesive biomaterials for tendon repair treatment. Tissue Engineering and Regenerative Medicine. 2021;18:1-14
  13. 13. Schnuriger B et al. Prevention of postoperative peritoneal adhesions: A review of the literature. American Journal of Surgery. 2011;201:111-121
  14. 14. Wu Y, Han Y, Wong YS, Fuh JYH. Fibre-based scaffolding techniques for tendon tissue engineering. Journal of Tissue Engineering and Regenerative Medicine. 2018;12:1798-1821
  15. 15. Rademakers T, Horvath JM, van Blitterswijk CA, LaPointe VLS. Oxygen and nutrient delivery in tissue engineering: Approaches to graft vascularization. Journal of Tissue Engineering and Regenerative Medicine. 2019;13:1815-1829
  16. 16. Cravedi P et al. Regenerative immunology: The immunological reaction to biomaterials. Transplant International. 2017;30:1199-1208
  17. 17. No YJ, Castilho M, Ramaswamy Y, Zreiqat H. Role of biomaterials and controlled architecture on tendon/ligament repair and regeneration. Advanced Materials. 2020;32:e1904511
  18. 18. Thomas SC, Jones LC, Hungerford DS. Hyaluronic acid and its effect on postoperative adhesions in the rabbit flexor tendon. A preliminary look. Clinical Orthopaedics and Related Research. 1986;206:281-289
  19. 19. Gomes SR et al. In vitro and in vivo evaluation of electrospun nanofibers of PCL, chitosan and gelatin: A comparative study. Materials Science & Engineering. C, Materials for Biological Applications. 2015;46:348-358
  20. 20. Zhang Q et al. Advanced technology-driven therapeutic interventions for prevention of tendon adhesion: Design, intrinsic and extrinsic factor considerations. Acta Biomaterialia. 2021;124:15-32
  21. 21. Brebels J, Mignon A. Polymer-based constructs for flexor tendon repair: A review. Polymers (Basel). 2022;14:867
  22. 22. El-Shafaey ESA, Karrouf GI, Zaghloul AE. Clinical and biomechanical evaluation of three bioscaffold augmentation devices used for superficial digital flexor tenorrhaphy in donkeys (Equus asinus): An experimental study. Journal of Advanced Research. 2013;4:103-113
  23. 23. Farnebo S, Wiig M, Holm B, Ghafouri B. Differentially expressed proteins in intra synovial compared to extra synovial flexor tendon grafts in a rabbit tendon transplantation model. Biomedicine. 2020;8:408
  24. 24. Sang R et al. Effect of acellular amnion with increased TGF-beta and bFGF levels on the biological behavior of tenocytes. Frontiers in Bioengineering and Biotechnology. 2020;8:446
  25. 25. Megerle K et al. Flexor tendon sheath engineering using decellularized porcine pericardium. Plastic and Reconstructive Surgery. 2016;138:630e-641e
  26. 26. Oryan A, Moshiri A, Meimandi-Parizi A. Implantation of a novel tissue-engineered graft in a large tendon defect initiated inflammation, accelerated fibroplasia and improved remodeling of the new Achilles tendon: A comprehensive detailed study with new insights. Cell and Tissue Research. 2014;355:59-80
  27. 27. Ding B, Wang X, Yao M. Photochemical tissue bonding technique for improving healing of hand tendon injury. Surgical Innovation. 2019;26:153-161
  28. 28. Kjaer M. Role of extracellular matrix in adaptation of tendon and skeletal muscle to mechanical loading. Physiological Reviews. 2004;84:649-698
  29. 29. Kjaer M et al. From mechanical loading to collagen synthesis, structural changes and function in human tendon. Scandinavian Journal of Medicine & Science in Sports. 2009;19:500-510
  30. 30. Birch HL. Tendon matrix composition and turnover in relation to functional requirements. International Journal of Experimental Pathology. 2007;88:241-248
  31. 31. Ackermann PW, Hart DA. Metabolic influences on risk for tendon disorders preface. Advances in Experimental Medicine and Biology. 2016;920:V-Vi
  32. 32. Thorpe CT, Screen HR. Tendon structure and composition. Advances in Experimental Medicine and Biology. 2016;920:3-10
  33. 33. Barnard K, Light ND, Sims TJ, Bailey AJ. Chemistry of the collagen cross-links. Origin and partial characterization of a putative mature cross-link of collagen. The Biochemical Journal. 1987;244:303-309
  34. 34. Canty EG, Kadler KE. Procollagen trafficking, processing and fibrillogenesis. Journal of Cell Science. 2005;118:1341-1353
  35. 35. Juncosa-Melvin N et al. Effects of mechanical stimulation on the biomechanics and histology of stem cell-collagen sponge constructs for rabbit patellar tendon repair. Tissue Engineering. 2006;12:2291-2300
  36. 36. Strickland JW. Development of flexor tendon surgery: Twenty-five years of progress. The Journal of Hand Surgery. 2000;25:214-235
  37. 37. Lin TW, Cardenas L, Soslowsky LJ. Biomechanics of tendon injury and repair. Journal of Biomechanics. 2004;37:865-877
  38. 38. Nichols AEC, Best KT, Loiselle AE. The cellular basis of fibrotic tendon healing: Challenges and opportunities. Translational Research. 2019;209:156-168
  39. 39. Millar NL, Murrell GA, McInnes IB. Inflammatory mechanisms in tendinopathy—Towards translation. Nature Reviews Rheumatology. 2017;13:110-122
  40. 40. Manning CN et al. Adipose-derived mesenchymal stromal cells modulate tendon fibroblast responses to macrophage-induced inflammation in vitro. Stem Cell Research & Therapy. 2015;6:74
  41. 41. Shen H et al. The effect of mesenchymal stromal cell sheets on the inflammatory stage of flexor tendon healing. Stem Cell Research & Therapy. 2016;7:144
  42. 42. Koob TJ, Summers AP. Tendon--bridging the gap. Comparative Biochemistry and Physiology. Part A, Molecular & Integrative Physiology. 2002;133:905-909
  43. 43. Strickland JW. Flexor tendons: Acute injuries. In: Green’s Operative Hand Surgery. London, UK: Churchill Livingstone; 1999
  44. 44. Das A et al. Monocyte and macrophage plasticity in tissue repair and regeneration. The American Journal of Pathology. 2015;185:2596-2606
  45. 45. Wojciak B, Crossan JF. The accumulation of inflammatory cells in synovial sheath and epitenon during adhesion formation in healing rat flexor tendons. Clinical and Experimental Immunology. 1993;93:108-114
  46. 46. Berglund ME et al. Neuropeptide, mast cell, and myofibroblast expression after rabbit deep flexor tendon repair. The Journal of Hand Surgery. 2010;35:1842-1849
  47. 47. Best KT, Loiselle AE. Scleraxis lineage cells contribute to organized bridging tissue during tendon healing and identify a subpopulation of resident tendon cells. The FASEB Journal. 2019;33:8578-8587
  48. 48. Bi Y et al. Identification of tendon stem/progenitor cells and the role of the extracellular matrix in their niche. Nature Medicine. 2007;13:1219-1227
  49. 49. Dyment NA et al. Lineage tracing of resident tendon progenitor cells during growth and natural healing. PLoS One. 2014;9:e96113
  50. 50. Yoshida R et al. Murine supraspinatus tendon injury model to identify the cellular origins of rotator cuff healing. Connective Tissue Research. 2016;57:507-515
  51. 51. Tempfer H et al. Perivascular cells of the supraspinatus tendon express both tendon- and stem cell-related markers. Histochemistry and Cell Biology. 2009;131:733-741
  52. 52. Mienaltowski MJ, Adams SM, Birk DE. Regional differences in stem cell/progenitor cell populations from the mouse achilles tendon. Tissue Engineering. Part A. 2013;19:199-210
  53. 53. Yin Z et al. Single-cell analysis reveals a nestin(+) tendon stem/progenitor cell population with strong tenogenic potentiality. Science Advances. 2016;2:e1600874
  54. 54. Sakabe T et al. Transcription factor scleraxis vitally contributes to progenitor lineage direction in wound healing of adult tendon in mice. The Journal of Biological Chemistry. 2018;293:5766-5780
  55. 55. Darby IA, Laverdet B, Bonte F, Desmouliere A. Fibroblasts and myofibroblasts in wound healing. Clinical, Cosmetic and Investigational Dermatology. 2014;7:301-311
  56. 56. Hu B, Phan SH. Myofibroblasts. Current Opinion in Rheumatology. 2013;25:71-77
  57. 57. Docheva D, Muller SA, Majewski M, Evans CH. Biologics for tendon repair. Advanced Drug Delivery Reviews. 2015;84:222-239
  58. 58. Branford OA, Klass BR, Grobbelaar AO, Rolfe KJ. The growth factors involved in flexor tendon repair and adhesion formation. The Journal of Hand Surgery (European Volume). 2014;39:60-70
  59. 59. Voleti PB, Buckley MR, Soslowsky LJ. Tendon healing: Repair and regeneration. Annual Review of Biomedical Engineering. 2012;14:47-71
  60. 60. Galvez MG, Crowe C, Farnebo S, Chang J. Tissue engineering in flexor tendon surgery: Current state and future advances. The Journal of Hand Surgery (European Volume). 2014;39:71-78
  61. 61. Wu YF et al. Adeno-associated virus-2-mediated TGF-beta1 microRNA transfection inhibits adhesion formation after digital flexor tendon injury. Gene Therapy. 2016;23:167-175
  62. 62. Lu H et al. Tanshinone IIA prevent tendon adhesion in the rat Achilles tendon model. Journal of Biomaterials and Tissue Engineering. 2016;6:739-744
  63. 63. Tang JB et al. Basic FGF or VEGF gene therapy corrects insufficiency in the intrinsic healing capacity of tendons. Scientific Reports. 2016;6:20643
  64. 64. Pan Q et al. Exosomes derived from mesenchymal stem cells ameliorate hypoxia/reoxygenation-injured ECs via transferring microRNA-126. Stem Cells International. 2019;2019:2831756
  65. 65. Evrova O, Buschmann J. In vitro and in vivo effects of Pdgf-bb delivery strategies on tendon healing: A review. European Cells & Materials. 2017;34:15-39
  66. 66. Yang QQ, Shao YX, Zhang LZ, Zhou YL. Therapeutic strategies for flexor tendon healing by nanoparticle-mediated co-delivery of bFGF and VEGFA genes. Colloids and Surfaces. B, Biointerfaces. 2018;164:165-176
  67. 67. Shen H et al. BMP12 induces tenogenic differentiation of adipose-derived stromal cells. PLoS One. 2013;8:e77613
  68. 68. Wang D et al. BMP14 induces tenogenic differentiation of bone marrow mesenchymal stem cells in vitro. Experimental and Therapeutic Medicine. 2018;16:1165-1174
  69. 69. Hyun SY, Lee JH, Kang KJ, Jang YJ. Effect of FGF-2, TGF-beta-1, and BMPs on Teno/Ligamentogenesis and osteo/Cementogenesis of human periodontal ligament stem cells. Molecules and Cells. 2017;40:550-557
  70. 70. de Aro AA et al. Injured Achilles tendons treated with adipose-derived stem cells transplantation and GDF-5. Cell. 2018;7:127
  71. 71. Niu X et al. Tendon cell regeneration is mediated by attachment site-resident progenitors and BMP Signaling. Current Biology. 2020;30:3277-3292 e3275
  72. 72. Chen S et al. Tenogenic adipose-derived stem cell sheets with nanoyarn scaffolds for tendon regeneration. Materials Science & Engineering. C, Materials for Biological Applications. 2021;119:111506
  73. 73. Sarikaya B, Gumusderelioglu M. Aligned silk fibroin/poly-3-hydroxybutyrate nanofibrous scaffolds seeded with adipose-derived stem cells for tendon tissue engineering. International Journal of Biological Macromolecules. 2021;193:276-286
  74. 74. Logerstedt DS et al. Effects of and response to mechanical loading on the knee. Sports Medicine. 2022;52:201-235
  75. 75. Rinoldi C et al. Tendon tissue engineering: Effects of mechanical and biochemical stimulation on stem cell alignment on cell-laden hydrogel yarns. Advanced Healthcare Materials. 2019;8:e1801218
  76. 76. Abdulmalik S et al. The glucagon-like peptide 1 receptor agonist Exendin-4 induces tenogenesis in human mesenchymal stem cells. Differentiation. 2021;120:1-9
  77. 77. Olesen JL et al. No treatment benefits of local administration of insulin-like growth factor-1 in addition to heavy slow resistance training in tendinopathic human patellar tendons a randomized, double-blind, placebo-controlled trial with 1-year follow-up. The American Journal of Sports Medicine. 2021;49:2361-2370
  78. 78. Farnebo S et al. Optimized repopulation of tendon hydrogel: Synergistic effects of growth factor combinations and adipose-derived stem cells. Hand (New York, N.Y.). 2017;12:68-77
  79. 79. Holladay C et al. Preferential tendon stem cell response to growth factor supplementation. Journal of Tissue Engineering and Regenerative Medicine. 2016;10:783-798
  80. 80. Chan KM et al. Expression of transforming growth factor beta isoforms and their roles in tendon healing. Wound Repair and Regeneration. 2008;16:399-407
  81. 81. Hoesel B, Schmid JA. The complexity of NF-kappa B signaling in inflammation and cancer. Molecular Cancer. 2013;12:86
  82. 82. Mitchell JP, Carmody RJ. NF-kappa B and the transcriptional control of inflammation. International Review of Cell and Molecular Biology. 2018;335:41-84
  83. 83. Abraham AC et al. Targeting the NF-kappa B signaling pathway in chronic tendon disease. Science Translational Medicine. 2019;11:eaav4319
  84. 84. Loiselle AE et al. Development of antisense oligonucleotide (ASO) technology against Tgf-beta Signaling to prevent scarring during flexor tendon repair. Journal of Orthopaedic Research. 2015;33:859-866
  85. 85. Zhang Y, Alexander PB, Wang XF. TGF-beta family Signaling in the control of cell proliferation and survival. Cold Spring Harbor Perspectives in Biology. 2017;9:a022145
  86. 86. Jiang S et al. Down-regulating ERK1/2 and SMAD2/3 phosphorylation by physical barrier of celecoxib-loaded electrospun fibrous membranes prevents tendon adhesions. Biomaterials. 2014;35:9920-9929
  87. 87. Yao Z et al. MicroRNA-21-3p engineered umbilical cord stem cell-derived exosomes inhibit tendon adhesion. Journal of Inflammation Research. 2020;13:303-316
  88. 88. Yao Z et al. MicroRNA engineered umbilical cord stem cell-derived exosomes direct tendon regeneration by mTOR signaling. Journal of Nanobiotechnology. 2021;19:169
  89. 89. Hata A, Chen YG. TGF-beta Signaling from receptors to Smads. Cold Spring Harbor Perspectives in Biology. 2016;8:a022061
  90. 90. Wu G et al. Three-dimensional tendon scaffold loaded with TGF-beta1 gene silencing plasmid prevents tendon adhesion and promotes tendon repair. ACS Biomaterials Science & Engineering. 2021;7:5739-5748
  91. 91. Weng CH et al. Interleukin-17A induces renal fibrosis through the ERK and Smad signaling pathways. Biomedicine & Pharmacotherapy. 2020;123:109741
  92. 92. Wu B et al. The TGF-beta superfamily cytokine activin-a is induced during autoimmune neuroinflammation and drives pathogenic Th17 cell differentiation. Immunity. 2021;54:308-323 e306
  93. 93. Katzel EB et al. Impact of Smad3 loss of function on scarring and adhesion formation during tendon healing. Journal of Orthopaedic Research. 2011;29:684-693
  94. 94. Xue M et al. Endogenous MMP-9 and not MMP-2 promotes rheumatoid synovial fibroblast survival, inflammation and cartilage degradation. Rheumatology (Oxford, England). 2014;53:2270-2279
  95. 95. Margulis A et al. MMP dependence of fibroblast contraction and collagen production induced by human mast cell activation in a three-dimensional collagen lattice. American Journal of Physiology. Lung Cellular and Molecular Physiology. 2009;296:L236-L247
  96. 96. Loiselle AE et al. Remodeling of murine Intrasynovial tendon adhesions following injury: MMP and Neotendon gene expression. Journal of Orthopaedic Research. 2009;27:833-840
  97. 97. Lu CC et al. Lateral slit delivery of bone marrow stromal cells enhances regeneration in the decellularized allograft flexor tendon. The Journal of Orthopaedic Translation. 2019;19:58-67
  98. 98. Cai C et al. Self-healing hydrogel embodied with macrophage-regulation and responsive-gene-silencing properties for synergistic prevention of peritendinous adhesion. Advanced Materials. 2022;34:e2106564
  99. 99. Blomgran P, Blomgran R, Ernerudh J, Aspenberg P. Cox-2 inhibition and the composition of inflammatory cell populations during early and mid-time tendon healing. Muscle, Ligaments and Tendons Journal. 2017;7:223-229
  100. 100. Cheng H et al. Role of prostaglandin E2 in tissue repair and regeneration. Theranostics. 2021;11:8836-8854
  101. 101. Dolkart O et al. Statins enhance rotator cuff healing by stimulating the COX2/PGE2/EP4 pathway an In vivo and In vitro study. The American Journal of Sports Medicine. 2014;42:2869-2876
  102. 102. Wang Y et al. High concentration of aspirin induces apoptosis in rat tendon stem cells via inhibition of the Wnt/beta-catenin pathway. Cellular Physiology and Biochemistry. 2018;50:2046-2059
  103. 103. Bindu S, Mazumder S, Bandyopadhyay U. Non-steroidal anti-inflammatory drugs (NSAIDs) and organ damage: A current perspective. Biochemical Pharmacology. 2020;180:114147
  104. 104. Guo C et al. Pharmacological properties and derivatives of shikonin-a review in recent years. Pharmacological Research. 2019;149:104463
  105. 105. Geary MB et al. Systemic EP4 inhibition increases adhesion formation in a murine model of flexor tendon repair. PLoS One. 2015;10:e0136351
  106. 106. Samora JB, Klinefelter RD. Flexor tendon reconstruction. The Journal of the American Academy of Orthopaedic Surgeons. 2016;24:28-36
  107. 107. Malizos KN et al. Infections of deep hand and wrist compartments. Microorganisms. 2020;8:838
  108. 108. Pickrell BB, Eberlin KR. Secondary surgery following replantation and revascularization. Hand Clinics. 2019;35:231-240
  109. 109. Shalumon KT et al. Multi-functional electrospun antibacterial core-shell nanofibrous membranes for prolonged prevention of post-surgical tendon adhesion and inflammation. Acta Biomaterialia. 2018;72:121-136
  110. 110. Altan E, Yalcin L. Release of adhesions after zone II flexor tendon repair: The ‘Gigli suture’ technique. The Journal of Hand Surgery (European Volume). 2017;42:525-526
  111. 111. Zhang CH et al. Evaluation of decellularized bovine tendon sheets for Achilles tendon defect reconstruction in a rabbit model. The American Journal of Sports Medicine. 2018;46:2687-2699
  112. 112. Van der Merwe W et al. Xenograft for anterior cruciate ligament reconstruction was associated with high graft processing infection. The Journal of Experimental Orthopaedics. 2020;7:79
  113. 113. Bai T, Zhang Q, Leng Y. A comparative study of biological and nonbiological sheath substitute in prevention of flexor tendon adhesions. Journal of Nanjing Railway Medical College. 1999;18:229-232
  114. 114. Kulick MI, Brazlow R, Smith S, Hentz VR. Injectable ibuprofen: Preliminary evaluation of its ability to decrease peritendinous adhesions. Annals of Plastic Surgery. 1984;13:459-467
  115. 115. Moore N et al. Pharmacoepidemiology of non-steroidal anti-inflammatory drugs. Thérapie. 2019;74:271-277
  116. 116. McGrath MH. Local steroid therapy in the hand. The Journal of Hand Surgery. 1984;9:915-921
  117. 117. Tatari H et al. Deleterious effects of local corticosteroid injections on the Achilles tendon of rats. Archives of Orthopaedic and Trauma Surgery. 2001;121:333-337
  118. 118. Yu HL et al. RGD-peptides modifying dexamethasone: To enhance the anti-inflammatory efficacy and limit the risk of osteoporosis. MedChemComm. 2015;6:1345-1351
  119. 119. Wu W et al. Advances in biomaterials for preventing tissue adhesion. Journal of Controlled Release. 2017;261:318-336
  120. 120. Alimohammadi M et al. Electrospun nanofibrous membranes for preventing tendon adhesion. ACS Biomaterials Science & Engineering. 2020;6:4356-4376
  121. 121. Khan F, Tanaka M. Designing smart biomaterials for tissue engineering. International Journal of Molecular Sciences. 2017;19:17
  122. 122. Jenkins MJ, Harrison KL. The effect of crystalline morphology on the degradation of polycaprolactone in a solution of phosphate buffer and lipase. Polymers for Advanced Technologies. 2008;19:1901-1906
  123. 123. Guo M et al. The effects of tensile stress on degradation of biodegradable PLGA membranes: A quantitative study. Polymer Degradation and Stability. 2016;124:95-100
  124. 124. Migliaresi C, Fambri L, Cohn D. A study on the in vitro degradation of poly(lactic acid). Journal of Biomaterials Science. Polymer Edition. 1994;5:591-606
  125. 125. Goh JC, Sahoo S. Scaffolds for tendon and ligament tissue engineering. In: Regenerative Medicine and Biomaterials for the Repair of Connective Tissues. Sawston Cambridge, UK: Woodhead Publishing; 2010
  126. 126. Cannella V et al. In vitro biocompatibility evaluation of nine dermal fillers on L929 cell line. BioMed Research International. 2020;2020:8676343
  127. 127. Chen SH, Chen CH, Fong YT, Chen JP. Prevention of peritendinous adhesions with electrospun chitosan-grafted polycaprolactone nanofibrous membranes. Acta Biomaterialia. 2014;10:4971-4982
  128. 128. Liu S et al. Prevention of peritendinous adhesions with electrospun ibuprofen-loaded poly(L-lactic acid)-polyethylene glycol fibrous membranes. Tissue Engineering. Part A. 2013;19:529-537
  129. 129. Uyanik O et al. Prevention of peritendinous adhesions with electrospun poly (lactic acid-co-glycolic acid) (PLGA) bioabsorbable nanofiber: An experimental study. Colloids and Surfaces. B, Biointerfaces. 2022;209:112181
  130. 130. Song A, Rane AA, Christman KL. Antibacterial and cell-adhesive polypeptide and poly(ethylene glycol) hydrogel as a potential scaffold for wound healing. Acta Biomaterialia. 2012;8:41-50
  131. 131. Karaaltin MV et al. The effects of 5-fluorouracil on flexor tendon healing by using a biodegradable gelatin, slow releasing system: Experimental study in a hen model. The Journal of Hand Surgery (European Volume). 2013;38:651-657
  132. 132. Moreland LW. Intra-articular hyaluronan (hyaluronic acid) and hylans for the treatment of osteoarthritis: Mechanisms of action. Arthritis Research & Therapy. 2003;5:54-67
  133. 133. Yilmaz E et al. The effect of seprafilm on adhesion formation and tendon healing after flexor tendon repair in chicken. Orthopedics. 2010;33:164-170
  134. 134. Kuo SM et al. Evaluation of the ability of xanthan gum/gellan gum/hyaluronan hydrogel membranes to prevent the adhesion of postrepaired tendons. Carbohydrate Polymers. 2014;114:230-237
  135. 135. Burgisser GM, Buschmann J. History and performance of implant materials applied as peritendinous antiadhesives. Journal of Biomedical Materials Research Part B. 2015;103:212-228
  136. 136. Ishiyama N et al. Reduction of peritendinous adhesions by hydrogel containing biocompatible phospholipid polymer MPC for tendon repair. The Journal of Bone and Joint Surgery. American Volume. 2011;93:142-149
  137. 137. Liu R, Zhang S, Chen X. Injectable hydrogels for tendon and ligament tissue engineering. Journal of Tissue Engineering and Regenerative Medicine. 2020;14:1333-1348
  138. 138. Chou PY et al. Thermo-responsive in-situ forming hydrogels as barriers to prevent post-operative peritendinous adhesion. Acta Biomaterialia. 2017;63:85-95
  139. 139. Zhou YL et al. Localized delivery of miRNAs targets cyclooxygenases and reduces flexor tendon adhesions. Acta Biomaterialia. 2018;70:237-248
  140. 140. Rinoldi C et al. Mechanical and biochemical stimulation of 3D multilayered scaffolds for tendon tissue engineering. ACS Biomaterials Science & Engineering. 2019;5:2953-2964
  141. 141. Madkhali O, Mekhail G, Wettig SD. Modified gelatin nanoparticles for gene delivery. International Journal of Pharmaceutics. 2019;554:224-234
  142. 142. Lee EJ, Kasper FK, Mikos AG. Biomaterials for tissue engineering. Annals of Biomedical Engineering. 2014;42:323-337
  143. 143. Lomas AJ et al. The past, present and future in scaffold-based tendon treatments. Advanced Drug Delivery Reviews. 2015;84:257-277
  144. 144. James R, Kesturu G, Balian G, Chhabra AB. Tendon: Biology, biomechanics, repair, growth factors, and evolving treatment options. The Journal of Hand Surgery. 2008;33:102-112
  145. 145. Zhou YL, Yang QQ, Zhang L, Tang JB. Nanoparticle-coated sutures providing sustained growth factor delivery to improve the healing strength of injured tendons. Acta Biomaterialia. 2021;124:301-314
  146. 146. Liu S et al. Biomimetic sheath membrane via electrospinning for antiadhesion of repaired tendon. Biomacromolecules. 2012;13:3611-3619
  147. 147. Yao Y et al. Effect of sustained heparin release from PCL/chitosan hybrid small-diameter vascular grafts on anti-thrombogenic property and endothelialization. Acta Biomaterialia. 2014;10:2739-2749
  148. 148. Zhang Q et al. Electrospun polymeric micro/nanofibrous scaffolds for long-term drug release and their biomedical applications. Drug Discovery Today. 2017;22:1351-1366
  149. 149. Luong-Van E et al. Controlled release of heparin from poly(epsilon-caprolactone) electrospun fibers. Biomaterials. 2006;27:2042-2050
  150. 150. Szentivanyi A, Chakradeo T, Zernetsch H, Glasmacher B. Electrospun cellular microenvironments: Understanding controlled release and scaffold structure. Advanced Drug Delivery Reviews. 2011;63:209-220
  151. 151. Casper CL, Yamaguchi N, Kiick KL, Rabolt JF. Functionalizing electrospun fibers with biologically relevant macromolecules. Biomacromolecules. 2005;6:1998-2007
  152. 152. Zeng J et al. Poly(vinyl alcohol) nanofibers by electrospinning as a protein delivery system and the retardation of enzyme release by additional polymer coatings. Biomacromolecules. 2005;6:1484-1488
  153. 153. Zomer Volpato F et al. Preservation of FGF-2 bioactivity using heparin-based nanoparticles, and their delivery from electrospun chitosan fibers. Acta Biomaterialia. 2012;8:1551-1559
  154. 154. Cheng L et al. Surface biofunctional drug-loaded electrospun fibrous scaffolds for comprehensive repairing hypertrophic scars. Biomaterials. 2016;83:169-181
  155. 155. Sultanova Z, Kaleli G, Kabay G, Mutlu M. Controlled release of a hydrophilic drug from coaxially electrospun polycaprolactone nanofibers. International Journal of Pharmaceutics. 2016;505:133-138
  156. 156. Jiang YN, Mo HY, Yu DG. Electrospun drug-loaded core-sheath PVP/zein nanofibers for biphasic drug release. International Journal of Pharmaceutics. 2012;438:232-239
  157. 157. Hu J et al. Drug-loaded emulsion electrospun nanofibers: Characterization, drug release and in vitro biocompatibility. RSC Advances. 2015;5:100256-100267
  158. 158. Sy JC, Klemm AS, Shastri VP. Emulsion as a means of controlling electrospinning of polymers. Advanced Materials. 2009;21:376-381
  159. 159. Zhao X et al. Optimization of intrinsic and extrinsic tendon healing through controllable water-soluble mitomycin-C release from electrospun fibers by mediating adhesion-related gene expression. Biomaterials. 2015;61:61-74
  160. 160. Xu X et al. Ultrafine medicated fibers electrospun from W/O emulsions. Journal of Controlled Release. 2005;108:33-42
  161. 161. Manning CN et al. Controlled delivery of mesenchymal stem cells and growth factors using a nanofiber scaffold for tendon repair. Acta Biomaterialia. 2013;9:6905-6914
  162. 162. Chainani A et al. Multilayered electrospun scaffolds for tendon tissue engineering. Tissue Engineering. Part A. 2013;19:2594-2604
  163. 163. Li LF et al. Release of celecoxib from a bi-layer biomimetic tendon sheath to prevent tissue adhesion. Materials Science and Engineering C: Materials for Biological Applications. 2016;61:220-226
  164. 164. Okuda T, Tominaga K, Kidoaki S. Time-programmed dual release formulation by multilayered drug-loaded nanofiber meshes. Journal of Controlled Release. 2010;143:258-264
  165. 165. Hu C, Cui W. Hierarchical structure of electrospun composite fibers for long-term controlled drug release carriers. Advanced Healthcare Materials. 2012;1:809-814
  166. 166. Hu C et al. Long-term drug release from electrospun fibers for in vivo inflammation prevention in the prevention of peritendinous adhesions. Acta Biomaterialia. 2013;9:7381-7388
  167. 167. Slowing II, Vivero-Escoto JL, Wu CW, Lin VS. Mesoporous silica nanoparticles as controlled release drug delivery and gene transfection carriers. Advanced Drug Delivery Reviews. 2008;60:1278-1288
  168. 168. Kim CH et al. An improved hydrophilicity via electrospinning for enhanced cell attachment and proliferation. Journal of Biomedical Materials Research. Part B, Applied Biomaterials. 2006;78:283-290
  169. 169. De Bartolo L, Morelli S, Bader A, Drioli E. The influence of polymeric membrane surface free energy on cell metabolic functions. Journal of Materials Science. Materials in Medicine. 2001;12:959-963
  170. 170. Fakhraei O et al. Nanofibrous polycaprolactone/chitosan membranes for preventing postsurgical tendon adhesion. Journal of Biomedical Materials Research. Part B, Applied Biomaterials. 2022;110:1279-1291
  171. 171. Van der Schueren L, Steyaert I, De Schoenmaker B, De Clerck K. Polycaprolactone/chitosan blend nanofibres electrospun from an acetic acid/formic acid solvent system. Carbohydrate Polymers. 2012;88:1221-1226
  172. 172. Wang YW et al. Effect of mitomycin on normal dermal fibroblast and HaCat cell: An in vitro study. Journal of Zhejiang University. Science. B. 2012;13:997-1005
  173. 173. Chen CH, Chen SH, Shalumon KT, Chen JP. Dual functional core-sheath electrospun hyaluronic acid/polycaprolactone nanofibrous membranes embedded with silver nanoparticles for prevention of peritendinous adhesion. Acta Biomaterialia. 2015;26:225-235
  174. 174. Chen CH et al. Functional hyaluronic acid-polylactic acid/silver nanoparticles Core-sheath nanofiber membranes for prevention of post-operative tendon adhesion. International Journal of Molecular Sciences. 2021;22:8781
  175. 175. El Khatib M et al. Fabrication and plasma surface activation of aligned electrospun PLGA Fiber fleeces with improved adhesion and infiltration of amniotic epithelial stem cells maintaining their Teno-inductive potential. Molecules. 2020;25:3176
  176. 176. Liu S et al. Tendon healing and anti-adhesion properties of electrospun fibrous membranes containing bFGF loaded nanoparticles. Biomaterials. 2013;34:4690-4701
  177. 177. Liao JCY et al. The effects of bi-functional anti-adhesion scaffolds on flexor tendon healing in a rabbit model. Journal of Biomedical Materials Research. Part B, Applied Biomaterials. 2018;106:2605-2614
  178. 178. Liu S et al. Gene silencing via PDA/ERK2-siRNA-mediated electrospun Fibers for peritendinous antiadhesion. Advanced Science (Weinheim, Baden-Württemberg, Germany). 2019;6:1801217
  179. 179. Li F et al. Efficient inhibition of the formation of joint adhesions by ERK2 small interfering RNAs. Biochemical and Biophysical Research Communications. 2010;391:795-799
  180. 180. Tang JB et al. Gene therapy strategies to improve strength and quality of flexor tendon healing. Expert Opinion on Biological Therapy. 2016;16:291-301
  181. 181. Liu S et al. Macrophage infiltration of electrospun polyester fibers. Biomaterials Science. 2017;5:1579-1587
  182. 182. Chen E et al. An asymmetric chitosan scaffold for tendon tissue engineering. Acta Biomaterialia. 2018;73:377-387
  183. 183. Jayasree A et al. Bioengineered braided micro-nano (multiscale) fibrous scaffolds for tendon reconstruction. ACS Biomaterials Science & Engineering. 2019;11:1476-1486
  184. 184. Park DY et al. Cross-linked cartilage acellular matrix film decreases postsurgical peritendinous adhesions. Artificial Organs. 2020;44:E136-E149
  185. 185. Song Y et al. Surface modification of electrospun fibers with mechano-growth factor for mitigating the foreign-body reaction. Bioactive Materials. 2021;6:2983-2998
  186. 186. Liu YS et al. No midterm advantages in the middle term using small intestinal submucosa and human amniotic membrane in Achilles tendon transverse tenotomy. Journal of Orthopaedic Surgery and Research. 2016;11:125
  187. 187. Koob TJ et al. Biological properties of dehydrated human amnion/chorion composite graft: Implications for chronic wound healing. International Wound Journal. 2013;10:493-500
  188. 188. Dogramaci Y, Duman IG. Reinforcement of the flexor tendon repair using human amniotic MembraneA biomechanical evaluation using the modified Kessler method of tendon repair. Journal of the American Podiatric Medical Association. 2016;106:319-322
  189. 189. Liu C et al. Experimental study of tendon sheath repair via decellularized amnion to prevent tendon adhesion. PLoS One. 2018;13:e0205811
  190. 190. Prakash S, Kalra P, Dhal A. Flexor tendon repair with amniotic membrane. International Orthopaedics. 2020;44:2037-2045
  191. 191. Bennett NT, Schultz GS. Growth-factors and wound-healing—Biochemical-properties of growth-factors and their receptors. American Journal of Surgery. 1993;165:728-737
  192. 192. Shu J et al. Human amnion mesenchymal cells inhibit lipopolysaccharide-induced TNF-alpha and IL-1beta production in THP-1 cells. Biological Research. 2015;48:69
  193. 193. Duan-Arnold Y et al. Retention of endogenous viable cells enhances the anti-inflammatory activity of cryopreserved amnion. Advances in Wound Care (New Rochelle). 2015;4:523-533
  194. 194. Yao M et al. Phototoxicity is not associated with photochemical tissue bonding of skin. Lasers in Surgery and Medicine. 2010;42:123-131
  195. 195. Liu CJ et al. Regulation of ERK1/2 and SMAD2/3 pathways by using multi-layered electrospun PCL-amnion nanofibrous membranes for the prevention of post-surgical tendon adhesion. International Journal of Nanomedicine. 2020;15:927-942
  196. 196. Liu C et al. Biological amnion prevents flexor tendon adhesion in zone II: A controlled, multicentre clinical trial. BioMed Research International. 2019;2019:2354325
  197. 197. Leppanen OV et al. Outcomes after flexor tendon repair combined with the application of human amniotic membrane allograft. The Journal of Hand Surgery. 2017;42:474 e471-474 e478
  198. 198. Kibadi K, Moutet F. Silicone infusion tubing instead of hunter rods for two-stage zone 2 flexor tendon reconstruction in a resource-limited surgical environment. Hand Surgery & Rehabilitation. 2017;36:384-387
  199. 199. Sedghizadeh P et al. Design and in vivo testing of novel bisphosphonate-fluoroquinolone conjugates chemisorbed to bone graft material. Journal of Bone and Mineral Research. 2019;34:367-368
  200. 200. Moshiri A, Oryan A, Meimandi-Parizi A. Synthesis, development, characterization and effectiveness of bovine pure platelet gel-collagen-polydioxanone bioactive graft on tendon healing. Journal of Cellular and Molecular Medicine. 2015;19:1308-1332
  201. 201. Harvey T, Flamenco S, Fan CM. A Tppp3(+)Pdgfra(+) tendon stem cell population contributes to regeneration and reveals a shared role for PDGF signalling in regeneration and fibrosis. Nature Cell Biology. 2019;21:1490-1503
  202. 202. Al-Qattan MM. Two-staged extensor tendon reconstruction for zone 6 extensor tendon loss of the fingers: Indications, technique and results. The Journal of Hand Surgery (European Volume). 2015;40:276-280
  203. 203. Sundine M, Scheker LR. A comparison of immediate and staged reconstruction of the dorsum of the hand. Journal of Hand Surgery (British). 1996;21:216-221
  204. 204. Adani R, Marcoccio I, Tarallo L. Flap coverage of dorsum of hand associated with extensor tendons injuries: A completely vascularized single-stage reconstruction. Microsurgery. 2003;23:32-39
  205. 205. Abdulaziz MKB et al. Two-stage reconstruction of hand extensor tendons using silicon rods. Plastic and Reconstructive Surgery—Global Open. 2021;9:e3858
  206. 206. Bashaireh KM, Audat Z, Radaideh AM, Aleshawi AJ. The effectiveness of autograft used in anterior cruciate ligament reconstruction of the knee: Surgical Records for the new Generations of Orthopedic surgeons and synthetic graft revisit. Orthopedic Research and Reviews. 2020;12:61-67
  207. 207. Satora W, Krolikowska A, Czamara A, Reichert P. Synthetic grafts in the treatment of ruptured anterior cruciate ligament of the knee joint. Polimery w Medycynie. 2017;47:55-59
  208. 208. Seiler JG 3rd et al. The Marshall R. Urist Young Investigator Award. Autogenous flexor tendon grafts. Biologic mechanisms for incorporation. Clinical Orthopaedics and Related Research. 1997;345:239-247
  209. 209. Berglund M et al. Assessment of mRNA levels for matrix molecules and TGF- beta1 in rabbit flexor and peroneus tendons reveals regional differences in steady-state expression. Journal of Hand Surgery (British). 2004;29:165-169
  210. 210. Peacock EE Jr, Madden JW. Human composite flexor tendon allografts. Annals of Surgery. 1967;166:624-629
  211. 211. Drake DB, Tilt AC, DeGeorge BR. Acellular flexor tendon allografts: A new horizon for tendon reconstruction. The Journal of Hand Surgery. 2013;38:2491-2495
  212. 212. DeGeorge BR Jr, Rodeheaver GT, Drake DB. The biophysical characteristics of human composite flexor tendon allograft for upper extremity reconstruction. Annals of Plastic Surgery. 2014;72:S184-S190
  213. 213. Kryger GS et al. A comparison of tenocytes and mesenchymal stem cells for use in flexor tendon tissue engineering. The Journal of Hand Surgery. 2007;32a:597-605
  214. 214. Basile P et al. Freeze-dried tendon Allografts as tissue-engineering scaffolds for gdf5 gene delivery. Molecular Therapy. 2008;16:466-473
  215. 215. Zhao C et al. Improvement of flexor tendon reconstruction with carbodiimide-derivatized hyaluronic acid and gelatin-modified intrasynovial allografts: Study of a primary repair failure model. The Journal of Bone and Joint Surgery. American Volume. 2010;92:2817-2828
  216. 216. Tang L et al. Long-term effectiveness of tendon allograft for repairing tendon defect. Zhongguo Xiu Fu Chong Jian Wai Ke Za Zhi. 2011;25:341-343
  217. 217. Baymurat AC et al. Bio-engineered synovial membrane to prevent tendon adhesions in rabbit flexor tendon model. Journal of Biomedical Materials Research. Part A. 2015;103:84-90
  218. 218. Sungur N et al. Prevention of tendon adhesions by the reconstruction of the tendon sheath with solvent dehydrated bovine pericard: An experimental study. The Journal of Trauma. 2006;61:1467-1472

Written By

Shen Liu, Qinglin Kang, Rui Zhang, Yanhao Li and Rong Bao

Submitted: 22 August 2022 Reviewed: 12 September 2022 Published: 03 November 2022