Abstract
Cardiovascular tissue constructs provide unique design requirements due to their functional responses to the substrate mechanical properties and cyclic stretching behavior of the tissue that requires the use of durable yet elastic materials. Given the diversity in polyester synthesis approaches, an opportunity exists to develop a new class of biocompatible, elastic, and immunomodulatory cardiovascular polymers. Furthermore, the elastomeric polyester materials have the capability to provide tailored biomechanical synergy with native tissue and hence reduce inflammatory response in vivo and better support tissue maturation in vitro. In this review, we highlight underlying chemistry and design strategies of polyester elastomers optimized for cardiac tissue scaffolds. Major advantages of these materials such as their tunable elasticity, desirable biodegradation, and potential for incorporation of bioactive compounds are further expanded. Their unique fabrication methods such as micromolding, 3D stamping, electrospinning, laser ablation and 3D printing are discussed. Moreover, application of these biomaterials in cardiovascular organ-on-a-chip devices and patches are analyzed. Finally, we outline unaddressed challenges in the field that need further study to enable impactful translation of soft polyesters to clinical applications.
Keywords: Tissue Engineering, Polyester, Biomaterials, Organ-on-a-chip, Microfabrication
1. Introduction
Biomaterial based technologies have played an indispensable role in the development of cardiovascular devices, tissue regeneration, acellular therapies, drug delivery, and in vitro scaffolds for tissue engineering. 1 The evolution of biomaterials has been a journey of continuous learning and innovation, marked by significant milestones that have shaped the field. In the early days, biomedical devices faced poor patient integration due to the lack of sterilization techniques and unadvanced knowledge of human biology. 2 Observations of healing of foreign bodies upon injuries during World War II inspired surgeons such as Sir Harold Ridley3, Sir John Charnley4, Dr. Charles Stent5 and many other surgeons to implant biomaterials into human body in many forms such as contact lenses, hip replacements, vascular stents and grafts. At the time, materials were deemed to be “biocompatible” if they did not provoke significant immune rejection within the host body. 6 Some of the commonly utilized materials at this era were silk, silicones, metals, poly(methyl methacrylates) and Teflon. 7
In 1960’s and 1970’s, biodegradability, the ability of the material to be broken down and removed from implant site became the major criteria for biocompatibility. 7 Polymers such as poly (lactic acid) (PLA), poly(glycolic) acid (PGA) or poly (lactic-co-glycolic) acid (PLGA) were developed. 8 Recent advances in understanding of concepts such as cell-surface receptors, cellular phenotypes, and cell-biomaterial crosstalk, the field has evolved to develop biomaterials that enhance the polymer-host tissue crosstalk. Novel techniques involving host immune response modulation, increased porosity to enhance vascularization, controlled drug release, biomaterials modified with bioactive molecules, nonfouling materials, thermoresponsive compounds and materials enhancing healing process have been realized. 9
Cardiovascular tissue engineering and biomaterials has recently focused on new approaches for regeneration of cardiac muscle, the myocardium, given promise to provide solutions for the failing heart. Given the low elasticity range of human myocardium during the diastole and systole cycles (10-300 kPa), an ideal biomaterial for cardiac tissue engineering should possess a low Young’s modulus to match the physiological range, as well as high elongation and tensile strength to support cyclic contractile behavior of cardiac tissue.40-43 Polyesters have drawn great attention toward this goal due to their notable potential to mimic physiological and biochemical properties of native tissues, and match the elasticity of cardiac tissue. 10
Polyesters are typically synthesized through esterification, which involves a reaction between an alcohol (or diol) and a carboxylic acid or (or diacid) group forming an ester linkages. 11 The ester linkage and control over crosslinking extent offers flexibility and mobility to the polymeric structure. Biodegradable polyesters are structurally and chemically stable for short- and medium-term applications and can provid mechanical cues to drive cellular behaviour12. They can also degrade over time via several mechanisms including hydrolysis, oxidation or enzymaticly. 13 The mechanical properties, degradation rate and biocompatibility of polyesters can be tuned by the choice of monomers and their molecular weights as well as the processing conditions (e.g., temperature, pressure, and reaction time). 14 Elastomeric polyesters have been frequently used in cardiovascular applications as artificial heart valves15, vascular prosthesis16, conduits17-18, and patches for repair of damaged myocardium or congenital malformations18. Biofunctionalization of these scaffold materials with natural materials such as laminin, fibronectin, collagen, and alginate can further promote cell attachment by modifying surface elasticity or hydrophobicity. 19-20
Mechanical properties, cytocompatibility, immunomodulation and fabrication of precise micron-scale structures are some of the major challenges that should be addressed during biomaterial design process. Building on the previous successes with polyester elastomers and given the diversity in polyester synthesis approaches, it is possible to develop a new class of biocompatible, elastic, and immunomodulatory cardiovascular polymers. This review compares the biological and mechanical properties of currently developed polyesters and evaluates their advantages and shortcomings in cardiac tissue engineering and organ-on-a-chip applications. We describe the challenges associated with fabricating scaffolds from polyesters, as well as innovative approaches that have been developed to overcome some of these limitations. This review also provides a comprehensive explanation of various in vitro and in vivo applications of polyesters in the field of cardiovascular research and foresees the clinical transition of such applications. Commonly used polyesters for soft tissue engineering developed in the last 20 years are highlighted in Figure 1.
Figure 1. Timeline of recent developments of soft polyester elastomers suitable for cardiovascular applications.

a. F-actin filament of cells attached on a poly(glycerol sebacate(PGS)-based patches seeded with cardiomyocytes. 21 Scale bar, 200 mm. Copyright 2008, Nature publication Group. 21 b. H&E staining of Poly(1,8-octanediol citrate) (POC) samples implanted in vivo. Copyright 2004, John Wiley & Sons. 22 Scale bar, 100 mm. c. Immunostaining of α-actinin and F-actin on cardiac patch made from poly(octamethylene maleate (anhydride) citrate (POMaC) and seeded with cardiomyocytes. Copyright 2015, Science publications. 23 Scale bar, 30 mm. d. 1,2,4 polymer developed in Radisic lab. 12 Engineered poly(octamethylene maleate (anhydride) 1,2,4-butanetricarboxylate) 1,2,4 polymer-based micro-scaled tube, AngioTube, seeded with endothelial cells and stained for CD31 to illustrate lumen sprouting. 12 Copyright 2021, Nature Publication Group. 24 Scale bar, 100 mm. Inset scale bar, 20 mm. e. Poly(itaconate-co-citrate-co-octanediol) (PICO)-based cardiac patch seeded with cardiomyocytes and stained for cardiac marker α-actinin and general filament marker, F-actin. 25 Copyright 2022, John Wiley & Sons. 25
2. Novel polyesters utilized in tissue engineering and organs-on-a-chip
Polyesters have been widely used as the material of choice for fabrication of tissue engineering scaffolds due to their degradation properties in vivo, cytocompatibility and ease of fabrication. 11, 26-28 Clinically used polyesters such as poly-L-lactic acid (PLLA) or polycaprolactone (PCL) often have a significantly higher elasticity compared to soft tissues. 29-30 Mechanical mismatch with the properties of native tissues can trigger a variety of adverse response including fibrosis31, and inflammation32 and reduce the fidelity and impaired cell maturation of in vitro models. 33 To address these shortcomings, polyesters with lower Young moduli (i.e. less than 1 MPa) have been of great interest in cardiac scaffold applications. 34 Several variations of polyester chemistries have been explored for scaffold development, including polyhydroxyalkanoates (PHAs), poly(ε-caprolactones) (PCL), poly(poly sebacate)s and poly(diol citrate)s (Figure 2). In this review, our emphasis is on soft polyesters suitable for engineering cardiac and vascular tissues, while a comprehensive overview of polyester biomaterials has been addressed elsewhere. 26
Figure 2. General polycondensation synthesis scheme of common polyesters for soft tissue engineering.

Polyesters are the product of condensation of an alcohol group (shown in black) and a carboxylic containing monomer (shown in blue) along with other organic monomers (shown in pink) a. Chemical structure of polyhydroxyalkanoates (PHA) groups. Given their natural source, length of carbon chains might be variable in different PHAs. 127 Poly(ε-caprolactone) can be fabricated by different chemistries such as b. polycondensation, or c. ring-opening via anionic, cationic or other catalytic agents. 61 d. Polyglycerol sebacate (PGS) polymer is synthesized from one-pot synthesis of glycerol and sebacic acid. 128 Other commonly used polyesters such as e. POC, f. p(OCS) g. POMaC, h. 1,2,4 polymer, and i. PICO are made from the reaction of 1,8-octanediol and e. citric acid79, f. citric acid and sebacic acid g. citric acid and maleic anhydride94, h. 1,2,4-butanetricarboxylic acid and maleic anhydride12, and i. triethyl citrate and dimethyl itaconate129, respectively. j. A type of adhesive polyester made from polycondensation of polyethylglycol (PEG), dopamine, citric acid and maleic anhydride. 126
2.1. Polyhydroxyalkanoate (PHA)
PHAs are biodegradable polyesters, naturally produced by both gram-positive35 and gram-negative bacteria36, in the presence of excess carbon sources and lack of alternative resources such as nitrogen, phosphorous, potassium and magnesium. They are often classified into short- (4 to 5 carbons), medium- (6 to 14 carbons), and long-chained (14 or more carbons) PHAs. 37 (Figure 2a) These polyesters provide multiple opportunities to adjust their utility by functionalization of unsaturated side chains. 38 For example, their hydrophobicity can be adjusted to control cell attachment and drug absorption to the scaffold. 39-40 Another advantage of PHA is their non-toxicity and hydrolytic degradation to carbon dioxide and water. 41 Commonly used PHAs are: poly(3-hydroxybutyrate) (PHB) 42, poly(4-hydroxybutyrate) (P4HB) 43, poly(3-hydroxybutyrate-co-4-hydroxybutyrate) (P34HB) 44 and poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV) 45. Their organic origins and biodegradability make them a great candidate for drug delivery46, biomedical device fabrication47 and tissue engineering48. Degradation of PHAs results in R-3HA groups which can be transformed to 2-alkylated 3HB or β-lactones, giving rise to antibiotic compounds such as carbapenem or macrolide. 49 Hemoembolizing agents such as rifampicin have also been encapsulated in PHB or PHVB microspheres for controlled drug delivery. 50 Tepha FLEX®, a PHB-based suture, has recently been approved by FDA for transplantation. 51 Non-porous PHAs exhibit a significantly higher elasticity (i.e., 3.7 MPa to 739.7 MPa) than myocardium; hence, highly porous cardiac patches from PHAs have been created to mimic properties of the native tissue. 42 Control over degradation52 , elasticity53 and batch-to-batch variability54 are the main challenges to be addressed when using this family of materials. Moreover, challenges with purification of PHAs should be considered for its application in tissue engineered platforms. 55
2.2. Poly (ε-caprolactone) (PCL)
PCL is widely used to its cytocompatibility56, ease of processing57 and viscoelastic properties58. This polyester, can be either derived directly from cyclic ester, ε-caprolactone or indirectly through decomposition of ε-caprolactone to 6-hydroxyhexanoic acid and its polycondensation59 or ring opening of ε-caprolactone60. (Figure 2b, 2c) PCL is soluble in various organic solvents such as chloroform, toluene, benzene, and dichloromethane and immiscible with alcohol and water. 61 As a result, PCL is one of the most common materials used within electrospinning or solvent casting methods. 62 Despite the fact that elasticity of pure PCL is high (i.e., 210-440 MPa) 61, by using different fabrication techniques and manipulating its molecular weight, its mechanical properties can be tuned. 63 PCL has been the frequent material of choice for construction of 3D cardiac64-65 or vascular66 scaffolds. PCL monomers have been frequently combined with other polymers to generate co-polymers, such as poly (lactic-co– ε-caprolactone) (PLCL) 67 or poly(glycolide-co- ε-caprolactone) (PGCL) 68, with modified properties. PCL is highly hydrophobic69, limiting cell adhesion and proliferation on its surface. As a result, the methodologies for fabrication of PCL-based cardiac substrates should overcome their high elastic modulus and surface hydrophobicity.
2.3. Poly (glycerol sebacate) (PGS)
PGS is developed by polycondensation of glycerol, the natural building block of lipids, and sebacic acid, a metabolic intermediate of fatty acid synthesis. 70 (Figure 1a, Figure 2d) Physical properties of PGS can be altered by manipulating monomer feed ratio or degree of esterification. 71 PGS spontaneously crosslinks after exposure to 120°C for a minimum duration of 72 hours; however, crosslinking agents such as methylene diphenyl diisocyanate (DMI) or hexamethylene diisocyanate (HDI) can be added to accelerate this synthesis process. 72-73 Acylating prepolymer to poly (glycerol sebacate) acrylate (PGSA) can provide a UV cross-linkable elastomer for uses such as an on-site surgical sealant. 74 Co-polymers of PGS family have also demonstrated great potential in soft tissue engineering. For instance, bonding of PGSA with hydroxyethyl methacrylate (HEMA) resulted in polyesters with tunable characteristics and shape-memory properties. 75 Although PGS has not yet been approved by FDA in a specific device, both of its monomers as well as similar polyester surgical sealants such as SETALUM™ (Gecko) have received FDA and CE Mark approvals. PGS and its composites were also used in fabrication of cardiac patches to enhance the cardiac function in vivo. 76-78
2.4. Poly (diol citrates)
Poly(diol citrates) are formed from a reaction of a diol- group and citrate group. Depending on the degree of crosslinking, these polyesters can have more physiological relevant mechanical properties for cardiac utility. Poly (1,8-octanediol citrate) (POC), poly (octanediol citrate-co-sebacate) (POCS), poly (octamethylene maleate anhydride) (POMaC), poly (octamethylene maleate anhydride 1,2,4-butanetricarboxylate) (1,2,4 polymer), poly (itaconate-co-citrate-co-octanediol) (PICO) and adhesive polyesters are some of the poly(diol citrates) currently utilized for cardiac or vascular tissue engineering.
2.4.1. Poly (1,8-octanediol citrate) (POC)
POC is a biodegradable and antimicrobial polyester for soft tissue engineering applications. 79 (Figure 1b, Figure 2e) POC has gained significant attention due to its high tensile strength and low elastic modulus. 80-81 POC has been shown to support growth of endothelial cells82 and differentiation of bone-marrow derived mesenchymal stem cells83. This polymer was also used to make composites with PCL, and electrospun to create aligned fibrous sheets for cardiac tissue culture. 84-85 POC/PCL composite is soluble in non-toxic solvents such as acetic acid and formic acid, reducing potential solvent toxicity issues in electrospinning process. 85 Furthermore, therapeutic agents such as lidocaine have been incorporated within the microstructure of POC to obtain tunable drug release kinetics. 86 Under physiological conditions, POC degrades into non-toxic products, citric acid and 1,8-octanediol, which can both be metabolized and eliminated by the body. 87-88 The two POC-based bone implants of CITRESPLINE® and CITRELOCK® (Accutive technologies) have been recently FDA-approved for surgical implantation.
2.4.2. Poly (octanediol citrate-co-sebacate) (POCS)
POCS is a tunable polymer formed from one-pot polycondensation of 1,8-octanediol, citric acid and sebacic acid. 89 (Figure 2f) Viscoelastic properties and degradation kinetics of POCS could be manipulated by changing the monomer ratio of citric acid to sebacate ratio. 89-90 Other groups have created conductive POCS by incorporating up to 5% carbon nanotubes using a novel screw-coating of the material with carbon nanotubes. 91 Pendant carboxyl groups on the surface of crosslinked POCS can regulate cellular attachment95 and covalently bond with extracellular matrix (ECM) molecules to enhance cell attachment. 92 POCS has been electrospun along with fibrinogen to enhance cell attachment and proliferation. 93
2.4.3. Poly (octamethylene maleate (anhydride) (POMaC)
POMaC is a biodegradable elastomer, synthesized through one-pot polycondensation reaction of citric acid, maleic anhydride and 1,8-octanediol. 94(Figure 1c, Figure 2g) Liquid pre-polymer generated through this synthesis, contains several vinyl and ester groups. Vinyl groups enable free radical polymerization with the addition of a photoinitiator and use of an appropriate wavelength of light. 94 In addition to the vinyl groups, POMaC structure is comprised of pendant carboxylic acid and alcohol groups undergoing esterification reactions, a post-polymerization process that is accelerated at higher temperatures. 94 Vinyl groups formed from UV crosslinking are non-polar and hence contribute to hydrophobic behaviour of the products, while ester bonds are highly susceptible to hydrolytic degradation. 95-96 The unique dual-crosslinking mechanism of POMaC creates an opportunity to tune degradation of the final biomaterial by altering monomer feed ratios. 94, 97 Moreover, pendant functional groups of POMaC are binding sites for bioactive molecules such as peptides or protein conjugates. 98
POMaC is a soft elastomer, with tunable properties that can match physiological properties of native myocardium and negligible changes upon cyclic stretching. This makes POMaC a good material for support of contractile behaviour of cardiac tissue both in vivo and in vitro. 18, 23, 99-102 Additionality, the ability to perfuse POMaC within polydimethylsiloxane (PDMS)-based microchannels facilitates microfabrication of a wide variety of micro-scaled geometries. 18, 23, 99 The anisotropic properties of POMaC-based micropatterned scaffolds are used to control cardiac tissue compaction, resulting in better cardiomyocyte elongation and maturation. 18, 23 Anisotropic patches fabricated with POMaC are not only well functional in vitro, but they have shown to significantly improve cardiac functional properties post myocardial infarction (MI) compared to other polymers such as poly(ethyl glycol) (PEG). 18 Other favorable properties of POMaC such as the ability to be partially UV-crosslinked and 3D stamped have resulted in AngioChip, a thick perfusable structure with the capability of forming 3D vascular networks. 99
2.4.4. Poly (octamethylene maleate (anhydride) 1,2,4-butanetricarboxylate) (1,2,4 polymer)
To further increase the elasticity of polymers for cardiovascular applications, we previously synthesized 1,2,4 polymer using a similar polycondensation reaction. 12 (Figure 1d, Figure 2h) One of the major advantages of 1,2,4 polymer is its elasticity variation with monomer composition, porosity and the degree of crosslinking. 12 The relatively lower Young’s modulus of 1,2,4 polymer (44 ± 7 kPa12) falls within the lower limits of mechanical properties of myocardium, while still enabling a reasonable ultimate tensile strength, thereby making this material a great candidate for cardiac tissue engineering. The polymer also degrades both hydrolytically and enzymatically under aqueous conditions and upon culture with cells, suitable for in vitro and in vivo applications. Finally, when compared to POMaC and the FDA approved PLLA, 1,2,4 polymer has demonstrated higher T cell recruitment and similar inflammatory response (with less overall macrophage response compared to POMaC). 12
2.4.5. Poly(itaconate-co-citrate-co-octanediol) (PICO)
PICO is a recently developed biodegradable elastomer, synthesized by a multi-step polycondensation reaction of triethyl citrate (TEC), 1,8-octanediol and dimethyl itaconate. 12 (Figure 1e, Figure 2i) Like POMaC, liquid PICO prepolymer can be injected into PDMS molds with small microchannels to create customizable scaffold designs. 12 Another recent design has looked into modification of patterns and porosity to obtain desirable elasticity matching that of the cardiac tissue. 25 Due to its tunable mechanical properties, ability to be micromolded to various 3D patterns and degradation into immunomodulatory monomers (i.e., citric acid103 and itaconate104), PICO has a great potential for application in scaffolds for various soft tissues and in vivo implantation. PICO degradation releases itaconate (ITA), which has demonstrated capacity as a small molecule to regulate inflammation105 but has been limited in efficacy through oral delivery due to rapid removal from circulation. 106 Synthesis of polyester materials containing itaconate, and subsequent degradation enables slower release of ITA. While yet to be investigated, PICO materials may offer intrinsic capacity to regulate inflammation upon implantation, providing utility in graft adoption of a tissue engineered construct.
2.4.6. Adhesive polyesters
Traditionally, sutures107, bioabsorbable wires108 and staples109-111 were used in surgical applications for holding tissues together, stopping body fluids and enabling healing mechanisms. These methods, not only require an expert surgeon to apply them, but they also promote chronic inflammation112, involve a risk of infection113, and are hard to apply depending on the mechanics of the tissue of interest. 114-115 Inspired by adhesion mechanistic of mussels, dopamine, a protein member of catecholamine family, has been of interest to scientists for creation of novel adhesive biomaterials as a surgical glue. 116-117 Dopamine undergoes a self-polymerization reaction and creates a thin layer of polydopamine which can enhance protein attachment and cell adhesion. 118 Embedding dopamine within the microstructure of many polymers have been shown to increase the adhesion properties of the polymer in the short term. 119-121 Injectable citrate-based mussel-inspired bioadhesives (iCMBAs) were developed by reaction of PEG, citric acid and catechol-containing molecules such as dopamine. 122 This material was applied as a wound dressing in vivo and resulted in the reduction of pro-inflammatory and the promotion of anti-inflammatory response within rodent models. It also led to a more physiological relevant mechanical properties of the skin compared to conventional suturing methods. 122 Later on, they developed an enhanced iCMBA, with reacting 10-undecylenic acid (UA), conjugated to citric acid, dopamine and PEG. 122-123 UA is a natural fatty acid with antimicrobial properties against several strands of bacteria, fungi and viruses, that acts by disrupting membrane integrity within these microbial cultures. 124-125 More recently, we created an adhesive material by one-pot synthesis of citric acid, PEG, maleic anhydride and dopamine for 28 to 72 hours. 126 (Figure 2j) The adhesive patch was demonstrated to have a superior adhesion to the heart tissue in vivo compared to POMaC or fibrin glue. 126 Further research is warranted to explore other adhesive molecules and optimize their properties for potential therapeutic applications such as minimization of inflammatory response and use of 3D human in vitro models for material evaluation.
3. Interplay between polyester chemistry and cardiovascular tissue physiology
Enhanced cell-biomaterial crosstalk25, cytocompatibility 130 and ease of chemical modification make polyesters an optimal choice for designing cellular microenvironments and deriving the desired cell response. In this section, we describe these unique advantages of polyester elastomers and highlight opportunities they provide to the engineered tissues. Table 1 summarizes some of the most important properties of commonly used soft polyesters in comparison to in vivo tissue properties.
Table 1.
Summary of common soft polyester materials, their mechanical properties and degradation kinetics of pure polymer material (prior to fabrication). Below-mentioned values may be modified depending on porosity and design of engineered scaffolds in different applications.
| Name | Young's Modulus |
Tensile Strength |
Elongation | Degradation | Reference |
|---|---|---|---|---|---|
| PHB * | 74.45-554 MPa | 1.3-87 MPa | 3.8-26% | In vitro: less than 10% in 6 weeks | 199-205 |
| P4HB * | 14.5-670 MPa | 2.3-70 MPa | 1000-1450% | In vitro: less than 1% in 28 days In vivo: 6 to 12 months | 43, 178, 206-207 |
| P34HB * | 13.9-902 MPa | 1.23-24.3 MPa | 7-17% | In vivo: no observed degradation over 8 moths | 208-209 |
| PHBV * | 67.7-106.7 MPa | 4.01-5.84 | 50.2-56.3% | In vitro: up to 16% in 6 weeks. | 210-213 |
| PCL | 30-241 MPa | 25.1-29.4 MPa | 450-772% | In vitro: less than 1% in 6 months | 206, 214-215 |
| PLCL | 8.4±0.9kPa | 4.7±2.1 kPa | 960±270% | In vitro: ~40% in 50 weeks In vivo: 61% after 24 weeks | 216-217 |
| PGCL | 292.98-263.26 MPa | 288.5-380 kPa | 43.7-59.2% | In vitro: 20-40% in 40 days | 218-219 |
| PGS | 0.282±0.0250 MPa | >0.5 MPa | 267±59.4% | In vitro:17 ± 6% after 60 days In vivo: 100 after 60 days | 220-221 |
| PGSA | 0.05-1.38 MPa | 0.05-0.5 MPa | 42-189% | In vitro: 15% after 10 weeks In vivo: fully degraded after 6 weeks | 74 |
| POC | 0.49-3.92 MPa | 6.1 ± 1.4 MPa | 40-50% | In vitro 100% after 15-68 week In vivo: 20% after 28 days | 22, 81 |
| POCS | 0.19-1.1 MPa | 0.2-0.6 MPa | 160-230% | In vitro: 9-70% in 4 weeks | 11, 90 |
| POMaC | 0.04 - 0.29 MPa | 245-611 kPa | 48 - 534% | In NaOH: 100% after 4 to 12 hours In vitro: 100% after 12 weeks | 94 |
| 1,2,4 polymer | 44±7 kPa | 34 ±13 | 99±32% | In vitro 60% (porous) and 40% (pure polymer) after 14 days Co-culture with cells: 60% (porous) and 30% (pure polymer) after 14 days | 12 |
| PICO | 36–1476 kPa | 50 -320 kPa | 10-45% | In NaOH: 25 to 80% (tunable) after 4 to 12 hours | 129, 222 |
| ITA polymer | N/A (liquid) | N/A | N/A | Enzymatic environment: 30-40% within 28 days. | 138 |
| Human left ventricle | 60–800 kPa | 2.51± 0.21 MPa | 34.9±1.1% | N/A | 223-224 |
| Rat Left ventricle | 20–54 kPa | 200-400 kPa | 100-175% | N/A | 21 |
| Rat aorta | 0.17-0.98 MPa | 0.40-1.88 MPa | 318±27% | N/A | 225-226 |
These values might vary as natural sourced PHA polyesters have batch-to-batch variations.
3.1. Cell adhesion mechanisms and surface modification
Cellular adhesion and tissue formation can be affected by surface chemistry or bulk material properties. Cell adhesion is a complex process involving surface receptors and their ligands. Integrin receptors are the key receptors for cell adhesion to ECM or scaffold materials and are heterodimers composed of two subunits: α and β. 131 In the extracellular domain, integrin receptors bind to the binding sites of ECM proteins derived from fibronectin (RGD, REDV, KQAGDV, and PHSRN),laminin (LRE, IKLLI, PDGSR, IKVAV, LRGDN, LGTIPG, and YIGSR), collagen (DGEA and GFOGER) (from collagen) or elastin (VAPG). 132 Upon binding of the extracellular domain, cytoplasmic protein talin binds to the β subunit of integrin and links it to actin cytoskeleton, initiating cytoskeletal remodeling. 133 Various surface modification techniques have been leveraged to enhance cell binding to the surface using reactions such as 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide/N-hydroxysuccinimide (EDC/NHS). 134 Many of the polyesters contain carboxylic group (-COOH) can be functionalized using EDC/NHS chemistry to immobilize biomolecules and crosslink integrin binding domains in order to improve cell adhesion. 135 For instance, PCL/ poly(m-anthranilic acid) (P3ANA) copolymer was functionalized by RGD-group to enhance cell attachment and proliferation on the scaffold. 136 Other groups have incorporated bioactive components within POC polyester and observed an enhanced vascularization of the in vivo model. 137
Given the wide choice of these building monomers, polyesters also have a unique ability to incorporate bioactive molecules like lidocaine86, itaconate138 or dopamine123, 126 within their backbone structure. These bioactive components will drive application-specific mechanical and physical properties. For instance, dopamine-embedded polymers had an enhanced ability to attach to surfaces123, 126 and consequently could be used as a surgical glue to hold cardiac patches in place. 77 Furthermore, highly polar groups in polyesters and their entanglement allow the physical incorporation of nano-scaled materials within polyester lattice. For example, Ahadian et al. incorporated carbon nanotubes (CNT) within 1,2,4 polymer and showed electrical conduction within the resultant polyester composite. 139 This nanocomposite was later used as a substrate for cardiac tissue engineering to enhance excitation threshold and induce cardiac fiber alignment. As soft polyesters are further explored in tissue engineering applications, modification of these materials with bioactive components would provide researchers a greater ability to model native tissue environment.
Bulk material properties such as porosity, hydrophobicity, surface energy, topography and elasticity can affect cellular attachment. 135 Porosity is another important factor as it can influence tissue integration and vascularization. 102, 140 Highly porous materials are generally less stiff compared to their pure solid counterparts, thus facilitating cell infiltration, neovascularization, and improved transport of nutrients and waste removal. 141 As an example, AngioChip platform, a POMaC-based blood vessel network on a chip, has incorporated both macro- and micro- porosity to enable cell extravasation and small molecule infiltration from the engineered micron-sized vessels. 99 Creation of patterned superhydrophilic and superhydrophobic surfaces can also engineer cellular attachment into desired shaped142. Cardiomyocytes attachment can also be achieved through topographical patterning of the surface. For instance, Au et al. have unidirectionally abraded surfaces to achieve parallel microgrooves and induce cellular alignment in cardiomyocytes. 143
3.2. Mechanotransduction and polyester mechanical properties
Engineered scaffolds provide a unique ability to design 3D mechanical cues and hence drive cellular responses resembling the native tissue. 144 Material elasticity145, porosity146 and topography147-148 are often optimized in tissue engineering to match the mechanical properties of the targeted tissue or the disease microenvironment. Mechanical properties of the cellular environment can activate different pathways related to cell migration 149, differentiation150, ECM deposition and degradation151-152, cytoskeletal arrangement149, muscle contraction153-154 and cellular maturity155 could be affected. For in vivo applications, implantation of stiffer materials has demonstrated impact on macrophage driven inflammation, ultimately leading to fibrosis. 156
In general cells can sense and respond to mechanical properties of their surrounding materials via intracellular factors such as talin and vinculin in processes known as mechanotransduction. 157 In the case of cardiac tissue, activation of these pathways causes rearrangement of actin and myosin fibers in cardiomyocytes, and hence affects cardiac contraction, alignment, and maturation. Mechanosensing is not limited to cardiomyocytes; cell subpopulations in cardiac tissue such as stromal cells, e.g. fibroblasts, undergo phenotypical changes when subjected to stiff substrates. 158 Topography or surface features, can be another driver of cellular behaviour. It has been shown that both micrometer25 and nanometer159 scaled topography can drive cardiac alignment and assist in formation of highly organized functional cardiac tissues. 143 Thus, when designing a biomaterial for cardiac tissue engineering, it is critical to biomimetically design these factors to minimize the impact on functionality of the cardiac tissue.
Polylactide and polylactone family (such as PCL, poly(D-lactide) (PDLA) and PLLA), despite their FDA approval in several medical devices, have a considerably high elasticity (~0.65-2.7 GPa160-161) compared to the cardiac tissue, which hinders their ability to recapitulate native tissue properties. 94, 162-163 Despite the fact that stiffer tissues such as trabecular and cortical bone (elasticity of ~10.4-20.7 GPa) 164 or human articular hip cartilage (elastic modulus of ~0.67-1.8 MPa) 165, have relatively similar mechanical properties, elasticity of many soft tissues are considerably lower. For instance, the elastic modulus is ~40-180 kPa for skin 166, ~280-300 for thoracic and abdominal aortas167, ~64-112 kPa for the bovine spinal cord 168 and ~200-500 kPa for the human myocardium in diastolic and systolic state. 169-172 Matched elasticity with the target tissue could prevent tissue deformation and significant irritation within the microenvironment. 173-174
Several biodegradable polyester materials such as POMaC and PICO offer the flexibility to tune their mechanical properties to closely resemble the native tissue. This can be achieved by adjusting the nano- and micro-scaled porosity of the materials as well as modifying the synthesis conditions and monomer ratios. 18, 94, 99, 129
3.3. Biodegradation
Depending on the chemical structure and surface area of polyester scaffolds, water can penetrate the scaffolds to different degrees, leading to hydrolysis of the polymers over time. 13 Controlled polymer degradation during integration with host is a critical factor for cardiovascular patches. Polyesters have attracted a lot of attention due to their moderate biodegradation enabling temporary implants175 and drug delivery systems176. Polyester degradation involves the breakage of ester bond to carboxylic acid and alcohol groups. 177-178 The increase of solution acidity will further autocatalyze degradation with raising the nucleophilic properties of water. 179 Degradation studies have been also conducted in basic conditions to accelerate hydrolysis process and predict long-time degradation capacity of polymers. 94, 129, 180 In other attempts, cells have been cultured adjacent to the material to capture crosstalk between the polymer and a specific cell type. Porous polyester structures have shown enhanced water absorption and hence an accelerated biodegradation. 12 Davenport Huyer et al. showed a greater mass loss in 1,2,4 polymer and POMaC in porous structures compared to the non-porous conditions. 12 The degradation process was further accelerated when scaffolds were further cultured with neonatal rat cardiomyocytes. 12 Degradation dynamics differs in vivo according to specific body conditions such as the local acidity of host tissue, mechanical loads181 and the cellular response to the implant182. The ideal kinetics of polymer degradation is also both species- and tissue- dependant. For instance, an ideal degradation of polymers for myocardial regeneration applications would happen between 1 week to 6 months in rodent cardiac patches aiming to restore tissue functionality post-MI. 183 This will allow sufficient time for cell remodelling whilst minimizing immune response to invasive materials, and matches the period needed for a complete remodelling post infarction. This timeframe would be at least 3 months in humans given the slower pace of remodelling and the larger surface area of human heart. 184
Polyesters degradation would ideally result in monomers that are naturally occurring in the body like citric acid94, glycolic acid76, and sebacate90. Citric acid, for instance is a crucial compound in the citric acid cycle (Kerb cycle), a fundamental cellular respiratory pathway. 185 Similarly, glycolic acid is part of glycolysis pathway, a metabolic pathway for glucose uptake. 186 Other polyesters, also degrade into molecules that can be easily metabolized by the body. For instance, sebacic acid and octanediol can be enzymatically digested in the body. 187 PCL can also degrade into caprolactones, further hydrolyzed into 6-hydroxyhexanoic acid, a naturally occurring carboxylic acid. 188 In conclusion, ideally the degradation of polyesters would result in chemicals that are recognized and metabolized by the body, hence reducing the risk of adverse reactions and improving biomaterial-host tissue integration. Polymer degradation also increases the porosity of the scaffold and hence promote the crosstalk between the tissue and biomaterial. For instance, several groups have attempted to drive tissue remodelling in situ after myocardial infarction (MI). Further work is needed to study the effect of monomer ratios on degradation properties, enabling engineers to fine-tune degradation kinetics upon translation into clinical applications.
3.4. Inflammation, proliferation, and tissue remodeling
The response of host tissue to implanted biomaterial has three main stages: inflammatory, proliferative, and remodelling phase. During the inflammatory phase, monocytes are activated to pro-inflammatory macrophages (M1 macrophages) after exposure to growth factors such as transforming growth factor beta (TGF-β) or monocyte chemoattractant protein (MCP)-1 and preliminary matrix layer is deposited on the biomaterial. 189 This is followed by the proliferative phase, in which anti-inflammatory macrophages (M2 macrophages) are activated by exposure to cytokines such as interleukin (IL)-4, IL-10, IL-14. 190 In this stage, other cell types such as fibroblasts, mesenchymal stem cells, smooth muscle progenitor cells and endothelial progenitor cells are summoned to the surface to deposit ECM proteins such as collagen and laminin and promote vascularization. 191 In the tissue remodelling phase, proteins from matrix metalloproteinase (MMP) family are secreted to facilitate ECM remodelling and rearrangement and resolution of inflammatory response. 192
Polyesters can be designed to modulate the regenerative capacity of the host tissue. 193 These materials also provide a unique opportunity for surface modification by incorporating different biologically active components such as peptides or growth factors, creating a selective environment for cell attachment. 194 For instance, ITA is a well-established molecule that plays a critical role in metabolic reprogramming of macrophages. 195 Metabolism of M1 macrophages is characterized by a broken tricarboxylic acid (TCA) cycle (Kerb cycle). 196 This leads to an accumulation of citrate and succinate through inhibition of isocitrate dehydrogenase (IDH). 196 Upon accumulation of citrate over time, citrate is metabolized to ITA, inhibiting succinate dehydrogenase (SDH) and activating anti-inflammatory pathways by engagement with cysteine-reactive protein residues to reduce macrophage inflammation. 197 Incorporation of ITA in the polyester backbone allows controlled release of the molecule and anti-inflammatory activity within the host. 138
Degradation of citrate-based polyesters provides similar bioactivity; the released citric acid can be absorbed by the adjacent cells as a metabolic fuel thereby creating better tissue integration and enhanced angiogenesis. 198 While there have been limited attempts as using polyesters as regenerative therapy, further work can be done on incorporation of the compounds targeting different phases of inflammatory cascade to drive tissue regeneration. Additionally, current polyesters need to be further studied for long-term safety and other regulatory requirement approvals.
4. Tailoring structure of polyester elastomers via microfabrication and 3D printing
Elastomeric polyesters discussed above can be crosslinked through different mechanisms such as UV23, 99, 102, 227-228 or heat17, 229-230 treatment, laying the foundation of several innovative techniques for fabrication of soft polyester-based scaffolds. 231-232 Methods such as electrospinning, laser ablation, micro-molding, 3D stamping and 3D printing have have emerged as effective approaches for achieving such scaffolds with high-resolution and accuracy at the microscale.
Electrospinning is a process where electrically charged streams of dissolved polymers are subjected to a high-voltage electric field, resulting in the production of elongated nanofibers as the solvents evaporate. Various polymers have been electrospun to create scaffolds with tunable mechanical properties and high porosity. 233-239 However, for soft tissue engineering, only a few candidates such as PGS, have been shown to be adaptable to the electrospinning process. 17, 238-241 The fabrication of PGS via electrospinning is challenging due to its low molecular weight and limited number of organic solvents that can dissolve cured PGS. 238 The addition of spinnable polymers such as PCL239-240, poly(methyl methacrylate) (PMMA) 241, polyvinyl chloride (PVA) 17, 242, or PLLA238 have been explored to overcome these challenges. (Figure 3a) Electrospinning of POC/PLCL84 or POC/PCL85 composites as well as soft polyester-urethane236-237 have also shown a potential in the fabrication of soft scaffolds. Laser ablation involves using brief laser bursts in the ultraviolet spectrum to rupture polymer chains, creating photo-ablated cavities. 243 This technique is commonly used for processing hard polyesters. 16, 244-246 Nevertheless, researchers have implemented laser ablation in the fabrication of PGS-based scaffolds with diamond-shaped holes for soft tissue engineering. 15, 21, 247 Laser ablation can also be used to create microholes on PICO or POMaC microtubes to enhance permeability and cell communication. 232 (Figure 3b) While laser ablation is a rapid and flexible approach, the surface finish quality, however, is poor and bulge formation along the scan route is a common problem in this technique.
Figure 3. Microfabrication approaches to tune micro-scale structure of polyester elastomers;

a. Electrospinning of PGS/PCL composite to create random or aligned fibers. Copyright 2014, Wiley Publication Group. 239 b. Creation of PICO/POMaC-based microtubes by (i) coaxial 3D printing in a Pluronic bath, followed by (ii) laser ablation to create micro-holes and provide higher permeability. Copyright 2022, Wiley Publication Group. 232 c. AngioChip assembly through (i) soft-lithography of SU-8 positive molds to create PDMS negative mold, (ii) reversible bonding of PDMS to PDMS or PDMS to glass, and injection of POMaC prepolymer into microchannels, (iii) UV treatment of for partial cross-linking and solidification. Copyright 2018, Nature Publication Group. 248 d. 3D stamping of AngioChip by (i) uncapping the top PDMS from both platforms, aligning the channels and stacking them up, and gently pressing under UV light to bond the layers, (ii) detaching the top PDMS mold and repeating the process to get multilayers. Copyright 2018, Nature Publication Group. 248 e. SEM image of a microchannel in (i) multi-layered AngioChip and (ii) features of each layer. Scale bar represents (i) 500μm and (ii) 200μm. Copyright 2018, Nature Publication Group. 248 f. SEM images of an AngioChip containing micro-holes with different sizes. The scaffold has shown different mechanical elasticity and anisotropy based on the size of the macro-porosity. Scale bars represent 1mm and 200μm (inset) for design A, 1mm and 200μm (inset) for design B and 1mm and 300μm (inset) for design C. Copyright 2016, Nature Publication Group. 99
Molding via microfabricated PDMS structures allows the development of injectable scaffolds with micron-scaled features. 18, 25, 129, 139 The process involves creating a mold using photolithography, adding PDMS to the mold to create the required indentations, and placing the PDMS mold onto a substrate to create microchannels/cavities. 97, 248 The polyester of interest is then added to the PDMS channels, and UV light or heat is used to crosslink the polymer. (Figure 3c) This technique has been used to create scaffolds with various lattice designs and pore sizes. 12, 18, 25, 126, 129, 139 3D stamping integrates soft polyesters into the fabrication of vascularized myocardium and multi-organs connected through endothelialized vasculature. 24, 99, 227, 249-250 The fabrication process involves creating polyester layers using PDMS molds followed by aligning and bonding the individual layers through UV or heat crosslinking.
3D stamping enables producing a perfusable and permeable tube, known as AngioTube, vascularized with endothelial cells. 249-252 (Figure 3d) Similarly, multiple polyester layers with microchannel cavities can be 3D stamped to create AngioChips, which offer branching vascular lumens for tissue engineering and organ-on-a-chip applications. 99, 248 (Figure 3e) The resulting AngioTubes or AngioChips can be incorporated into well plate-typed bioreactors, where the tubes extend over multiple wells to form an integrated vasculature for studying dynamic events (InVADE). 24 Control of lattice shape in such platforms allows matching apparent elasticity and scaffold anisotropy to those of the native tissues. (Figure 3f) Despite their potential, PDMS micro-molding and 3D stamping methods face challenges in scalability, requiring complex equipment, manual skill, and time-consuming fabrication processes. Additionally, the rectangular cross-sections of the produced tubes and channels limit their ability to mimic native vascular tissue accurately.
3D printing of polyesters is challenging due to their low elastic modulus and long gelation time. Secondary hydrogels, such as gelatin or Carbomer, are used as supporting materials during the printing process and can be removed later. 232, 253-254 Techniques like extrusion-based 3D printing254 and the freeform reversible embedding of suspended hydrogels (FRESH) method232 have been employed to print polyester structures and vascular tubes. (Figure 3b) Stiffer UV crosslinkable polyesters such as poly(propylene fumarate) (PPF) 255, PCL256-257 and POMaC/poly(ethylene glycol) diacrylate (PEGDA700) composite258 were also printed using stereolithography (SLA) techniques.
5. In vitro and in vivo applications of soft polyesters
Soft polyesters have a great potential for fabrication of in vitro cardiovascular models, as well as scaffolds that can be implanted in vivo. In this section, we will review recent advances in using polyesters in cardiovascular tissue engineering models and their advantages. Although these models can properly recapitulate native environment, their translation into clinical use and pharmaceutical applications requires further studies on the application-specific optimizations of scaffold properties.
5.1. In vitro
In vitro models aim to recapitulate the functionality of native tissues in the format of organ-on-a-chip devices or larger-scaled 3D tissues. Cardiac-on-a-chip models should be able to produce a contraction force of 2-4mN/mm2 in a cyclic manner and transmit electrical signals at a ~25cm/s. 259 Although natural polymers such as gelatin, fibrin, alginate, and hyaluronic acid can provide ECM-like microenvironments for different cell types, they are often lack mechanical stability and the elastic modulus of the cardiovascular tissue. 74 Synthetic polyesters, on the other hand, offer several advantages such as mechanical strengths, tunable elasticity, controllable degradation and ease of fabrication. 139, 259
5.1.1. Organ-on -a-chip
Conventionally, organ-on-a-chip platforms are fabricated with PDMS due to its biocompatibility, optical transparency, gas permeability and high elastic modulus. 260-262 However, PDMS can non-specifically adsorb biomolecules due to the hydrophobic interactions, resulting in variations in the concentrations of culture media compounds and the introduced drugs/reagents. 263 Polyester scaffolds have been carefully designed to provide structural support, imitate native ECM, and promote cell adhesion, proliferation and differentiation139, 263 and provide the capability of tuning their mechanical properties, degradation profile, tissue culture conditions, fabrication technique and porosity. 99
Given the low elastic moduli of POMaC and its mechanical stability, this polymer can be cast into microscale wires and assembled into miniature cardiac tissue culture wells with two microwires spaced apart, giving rise to Biowire platform. Autofluoresent properties of POMaC allows tracking of wire displacement and correlating it to cardiac force, hence warrants a non-invasive in situ method for tracking cardiomyocyte contraction and electrophysiology. 264 Long term electrical stimulation of the obtained tissues would result in cardiomyocyte maturation101-102, 264, modelling of complex diseases such as cardiac fibrosis100, 265 and more physiological relevant drug testing results101-102, 264. (Figure 4a) Biowire platform, acquired by Valo Health, is currently being used in industry for drug discovery and development.
Figure 4. Applications of polyester elastomers in cardiovascular tissue engineering and organs-on-a-chip.

a. Biowire is a cardiac-on-a-chip model that (i) enables studying complex disease mechanisms such as fibrosis. Copyright 2019, ACS Publication Group. 100 (ii) This model allows for a formation of a dense cardiac tissue in between parallel POMaC wires. Scale bar represents 1mm. Copyright 2019, Elsevier. 264 (iii) Comparison of electrical propagation patterns in healthy and fibrotic tissue. Scale bar represents 500μm. Copyright 2019, ACS Publication Group. 100 (iv) Collagen deposition was observed in the control sample (left) vs fibrotic Biowire model (right). Scale bar represents 100μm. Copyright 2020, Nature Publication Group. 290 (v) Different sides of the Biowire could be seeded with atrial and ventricular cardiomyocytes to study their crosstalk. Scale bar represents 0.5mm. Copyright 2019, Elsevier. 264 b. AngioTube is a polyester-based microtube. (i) fabricated by 3D stamping of a hollow bottom and flat top structures (Copyright 2021, Nature Publication Group24) to give rise to (ii) micro-scaled porosity. Scale bar represents 200μm Copyright 2021, Nature Publication Group. 24 (iii) AngioTube is assembled on a polystyrene sheet and capped with a bottom-less 96-well plate to create gravity-driven flow. Copyright 2021, Nature Publication Group. 24 (iv) seeding of parenchymal tissue around the AngioTube. Scale bar represents 200μm. Copyright 2017, John Wiley & Sons. 227 (v) Immunostaining for cardiomyocyte marker, sarcomeric alpha-actinin (Scale bar represents 200μm. Copyright 2017, John Wiley & Sons. 227 (vi) Cardiac marker, Troponin-T, and nuclei marker, DAPI, staining of AngioTube after exposure to 50μgml−1 of SiO2 nanoparticles for 24hr. Scale bar represents 50μm. Copyright 2017, John Wiley & Sons. 249 c. mm-scaled 3D left ventricle model (i) fabricated by rolling a flat seeded patch around a central mandrel to create (ii) a 3D model capable of containing fluid within its cavity. Scale bar 1cm. (iii) Computer aided (CAD) model of the fabricated in vitro left ventricle. (iv) staining of alpha-sarcomeric actinin and F-actin filaments on the scaffold and (v) cardiomyocyte crosstalk between different microgrooves due to the unique design of the patch. Scalebar (iv)100μm and (v)50μm. Copyright 2022, ACS Publication Group. 25 d. Flexible shape-memory scaffold that can be (i) rolled within a 1 mm diameter opening needle and (ii) injected in situ, while maintaining its original shape. Scale bar (i) and (ii) 5mm.(iii) CFDA (live/green) and propidium iodide (PI)(dead/red) staining of the patch seeded with cardiomyocytes after injection. Scale bar 2.5mm. (iv) implantation of stem-cell derived cardiomyocytes on porcine heart. (v)SEM images of seeded patch with after implantation on porcine epicardium. Scale bar 1mm (main) and 100μm (inset) (vi) comparison of myocardial thickness of rodent model with (left) and without (right) implantation of the patch. Copyright 2017, Nature Publication Group. 18
Vascularization of the engineered microtissues plays a crucial role in their functionality when studied on a chip, since the cell metabolism is highly susceptible to exchange rate of nutrients, metabolite, and oxygen in such organs. AngioChip platform, developed by our group, is capable of creating permeable vascular lumens for organ-on-a-chip engineering. 99 This technology combines two unique features: (1) a polymer-based 3D branched microchannel network with thin and permeable, yet mechanically stable walls to support the built-in vasculature coated with endothelial cells, and (2) a hydrogel embedding the cells which is cast into the polymer mesh around the network such that the parenchymal cells remodel the matrix and compact around the built-in vasculature to form a functional tissue. POMaC-based polymer lumens of AngioChip, with intricate pore structures are controlled from the nanometer to millimeter scale to model transfer of molecules and cells between vasculature and parenchymal space. 99 The lumen could be confluently endothelialized, and vessels are able to sprout in response to an angiogenic stimulus. AngioChip was utilized to recapitulate vascularized cardiac tissue and synchronized beating that could macroscopically compress the chip without damaging the vascularized network during the perfusion. Application of AngioChip with HEPG2 liver cell lines also allowed investigations of urea secretion and terfenadine metabolism in vitro. 99
As an alternative approach, a single porous microchannel, AngioTube, could be fabricated by combining micromolding and 3D stamping techniques. AngioTube was embedded in 96-well plates to create InVADE system. 24, 227, 266 A programmed rocking system facilitated fluid circulation in the wells, relying on gravity-driven flow. The channel permeability, cell migration, and vessel sprouting is enabled in AngioTubes through nano and micro scaled designed porosity. 227. Lai et al. recapitulated liver, cardiac tissue, and cancer invasion cascade via InVADE platform. 227 The open 96-well plate concept enables ease of tissue extraction and seeding. In later studies, InVADE system was utilized to investigate the effect of CuO and SiO2 nanoparticles on cardiovascular system under perfusion. 249 The platform demonstrated that the release of reactive oxygen species (ROS) and secretion of pro-inflammatory cytokines under the influence of nanoparticles brought about electrical and contractile dysfunction of cardiac tissue. 249 InVADE platform has also been utilized for the evaluation of chemotherapeutic medication, gemcitabine, on pancreatic ductal adenocarcinoma (PDAC) co-cultured with stromal fibroblast. 266 The result demonstrated an enhanced tumor viability and higher dose requirement for PDACs in perfused conditions, compared to stationary tissue culture. 266 In another study, InVADE was employed to assess the impact of SARS-CoV-2 on the endothelialized AngioTubes and therapeutic efficacy of an antiopoietin-1 derived QHREDGS peptide. 250, 267 By circulation of immune cells in the engineered vasculature, cytokine storm was induced, where vascular dysfunction is advanced due to the over-secretion of cytokines from activated endothelial cells and monocytes.135 Three-dimensional printing of polyester microtubes has also been shown effective in organ-on-a-chip applications, enabling cardiomyocytes and endothelial cell attachment, upon placement of the tubes into customized 96 well plates. 232, 254 (Figure 4b)
5.1.2. In vitro cardiac tissue models
Three-dimensional printing of CMs embedded in polyester-based scaffolds 268-269 has resulted in successful cm-scaled heart models. 268-269 However, many of these models are still incapable of recapitulating spatial changes and physiological phenomena such as ejection fraction and MI. Zhang et al. developed a 3D scaffold made of multiple 2D POMaC scaffold meshes assembled together through a hook and loop system, termed as Tissue-Velcro. 23 The design provided anisotropic mechanical properties and allowed for culture of multiple cell types. Moreover, the disassembly of the layers with preserved structure for further analysis of the cells was quite straightforward. 23 Other studies demonstrated that up to 0.5wt% carbon nanotubes (CNTs) could be embedded in 1,2,4 polymer to increase the electrical conductivity of the fabricated patches and hence improve cardiac electrical excitability. 139. Similarly, in a model demonstrated by Mohammadi et al., the myocardial architecture of left ventricle and its various myofiber anisotropic angles were recapitulated in vitro. 25 (Figure 4c) The model comprised multi-stacked 2D CMs coated sheets assembled as a conical cardiac ventricle. The design incorporated both elastic PICO as a backbone and a collagen/Matrigel hydrogel for encapsulating CMs and enhancing cell attachment. The embedded holes in each sheet enhanced the delivery of oxygen and nutrients and promoted cellular crosstalk. 25 Despite the significant work in 3D tissue models, further work needs to be done to create better physiologically relevant cardiac models with the ability to reconstruct the complexities of native organs.
5.2. In vivo applications
Poor biological integration, low biocompatibility, and high chance of fibrosis often hinder application of engineered scaffolds in vivo. 270-274 Given the low elastic modulus of human myocardium during the diastole and systole cycles (10-300 kPa), an ideal elastomer for implantable patches should possess similar Young’s modulus, as well as high elongation and tensile strength in order to support contractile behavior of cardiac tissue. 12, 97, 275-276 Implementation of cardiac patches with better physiological integration has raised a great interest towards restoring the functionality of the heart upon injury. 277 Patches could be made of a combination of synthetic polymers and natural hydrogels278-279 280 259, 281-285 A landmark study by Zimmermann et al. demonstrated that the use of cardiac patches, made of heart cells incorporated in collagen I and Matrigel, on the epicardial surface of the heart can result in functional improvement post-MI. 286 One issue with the current cardiac patches is the invasiveness of open-heart surgery for placement of the patch on the cardiac surface, limiting their applicability and posing risks such as chest wound infection. Montgomery et al. have developed a shape-memory scaffold constructed of POMaC which could enable precise injection of fully functional tissues to the heart, noninvasively. 18 The scaffold with mechanical anisotropy could be delivered through a 1mm I.D. needle, recovering its initial shape upon injection without affecting cardiomyocyte viability or function. The patches could significantly improve the cardiac function in post-MI rats, and demonstrated similar vascularization, macrophage recruitment and cell survival compared with open heart surgery. 287 POMaC-based AngioChips, cultured with neonatal rat cardiomyocytes, were also shown to enable direct surgical anastomosis with different setups, artery-to-artery or artery-to-vein and support cardiomyocytes elongation and mural cell penetration after a week. 99 (Figure 4d)
One important issue with cardiovascular grafts and catheters is contamination and bacterial infection. 288 Integration of ITA, an anti-microbial compound, into the backbone of polyesters could potentially reduce the chance of infection in the implants made of such materials. 129, 138 289 In PICO and ITA polymer, a stable release of ITA was observed through degradation of the polyester backbone in neutral hydrolytic or slightly alkaline conditions (pH=8) mimicking the body environment. 129, 138
6. Conclusions and Future Perspectives
Soft polyesters often exhibit desirable properties including low elastic moduli97, versatile methods of crosslinkability 291, short term stability, long-term biodegradation101 and ability to crosslink bioactive chemicals to the polymeric backbone97. These properties enable tissue engineers to design higher fidelity models using these polymers. 21, 28, 102, 264 Although many soft polyesters such as PGS220, POC22, POMaC94, 124 polymer12 and PICO129 have been developed in the past 20 years with superior biodegradability and tissue mimicking elasticity, their long-term response in the host needs to be further studied prior to clinical translation. On the other hand, some of these materials have recently been FDA-approved in specific devices, further emphasizing the potential of these materials in implantation. They have also been incorporated into organ-on-a-chip devices such as Biowire264, AngioChip99 or AngioTube227 some of which are currently used in the market for drug development and assessment. Yet, mainstream applications require scalable fabrication techniques. Figure 5 summarizes some of advantages and constraints of commonly utilized polyesters for cardiac engineering.
Figure 5. Comparison of advantages and shortcomings of each polyester group for fabrication of engineered cardiac tissues.

To overcome these challenges, several strategies can be explored for development of novel chemical or fabrication techniques. Current polyesters seem to have a mechanical mismatch in terms of the visco-elastic properties with the myocardium, resulting in poor integration between the fabricated scaffolds and native myocardium. (Figure 5) The ability to fine-tune mechanical properties of polyesters using different molecular weights of monomers, degree of crosslinking and incorporation of porosity can be leveraged to achieve more cardiac-friendly constructs. Additionally, techniques such as surface modification can be applied to better mimic factors present in native ECM and establish cellular adhesion and communication with the biomaterial. Degradation kinetics along with the ability to incorporate a wide range of monomers in polymeric backbone allows for controlled release of different biochemical agents such as immunomodulatory129, 138, 193, conductive139, or adhesive compounds126. Incorporating other factors important in mediating the inflammatory response of the body and tissue remodeling in the material will be beneficial for scaffolds designed to be implanted. For instance, following a myocardial infarction, spatiotemporal release of anti-inflammatory and proangiogenetic factors in the myocardium could enhance regeneration process and potentially restore the native tissue. 292 Other strategies such as co-polymerization of different polyester monomers could also be beneficial in such applications.
Biofabrication techniques also provide a distinctive opportunity to fabricate unique bioengineered designs for in vivo and in vitro models. Currently, electrospinning, laser ablation, micromolding, 3D stamping and 3D printing are some of the most common methods of scaffold fabrication. However, due to relatively long crosslinking time of polyesters12, 22, 94, 129, 220 additional methods of high throughput fabrication such as droplet-based bioprinting, extrusion bioprinting and stereolithographical bioprinting could be applied to obtain higher resolution features, in a more reproducible manner. Delivery of the biomaterials into affected tissue could also be improved by enhancing the shape-memory properties of the designed scaffold. Currently an open-heart surgery is required for application of cardiac patches in vivo, which requires a longer recovery time and could result in in-patient post-surgical complications. Shape-memory properties of designed scaffold will enables application of cardiac patches with minimal invasiveness. 18 Finally, more thorough understanding of long-term effects of these constructs in vivo is vital to enable transition into clinical trials.
In summary, recently developed polyesters have demonstrated a great potential for fabrication of in vitro cardiovascular models102, 232, 264, as well as implantable patches and grafts18, 23, 99, 248, 293. Given their potential, more research in this area is needed to explore novel chemistries, fabricate the models in higher throughput and functionalize them with application-specific bioactive compounds for regenerative medicine or drug delivery purposes.
Acknowledgement
Our work is funded by the Canadian Institutes of Health Research (CIHR) Foundation Grant FDN-167274, Natural Sciences and Engineering Research Council of Canada (NSERC) Discovery Grant (RGPIN 326982-10), NSERC-CIHR Collaborative Health Research Grant (CHRP 493737-16), National Institutes of Health Grant 2R01 HL076485. M.R. was supported by Killam Fellowship and Canada Research Chair. SO was supported by CIHR Canada Graduate Scholarship. AS is supported by NSERC Post-doctoral Fellowship.
8. References:
- (1).Khanna A; Zamani M; Huang NF Extracellular Matrix-Based Biomaterials for Cardiovascular Tissue Engineering. J. Cardiovasc. Dev. Dis 2021, 8 (11), 137–137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (2).Ratner BD; Zhang G A history of biomaterials. In Biomater. Sci, Elsevier, 2020; pp 21–34. [Google Scholar]
- (3).Apple DJ; Sims J Harold Ridley and the invention of the intraocular lens. Surv. Ophthalmol 1996, 40 (4), 279–292. [DOI] [PubMed] [Google Scholar]
- (4).CHARNLEY J; HALLEY DK Rate of wear in total hip replacement. Clin. Orthop. Relat. Res 1975, 112, 170–179. [PubMed] [Google Scholar]
- (5).MULLIKEN JB; GOLDWYN RM Impressions of Charles stent. Plast. Reconstr. Surg 1978, 62 (2), 173–176. [DOI] [PubMed] [Google Scholar]
- (6).Mertz L. What is biocompatibility?: a new definition based on the latest technology. IEEE pulse 2013, 4 (4), 14–15. [DOI] [PubMed] [Google Scholar]
- (7).Ratner BD; Zhang G 1.1.2 - A History of Biomaterials. In Biomaterials Science (Fourth Edition), Wagner WR, Sakiyama-Elbert SE, Zhang G, Yaszemski MJ Eds.; Academic Press, 2020; pp 21–34. [Google Scholar]
- (8).Blasi P. Poly (lactic acid)/poly (lactic-co-glycolic acid)-based microparticles: an overview. J. Pharm. Invest 2019, 49, 337–346. [Google Scholar]
- (9).Goonoo N. Tunable biomaterials for myocardial tissue regeneration: promising new strategies for advanced biointerface control and improved therapeutic outcomes. Biomater. Sci 2022, 10 (7), 1626–1646. [DOI] [PubMed] [Google Scholar]
- (10).Manavitehrani I; Fathi A; Badr H; Daly S; Negahi Shirazi A; Dehghani F Biomedical Applications of Biodegradable Polyesters. Polymers 2016, 8 (1), 20–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (11).Ye H; Zhang K; Kai D; Li Z; Loh XJ Polyester elastomers for soft tissue engineering. Chem. Soc. Rev 2018, 47 (12), 4545–4580. [DOI] [PubMed] [Google Scholar]
- (12).Davenport Huyer L; Zhang B; Korolj A; Montgomery M; Drecun S; Conant G; Zhao Y; Reis L; Radisic M Highly Elastic and Moldable Polyester Biomaterial for Cardiac Tissue Engineering Applications. ACS Biomater. Sci. Eng 2016, 2 (5), 780–788. [DOI] [PubMed] [Google Scholar]
- (13).Engineer C; Parikh J; Raval A Review on hydrolytic degradation behavior of biodegradable polymers from controlled drug delivery system. Trends in Biomaterials & Artificial Organs 2011, 25 (2), 79–85. [Google Scholar]
- (14).Seyednejad H; Ghassemi AH; van Nostrum CF; Vermonden T; Hennink WE Functional aliphatic polyesters for biomedical and pharmaceutical applications. J. Controlled Release 2011, 152 (1), 168–176. [DOI] [PubMed] [Google Scholar]
- (15).Masoumi N; Jean A; Zugates JT; Johnson KL; Engelmayr GC Laser microfabricated poly (glycerol sebacate) scaffolds for heart valve tissue engineering. J. Biomed. Mater. Res 2012, 2, 104–114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (16).Bot VA; Shakeri A; Weitz JI; Didar TF A Vascular Graft On-a-Chip Platform for Assessing the Thrombogenicity of Vascular Prosthesis and Coatings with Tuneable Flow and Surface Conditions. Adv. Funct. Mater 2022, 32 (41), 2205078–2205078. [Google Scholar]
- (17).Jeffries EM; Allen RA; Gao J; Pesce M; Wang Y Highly elastic and suturable electrospun poly (glycerol sebacate) fibrous scaffolds. Acta Biomater. 2015, 18, 30–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).Montgomery M; Ahadian S; Davenport Huyer L; Lo Rito M; Civitarese RA; Vanderlaan RD; Wu J; Reis LA; Momen A; Akbari S Flexible shape-memory scaffold for minimally invasive delivery of functional tissues. Nat. Mater 2017, 16 (10), 1038–1046. [DOI] [PubMed] [Google Scholar]
- (19).Casanova MR; Reis RL; Martins A; Neves NM Surface biofunctionalization to improve the efficacy of biomaterial substrates to be used in regenerative medicine. Mater. Horiz 2020, 7 (9), 2258–2275. [Google Scholar]
- (20).Yu W; Jiang Y; Lin F; Liu J; Zhou J Surface Biofunctionalization of Tissue Engineered for the Development of Biological Heart Valves: A Review. Coatings 2022, 12 (9), 1322–1322. [Google Scholar]
- (21).Engelmayr GC; Cheng M; Bettinger CJ; Borenstein JT; Langer R; Freed LE Accordion-like honeycombs for tissue engineering of cardiac anisotropy. Nat. Mater 2008, 7 (12), 1003–1010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).Yang J; Webb AR; Ameer GA Novel citric acid-based biodegradable elastomers for tissue engineering. Adv. Mater 2004, 16 (6), 511–516. [Google Scholar]
- (23).Zhang B; Montgomery M; Davenport-Huyer L; Korolj A; Radisic M Platform technology for scalable assembly of instantaneously functional mosaic tissues. Sci. Adv 2015, 1 (7), e1500423. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (24).Lai BFL; Lu RXZ; Davenport Huyer L; Kakinoki S; Yazbeck J; Wang EY; Wu Q; Zhang B; Radisic M A well plate–based multiplexed platform for incorporation of organoids into an organ-on-a-chip system with a perfusable vasculature. Nat. Protoc 2021, 16 (4), 2158–2189. [DOI] [PubMed] [Google Scholar]
- (25).Mohammadi MH; Okhovatian S; Savoji H; Campbell SB; Lai BFL; Wu J; Pascual-Gil S; Bannerman D; Rafatian N; Li RK Toward Hierarchical Assembly of Aligned Cell Sheets into a Conical Cardiac Ventricle Using Microfabricated Elastomers. Adv. Biol 2022, 6 (11), 2101165. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (26).Webb AR; Yang J; Ameer GA Biodegradable polyester elastomers in tissue engineering. Expert Opin. Biol. Ther 2004, 4 (6), 801–812. [DOI] [PubMed] [Google Scholar]
- (27).Gonçalves FA; Fonseca A; Domingos M; Gloria A; Serra A; Coelho JF The potential of unsaturated polyesters in biomedicine and tissue engineering: Synthesis, structure-properties relationships and additive manufacturing. Prog. Polym. Sci 2017, 68, 1–34. [Google Scholar]
- (28).Huyer LD; Montgomery M; Zhao Y; Xiao Y; Conant G; Korolj A; Radisic M Biomaterial based cardiac tissue engineering and its applications. Biomed. Mater 2015, 10 (3), 034004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).Sabir MI; Xu X; Li L A review on biodegradable polymeric materials for bone tissue engineering applications. J. Mater. Sci 2009, 44, 5713–5724. [Google Scholar]
- (30).Yilgor P; Sousa RA; Reis RL; Hasirci N; Hasirci V 3D plotted PCL scaffolds for stem cell based bone tissue engineering. In Macromol. Symp, 2008; Wiley Online Library: Vol. 269, pp 92–99. [Google Scholar]
- (31).Noskovicova N; Hinz B; Pakshir P Implant fibrosis and the underappreciated role of myofibroblasts in the foreign body reaction. Cells 2021, 10 (7), 1794. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (32).Padmanabhan J; Kyriakides TR Nanomaterials, inflammation, and tissue engineering. Wiley Interdiscip. Rev.: Nanomed. Nanobiotechnol 2015, 7 (3), 355–370. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (33).Guimarães CF; Gasperini L; Marques AP; Reis RL The stiffness of living tissues and its implications for tissue engineering. Nat. Rev. Mater 2020, 5 (5), 351–370. [Google Scholar]
- (34).Anderson JM; Shive MS Biodegradation and biocompatibility of PLA and PLGA microspheres. Adv. Drug Delivery Rev 2012, 64, 72–82. [DOI] [PubMed] [Google Scholar]
- (35).Misra S; Srivastava A; Raghuwanshi S; Sharma V; Bisen P Medical grade biodegradable polymers: A perspective from gram-positive bacteria. Fundam. Biomater.: Polym 2018, 1, 267–286. [Google Scholar]
- (36).Zinn M; Witholt B; Egli T Occurrence, synthesis and medical application of bacterial polyhydroxyalkanoate. Adv. Drug Delivery Rev 2001, 53 (1), 5–21. [DOI] [PubMed] [Google Scholar]
- (37).Reddy VUN; Ramanaiah S; Reddy MV; Chang Y-C Review of the developments of bacterial medium-chain-length polyhydroxyalkanoates (mcl-PHAs). Bioengineering 2022, 9 (5), 225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (38).Grande D; Ramier J; Versace DL; Renard E; Langlois V Design of functionalized biodegradable PHA-based electrospun scaffolds meant for tissue engineering applications. N. Biotechnol 2017, 37, 129–137. [DOI] [PubMed] [Google Scholar]
- (39).Dong C-L; Li S-Y; Wang Y; Dong Y; Tang JZ; Chen J-C; Chen G-Q The cytocompatability of polyhydroxyalkanoates coated with a fusion protein of PHA repressor protein (PhaR) and Lys-Gln-Ala-Gly-Asp-Val (KQAGDV) polypeptide. Biomaterials 2012, 33 (9), 2593–2599. [DOI] [PubMed] [Google Scholar]
- (40).Xie H; Li J; Li L; Dong Y; Chen G-Q; Chen KC Enhanced proliferation and differentiation of neural stem cells grown on PHA films coated with recombinant fusion proteins. Acta Biomater. 2013, 9 (8), 7845–7854. [DOI] [PubMed] [Google Scholar]
- (41).Goonoo N; Bhaw-Luximon A; Passanha P; Esteves SR; Jhurry D Third generation poly (hydroxyacid) composite scaffolds for tissue engineering. J. Biomed. Mater. Res., Part B 2017, 105 (6), 1667–1684. [DOI] [PubMed] [Google Scholar]
- (42).Castellano D; Blanes M; Marco B; Cerrada I; Ruiz-Saurí A; Pelacho B; Arana M; Montero JA; Cambra V; Prosper F A comparison of electrospun polymers reveals poly (3-hydroxybutyrate) fiber as a superior scaffold for cardiac repair. Stem Cells Dev. 2014, 23 (13), 1479–1490. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (43).Williams SF; Rizk S; Martin DP Poly-4-hydroxybutyrate (P4HB): a new generation of resorbable medical devices for tissue repair and regeneration. Biomedizinische Technik/Biomedical Engineering 2013, 58 (5), 439–452. [DOI] [PubMed] [Google Scholar]
- (44).Fu N; Meng Z; Jiao T; Luo X; Tang Z; Zhu B; Sui L; Cai X P34HB electrospun fibres promote bone regeneration in vivo. Cell Prolif. 2019, 52 (3), e12601. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (45).Doi Y; Kanesawa Y; Kunioka M; Saito T Biodegradation of microbial copolyesters: poly (3-hydroxybutyrate-co-3-hydroxyvalerate) and poly (3-hydroxybutyrate-co-4-hydroxybutyrate). Macromolecules 1990, 23 (1), 26–31. [Google Scholar]
- (46).Michalak M; Marek AA; Zawadiak J; Kawalec M; Kurcok P Synthesis of PHB-based carrier for drug delivery systems with pH-controlled release. Eur. Polym. J 2013, 49 (12), 4149–4156. [Google Scholar]
- (47).Koseva N; Kurcok P; Adamus G; Troev K; Kowalczuk M Polyester-based Copolymers for Biomaterials Fabrication. In Macromol. Symp, 2007; Wiley Online Library: Vol. 253, pp 24–32. [Google Scholar]
- (48).Ye C; Hu P; Ma M-X; Xiang Y; Liu R-G; Shang X-W PHB/PHBHHx scaffolds and human adipose-derived stem cells for cartilage tissue engineering. Biomaterials 2009, 30 (26), 4401–4406. [DOI] [PubMed] [Google Scholar]
- (49).Kalia VC; Patel SK; Lee J-K Exploiting Polyhydroxyalkanoates for Biomedical Applications. Polymers 2023, 15 (8), 1937. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).Türesin F; Gürsel I; Hasirci V Biodegradable polyhydroxyalkanoate implants for osteomyelitis therapy: in vitro antibiotic release. J. Biomater. Sci. Polym. Ed 2001, 12 (2), 195–207. [DOI] [PubMed] [Google Scholar]
- (51).Sirohi R; Pandey JP; Tarafdar A; Sindhu R; Binod P; Pandey A Applications of poly-3-hydroxybutyrate based composite. Adv Appl Bio Degradable Green Comp 2020, 68, 45. [Google Scholar]
- (52).Sridewi N; Bhubalan K; Sudesh K Degradation of commercially important polyhydroxyalkanoates in tropical mangrove ecosystem. Polym. Degrad. Stab 2006, 91 (12), 2931–2940. [Google Scholar]
- (53).Misra SK; Valappil SP; Roy I; Boccaccini AR Polyhydroxyalkanoate (PHA)/inorganic phase composites for tissue engineering applications. Biomacromolecules 2006, 7 (8), 2249–2258. [DOI] [PubMed] [Google Scholar]
- (54).Zubairi SI; Mantalaris A; Bismarck A; Aizad S Polyhydroxyalkanoates (PHAs) for tissue engineering applications: Biotransformation of palm oil mill effluent (POME) to value-added polymers. Jurnal Teknologi 2016, 78 (1), 13–29. [Google Scholar]
- (55).Ray S; Kalia VC Biomedical Applications of Polyhydroxyalkanoates. Indian J. Microbiol 2017, 57 (3), 261–269. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (56).Serrano M; Pagani R; Vallet-Regı M; Pena J; Ramila A; Izquierdo I; Portolés M In vitro biocompatibility assessment of poly (ε-caprolactone) films using L929 mouse fibroblasts. Biomaterials 2004, 25 (25), 5603–5611. [DOI] [PubMed] [Google Scholar]
- (57).Abedalwafa M; Wang F; Wang L; Li C Biodegradable poly-epsilon-caprolactone (PCL) for tissue engineering applications: A review. Rev. Adv. Mater. Sci 2013, 34 (2), 123–140. [Google Scholar]
- (58).Shahin-Shamsabadi A; Hashemi A; Tahriri M; Bastami F; Salehi M; Abbas FM Mechanical, material, and biological study of a PCL/bioactive glass bone scaffold: Importance of viscoelasticity. Materials Science and Engineering: C 2018, 90, 280–288. [DOI] [PubMed] [Google Scholar]
- (59).Braud C; Devarieux R; Atlan A; Ducos C; Michel V Capillary zone electrophoresis in normal or reverse polarity separation modes for the analysis of hydroxy acid oligomers in neutral phosphate buffer. J. Chromatogr. B: Biomed. Sci. Appl 1998, 706 (1), 73–82. [DOI] [PubMed] [Google Scholar]
- (60).Kim MS; Seo KS; Khang G; Lee HB Ring-Opening Polymerization of ε-Caprolactone by Poly(ethylene glycol) by an Activated Monomer Mechanism. Macromol. Rapid Commun 2005, 26 (8), 643–648. [Google Scholar]
- (61).Labet M; Thielemans W Synthesis of polycaprolactone: a review. Chem. Soc. Rev 2009, 38 (12), 3484–3504. [DOI] [PubMed] [Google Scholar]
- (62).Sant S; Hwang CM; Lee S-H; Khademhosseini A Hybrid PGS–PCL microfibrous scaffolds with improved mechanical and biological properties. J. Tissue Eng. Regen. Med 2011, 5 (4), 283–291. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (63).Tsuji H; Ishizaka T Porous biodegradable polyesters. II. Physical properties, morphology, and enzymatic and alkaline hydrolysis of porous poly (ϵ-caprolactone) films. J. Appl. Polym. Sci 2001, 80 (12), 2281–2291. [Google Scholar]
- (64).Grosberg A; Alford PW; McCain ML; Parker KK Ensembles of engineered cardiac tissues for physiological and pharmacological study: heart on a chip. Lab Chip 2011, 11 (24), 4165–4173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (65).Chun YW; Balikov DA; Feaster TK; Williams CH; Sheng CC; Lee J-B; Boire TC; Neely MD; Bellan LM; Ess KC Combinatorial polymer matrices enhance in vitro maturation of human induced pluripotent stem cell-derived cardiomyocytes. Biomaterials 2015, 67, 52–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (66).Landau S; Szklanny AA; Machour M; Kaplan B; Shandalov Y; Redenski I; Beckerman M; Harari-Steinberg O; Zavin J; Karni-Katovitch O Human-engineered auricular reconstruction (hEAR) by 3D-printed molding with human-derived auricular and costal chondrocytes and adipose-derived mesenchymal stem cells. Biofabrication 2021, 14 (1), 015010. [DOI] [PubMed] [Google Scholar]
- (67).Nalampang K; Molloy R; Punyodom W Synthesis and characterization of poly(L-lactide-co-ε- caprolactone) copolymers: influence of sequential monomer addition on chain microstructure. Polym. Adv. Technol 2007, 18 (3), 240–248, 10.1002/pat.880. [DOI] [Google Scholar]
- (68).Min C; Cui W; Bei J; Wang S Biodegradable shape-memory polymer—polylactide-co-poly (glycolide-co-caprolactone) multiblock copolymer. Polym. Adv. Technol 2005, 16 (8), 608–615. [Google Scholar]
- (69).Tran PA; Hocking DM; O'Connor AJ In situ formation of antimicrobial silver nanoparticles and the impregnation of hydrophobic polycaprolactone matrix for antimicrobial medical device applications. Mater. Sci. Eng 2015, 47, 63–69. [DOI] [PubMed] [Google Scholar]
- (70).Wang Y; Ameer GA; Sheppard BJ; Langer R A tough biodegradable elastomer. Nat. Biotechnol 2002, 20 (6), 602–606. [DOI] [PubMed] [Google Scholar]
- (71).Kemppainen JM; Hollister SJ Tailoring the mechanical properties of 3D-designed poly (glycerol sebacate) scaffolds for cartilage applications. Journal of Biomedical Materials Research Part A: An Official Journal of The Society for Biomaterials, The Japanese Society for Biomaterials, and The Australian Society for Biomaterials and the Korean Society for Biomaterials 2010, 94 (1), 9–18. [DOI] [PubMed] [Google Scholar]
- (72).Singh D; Harding AJ; Albadawi E; Boissonade FM; Haycock JW; Claeyssens F Additive manufactured biodegradable poly (glycerol sebacate methacrylate) nerve guidance conduits. Acta Biomater. 2018, 78, 48–63. [DOI] [PubMed] [Google Scholar]
- (73).Li X; Hong AT-L; Naskar N; Chung H-J Criteria for quick and consistent synthesis of poly (glycerol sebacate) for tailored mechanical properties. Biomacromolecules 2015, 16 (5), 1525–1533. [DOI] [PubMed] [Google Scholar]
- (74).Nijst CL; Bruggeman JP; Karp JM; Ferreira L; Zumbuehl A; Bettinger CJ; Langer R Synthesis and characterization of photocurable elastomers from poly (glycerol-co-sebacate). Biomacromolecules 2007, 8 (10), 3067–3073. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (75).Qu R; Zhou D; Guo T; He W; Cui C; Zhou Y; Zhang Y; Tang Z; Zhang X; Wang Q 4D printing of shape memory inferior vena cava filters based on copolymer of poly (glycerol sebacate) acrylate-co-hydroxyethyl methacrylate (PGSA-HEMA). Mater. Des 2023, 225, 111556. [Google Scholar]
- (76).Rai R; Tallawi M; Barbani N; Frati C; Madeddu D; Cavalli S; Graiani G; Quaini F; Roether JA; Schubert DW Biomimetic poly (glycerol sebacate)(PGS) membranes for cardiac patch application. Mater. Sci. Eng., C 2013, 33 (7), 3677–3687. [DOI] [PubMed] [Google Scholar]
- (77).Tallawi M; Dippold D; Rai R; D'atri D; Roether J; Schubert D; Rosellini E; Engel F; Boccaccini A Novel PGS/PCL electrospun fiber mats with patterned topographical features for cardiac patch applications. Mater. Sci. Eng., C 2016, 69, 569–576. [DOI] [PubMed] [Google Scholar]
- (78).Ravichandran R; Venugopal JR; Sundarrajan S; Mukherjee S; Sridhar R; Ramakrishna S Expression of cardiac proteins in neonatal cardiomyocytes on PGS/fibrinogen core/shell substrate for Cardiac tissue engineering. Int. J. Cardiol 2013, 167 (4), 1461–1468. [DOI] [PubMed] [Google Scholar]
- (79).Kang Y; Yang J; Khan S; Anissian L; Ameer GA A new biodegradable polyester elastomer for cartilage tissue engineering. Journal of Biomedical Materials Research Part A: An Official Journal of The Society for Biomaterials, The Japanese Society for Biomaterials, and The Australian Society for Biomaterials and the Korean Society for Biomaterials 2006, 77 (2), 331–339. [DOI] [PubMed] [Google Scholar]
- (80).Ji Y; Wang X; Liang K Regulating the mechanical properties of poly (1, 8-octanediol citrate) bioelastomer via loading of chitin nanocrystals. RSC advances 2014, 4 (78), 41357–41363. [Google Scholar]
- (81).Wan L; Lu L; Zhu T; Liu Z; Du R; Luo Q; Xu Q; Zhang Q; Jia X Bulk Erosion Degradation Mechanism for Poly (1, 8-octanediol-co-citrate) Elastomer: An In Vivo and In Vitro Investigation. Biomacromolecules 2022, 23 (10), 4268–4281. [DOI] [PubMed] [Google Scholar]
- (82).Motlagh D; Allen J; Hoshi R; Yang J; Lui K; Ameer G Hemocompatibility evaluation of poly (diol citrate) in vitro for vascular tissue engineering. Journal of Biomedical Materials Research Part A: An Official Journal of The Society for Biomaterials, The Japanese Society for Biomaterials, and The Australian Society for Biomaterials and the Korean Society for Biomaterials 2007, 82 (4), 907–916. [DOI] [PubMed] [Google Scholar]
- (83).Sharma AK; Hota PV; Matoka DJ; Fuller NJ; Jandali D; Thaker H; Ameer GA; Cheng EY Urinary bladder smooth muscle regeneration utilizing bone marrow derived mesenchymal stem cell seeded elastomeric poly (1, 8-octanediol-co-citrate) based thin films. Biomaterials 2010, 31 (24), 6207–6217. [DOI] [PubMed] [Google Scholar]
- (84).Prabhakaran MP; Nair AS; Kai D; Ramakrishna S Electrospun composite scaffolds containing poly (octanediol-co-citrate) for cardiac tissue engineering. Biopolymers 2012, 97 (7), 529–538. [DOI] [PubMed] [Google Scholar]
- (85).Vogt L; Boccaccini AR Random and aligned electrospun poly (ε-caprolactone)(PCL)/poly (1, 8-octanediol-co-citrate)(POC) fiber mats for cardiac tissue engineering using benign solvents. Eur. Polym. J 2021, 160, 110772. [Google Scholar]
- (86).Serrano MC; Gutiérrez MC; Jiménez R; Ferrer ML; del Monte F Synthesis of novel lidocaine-releasing poly (diol-co-citrate) elastomers by using deep eutectic solvents. Chem. Commun 2012, 48 (4), 579–581. [DOI] [PubMed] [Google Scholar]
- (87).Huynen MA; Dandekar T; Bork P Variation and evolution of the citric-acid cycle: a genomic perspective. Trends Microbiol. 1999, 7 (7), 281–291. [DOI] [PubMed] [Google Scholar]
- (88).Jiang X; Abgottspon D; Kleeb S; Rabbani S; Scharenberg M; Wittwer M; Haug M; Schwardt O; Ernst B Antiadhesion Therapy for Urinary Tract Infection A Balanced PK/PD Profile Proved To Be Key for Success. J. Med. Chem 2012, 55 (10), 4700–4713. [DOI] [PubMed] [Google Scholar]
- (89).Djordjevic I; Choudhury NR; Dutta NK; Kumar S Synthesis and characterization of novel citric acid-based polyester elastomers. Polymer 2009, 50 (7), 1682–1691. [Google Scholar]
- (90).Djordjevic I; Choudhury NR; Dutta NK; Kumar S Poly [octanediol-co-(citric acid)-co-(sebacic acid)] elastomers: novel bio-elastomers for tissue engineering. Polym. Int 2011, 60 (3), 333–343. [Google Scholar]
- (91).Shah AH; Bhusari S; Djordjevic I; Steele TW Twin screw extrusion of conductive citrate-based biomaterials. Eur. Polym. J 2019, 110, 176–182. [Google Scholar]
- (92).Liu X; Holzwarth JM; Ma PX Functionalized Synthetic Biodegradable Polymer Scaffolds for Tissue Engineering. Macromol. Biosci 2012, 12 (7), 911–919. [DOI] [PubMed] [Google Scholar]
- (93).Sireesha M; Babu VJ; Ramakrishna S Biocompatible and biodegradable elastomer/fibrinogen composite electrospun scaffolds for cardiac tissue regeneration. RSC advances 2015, 5 (125), 103308–103314. [Google Scholar]
- (94).Tran RT; Thevenot P; Gyawali D; Chiao J-C; Tang L; Yang J Synthesis and characterization of a biodegradable elastomer featuring a dual crosslinking mechanism. Soft matter 2010, 6 (11), 2449–2461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (95).Visco A; Brancato V; Campo N Degradation effects in polyester and vinyl ester resins induced by accelerated aging in seawater. J. Compos. Mater 2012, 46 (17), 2025–2040. [Google Scholar]
- (96).Codari F; Lazzari S; Soos M; Storti G; Morbidelli M; Moscatelli D Kinetics of the hydrolytic degradation of poly (lactic acid). Polym. Degrad. Stab 2012, 97 (11), 2460–2466. [Google Scholar]
- (97).Montgomery M; Davenport Huyer L; Bannerman D; Mohammadi MH; Conant G; Radisic M Method for the fabrication of elastomeric polyester scaffolds for tissue engineering and minimally invasive delivery. ACS Biomater. Sci. Eng 2018, 4 (11), 3691–3703. [DOI] [PubMed] [Google Scholar]
- (98).Gyawali D; Tran RT; Guleserian KJ; Tang L; Yang J Citric-acid-derived photo-cross-linked biodegradable elastomers. J. Biomater. Sci. Polym. Ed 2010, 21 (13), 1761–1782. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (99).Zhang B; Montgomery M; Chamberlain MD; Ogawa S; Korolj A; Pahnke A; Wells LA; Massé S; Kim J; Reis L Biodegradable scaffold with built-in vasculature for organ-on-a-chip engineering and direct surgical anastomosis. Nat. Mater 2016, 15 (6), 669–678. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (100).Wang EY; Rafatian N; Zhao Y; Lee A; Lai BFL; Lu RX; Jekic D; Davenport Huyer L; Knee-Walden EJ; Bhattacharya S Biowire model of interstitial and focal cardiac fibrosis. ACS central science 2019, 5 (7), 1146–1158. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (101).Zhao Y; Rafatian N; Wang EY; Feric NT; Lai BF; Knee-Walden EJ; Backx PH; Radisic M Engineering microenvironment for human cardiac tissue assembly in heart-on-a-chip platform. Matrix Biol. 2020, 85, 189–204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (102).Zhao Y; Wang EY; Davenport LH; Liao Y; Yeager K; Vunjak-Novakovic G; Radisic M; Zhang B A multimaterial microphysiological platform enabled by rapid casting of elastic microwires. Adv. Healthcare Mater 2019, 8 (5), 1801187. [DOI] [PubMed] [Google Scholar]
- (103).Cumpata AJ; Peptanariu D; Lungoci A-L; Labusca L; Pinteala M; Radulescu L Towards Regenerative Audiology: Immune Modulation of Adipose-Derived Mesenchymal Cells Preconditioned with Citric Acid-Coated Antioxidant-Functionalized Magnetic Nanoparticles. Medicina 2023, 59 (3), 587. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (104).O’Neill LA; Artyomov MN Itaconate: the poster child of metabolic reprogramming in macrophage function. Nat. Rev. Immunol 2019, 19 (5), 273–281. [DOI] [PubMed] [Google Scholar]
- (105).Lampropoulou V; Sergushichev A; Bambouskova M; Nair S; Vincent EE; Loginicheva E; Cervantes-Barragan L; Ma X; Huang SC-C; Griss T Itaconate links inhibition of succinate dehydrogenase with macrophage metabolic remodeling and regulation of inflammation. Cell Metab. 2016, 24 (1), 158–166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (106).Mills EL; Ryan DG; Prag HA; Dikovskaya D; Menon D; Zaslona Z; Jedrychowski MP; Costa ASH; Higgins M; Hams E; et al. Itaconate is an anti-inflammatory metabolite that activates Nrf2 via alkylation of KEAP1. Nature 2018, 556 (7699), 113–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (107).Regula CG; Yag-Howard C Suture products and techniques: what to use, where, and why. Dermatol. Surg 2015, 41, S187–S200. [DOI] [PubMed] [Google Scholar]
- (108).Törmälä P; Pohjonen T; Rokkanen P Bioabsorbable polymers: materials technology and surgical applications. Proc. Inst. Mech. Eng., Part H 1998, 212 (2), 101–111. [DOI] [PubMed] [Google Scholar]
- (109).Yao D-B; Wu S-D Application of stapling devices in liver surgery: Current status and future prospects. World J. Gastroenterol 2016, 22 (31), 7091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (110).Rhee P; Sharpe R; Huynh T; Burris D; Wang D; Pikoulis E; Sun L; Rich NM; Wherry D Use of titanium vascular staples in trauma. J. Trauma Acute Care Surg 1998, 45 (6), 1097–1099. [DOI] [PubMed] [Google Scholar]
- (111).Abu-Ghanem Y; Meydan C; Segev L; Rubin M; Blumenfeld O; Spivak H Gastric wall thickness and the choice of linear staples in laparoscopic sleeve gastrectomy: challenging conventional concepts. Obes. Surg 2017, 27, 837–843. [DOI] [PubMed] [Google Scholar]
- (112).Yüksel M; Akgül AG; Evman S; Batirel HF Suture and stapler granulomas: a word of caution. Eur. J. Cardiothorac. Surg 2007, 31 (3), 563–565. [DOI] [PubMed] [Google Scholar]
- (113).Ballantyne GH The experimental basis of intestinal suturing: effect of surgical technique, inflammation, and infection on enteric wound healing. Dis. Colon Rectum 1984, 27, 61–71. [DOI] [PubMed] [Google Scholar]
- (114).de la Puerta B; Parsons KJ; Draper ER; Moores AL; Moores AP In vitro comparison of mechanical and degradation properties of equivalent absorbable suture materials from two different manufacturers. Vet. Surg 2011, 40 (2), 223–227. [DOI] [PubMed] [Google Scholar]
- (115).Bao X; Li W; Lu M; Zhou Z Experiment study on puncture force between MIS suture needle and soft tissue. Biosurface and Biotribology 2016, 2 (2), 49–58. [Google Scholar]
- (116).Chakroborty D; Goswami S; Basu S; Sarkar C Catecholamines in the regulation of angiogenesis in cutaneous wound healing. FASEB journal: official publication of the Federation of American Societies for Experimental Biology 2020, 34 (11), 14093. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (117).Waite J; Tanzer M The bioadhesive of Mytilus byssus: a protein containing L-dopa. Biochem. Biophys. Res. Commun 1980, 96 (4), 1554–1561. [DOI] [PubMed] [Google Scholar]
- (118).Chen X; Gao Y; Wang Y; Pan G Mussel-inspired peptide mimicking: An emerging strategy for surface bioengineering of medical implants. Smart Materials in Medicine 2021, 2, 26–37. [Google Scholar]
- (119).Xi Z-Y; Xu Y-Y; Zhu L-P; Wang Y; Zhu B-K A facile method of surface modification for hydrophobic polymer membranes based on the adhesive behavior of poly (DOPA) and poly (dopamine). J. Membr. Sci 2009, 327 (1-2), 244–253. [Google Scholar]
- (120).Chung H; Glass P; Pothen JM; Sitti M; Washburn NR Enhanced adhesion of dopamine methacrylamide elastomers via viscoelasticity tuning. Biomacromolecules 2011, 12 (2), 342–347. [DOI] [PubMed] [Google Scholar]
- (121).Cencer M; Liu Y; Winter A; Murley M; Meng H; Lee BP Effect of pH on the rate of curing and bioadhesive properties of dopamine functionalized poly (ethylene glycol) hydrogels. Biomacromolecules 2014, 15 (8), 2861–2869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (122).Mehdizadeh M; Weng H; Gyawali D; Tang L; Yang J Injectable citrate-based mussel-inspired tissue bioadhesives with high wet strength for sutureless wound closure. Biomaterials 2012, 33 (32), 7972–7983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (123).Guo J; Kim GB; Shan D; Kim JP; Hu J; Wang W; Hamad FG; Qian G; Rizk EB; Yang J Click chemistry improved wet adhesion strength of mussel-inspired citrate-based antimicrobial bioadhesives. Biomaterials 2017, 112, 275–286. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (124).Van der Steen M; Stevens CV Undecylenic acid: a valuable and physiologically active renewable building block from castor oil. ChemSusChem: Chemistry & Sustainability Energy & Materials 2009, 2 (8), 692–713. [DOI] [PubMed] [Google Scholar]
- (125).Li X-C; Jacob MR; Khan SI; Ashfaq MK; Babu KS; Agarwal AK; ElSohly HN; Manly SP; Clark AM Potent in vitro antifungal activities of naturally occurring acetylenic acids. Antimicrob. Agents Chemother 2008, 52 (7), 2442–2448. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (126).Bannerman AD; Davenport Huyer L; Montgomery M; Zhao N; Velikonja C; Bender TP; Radisic M Elastic biomaterial scaffold with spatially varying adhesive design. Adv. Biosyst 2020, 4 (8), 2000046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (127).Philip S; Keshavarz T; Roy I Polyhydroxyalkanoates: biodegradable polymers with a range of applications. J. Chem. Technol. Biotechnol 2007, 82 (3), 233–247. [Google Scholar]
- (128).Wang Y; Sheppard BJ; Langer R Poly(glycerol sebacate)—A Novel Biodegradable Elastomer for Tissue Engineering. MRS Online Proc. Libr 2002, 724, N11.11, From Cambridge University Press Cambridge Core. [Google Scholar]
- (129).Davenport Huyer L; Bannerman AD; Wang Y; Savoji H; Knee-Walden EJ; Brissenden A; Yee B; Shoaib M; Bobicki E; Amsden BG; et al. One-Pot Synthesis of Unsaturated Polyester Bioelastomer with Controllable Material Curing for Microscale Designs. Adv. Healthcare Mater 2019, 8 (16), e1900245. [DOI] [PubMed] [Google Scholar]
- (130).Tokiwa Y; Calabia BP Biodegradability and biodegradation of polyesters. J. Polym. Environ 2007, 15, 259–267. [Google Scholar]
- (131).Assa-Munt N; Jia X; Laakkonen P; Ruoslahti E Solution structures and integrin binding activities of an RGD peptide with two isomers. Biochemistry 2001, 40 (8), 2373–2378. [DOI] [PubMed] [Google Scholar]
- (132).Rahmany MB; Van Dyke M Biomimetic approaches to modulate cellular adhesion in biomaterials: A review. Acta Biomater. 2013, 9 (3), 5431–5437. [DOI] [PubMed] [Google Scholar]
- (133).Lee H-S; Bellin RM; Walker DL; Patel B; Powers P; Liu H; Garcia-Alvarez B; de Pereda JM; Liddington RC; Volkmann N Characterization of an actin-binding site within the talin FERM domain. J. Mol. Biol 2004, 343 (3), 771–784. [DOI] [PubMed] [Google Scholar]
- (134).Fischer MJ Amine coupling through EDC/NHS: a practical approach. Surface plasmon resonance: methods and protocols 2010, 627, 55–73. [DOI] [PubMed] [Google Scholar]
- (135).Amani H; Arzaghi H; Bayandori M; Dezfuli AS; Pazoki-Toroudi H; Shafiee A; Moradi L Controlling Cell Behavior through the Design of Biomaterial Surfaces: A Focus on Surface Modification Techniques. Adv. Mater. Interfaces 2019, 6 (13), 1900572. [Google Scholar]
- (136).Polat E; Güler Z; Balkan T; Sarac AS Covalent streptavidin immobilization on electrospun poly (m-anthranilic acid)/polycaprolactone nanofibers and cytocompatibility. J. Bioact. Compat. Polym 2016, 31 (3), 291–303. [Google Scholar]
- (137).Siddiqui Z; Sarkar B; Kim KK; Kumar A; Paul R; Mahajan A; Grasman JM; Yang J; Kumar VA Self-assembling peptide hydrogels facilitate vascularization in two-component scaffolds. Chem. Eng. J 2021, 422, 130145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (138).Davenport Huyer L; Mandla S; Wang Y; Campbell SB; Yee B; Euler C; Lai BF; Bannerman AD; Lin DS; Montgomery M Macrophage immunomodulation through new polymers that recapitulate functional effects of itaconate as a power house of innate immunity. Adv. Funct. Mater 2021, 31 (6), 2003341. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (139).Ahadian S; Huyer LD; Estili M; Yee B; Smith N; Xu Z; Sun Y; Radisic M Moldable elastomeric polyester-carbon nanotube scaffolds for cardiac tissue engineering. Acta Biomater. 2017, 52, 81–91. [DOI] [PubMed] [Google Scholar]
- (140).Morgan KY; Sklaviadis D; Tochka ZL; Fischer KM; Hearon K; Morgan TD; Langer R; Freed LE Multi-material tissue engineering scaffold with hierarchical pore architecture. Adv. Funct. Mater 2016, 26 (32), 5873–5883. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (141).Park H; Radisic M; Lim JO; Chang BH; Vunjak-Novakovic G A novel composite scaffold for cardiac tissue engineering. In Vitro Cellular & Developmental Biology-Animal 2005, 41, 188–196. [DOI] [PubMed] [Google Scholar]
- (142).Ishizaki T; Saito N; Takai O Correlation of Cell Adhesive Behaviors on Superhydrophobic, Superhydrophilic, and Micropatterned Superhydrophobic/Superhydrophilic Surfaces to Their Surface Chemistry. Langmuir 2010, 26 (11), 8147–8154. [DOI] [PubMed] [Google Scholar]
- (143).Au HTH; Cheng I; Chowdhury MF; Radisic M Interactive effects of surface topography and pulsatile electrical field stimulation on orientation and elongation of fibroblasts and cardiomyocytes. Biomaterials 2007, 28 (29), 4277–4293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (144).Chan BP; Leong KW Scaffolding in tissue engineering: general approaches and tissue-specific considerations. Eur. Spine J 2008, 17 (4), 467–479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (145).Breuls RG; Jiya TU; Smit TH Scaffold stiffness influences cell behavior: opportunities for skeletal tissue engineering. The open orthopaedics journal 2008, 2, 103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (146).Hollister SJ Porous scaffold design for tissue engineering. Nat. Mater 2005, 4 (7), 518–524. [DOI] [PubMed] [Google Scholar]
- (147).Chiu LL; Janic K; Radisic M Engineering of oriented myocardium on three-dimensional micropatterned collagen-chitosan hydrogel. The International journal of artificial organs 2012, 35 (4), 237–250. [DOI] [PubMed] [Google Scholar]
- (148).Al-Haque S; Miklas JW; Feric N; Chiu LL; Chen WLK; Simmons CA; Radisic M Hydrogel substrate stiffness and topography interact to induce contact guidance in cardiac fibroblasts. Macromol. Biosci 2012, 12 (10), 1342–1353. [DOI] [PubMed] [Google Scholar]
- (149).Ghosh K; Pan Z; Guan E; Ge S; Liu Y; Nakamura T; Ren XD; Rafailovich M; Clark RA Cell adaptation to a physiologically relevant ECM mimic with different viscoelastic properties. Biomaterials 2007, 28 (4), 671–679. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (150).Engler AJ; Sen S; Sweeney HL; Discher DE Matrix elasticity directs stem cell lineage specification. Cell 2006, 126 (4), 677–689. [DOI] [PubMed] [Google Scholar]
- (151).Wells RG Tissue mechanics and fibrosis. Biochim. Biophys. Acta 2013, 1832 (7), 884–890. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (152).Najafi M; Farhood B; Mortezaee K Extracellular matrix (ECM) stiffness and degradation as cancer drivers. J. Cell. Biochem 2019, 120 (3), 2782–2790. [DOI] [PubMed] [Google Scholar]
- (153).Bhana B; Iyer RK; Chen WL; Zhao R; Sider KL; Likhitpanichkul M; Simmons CA; Radisic M Influence of substrate stiffness on the phenotype of heart cells. Biotechnol Bioeng 2010, 105 (6), 1148–1160. [DOI] [PubMed] [Google Scholar]
- (154).Engler AJ; Carag-Krieger C; Johnson CP; Raab M; Tang HY; Speicher DW; Sanger JW; Sanger JM; Discher DE Embryonic cardiomyocytes beat best on a matrix with heart-like elasticity: scar-like rigidity inhibits beating. J. Cell Sci 2008, 121 (Pt 22), 3794–3802. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (155).Dhahri W; Sadikov Valdman T; Wilkinson D; Pereira E; Ceylan E; Andharia N; Qiang B; Masoudpour H; Wulkan F; Quesnel E In Vitro Matured Human Pluripotent Stem Cell–Derived Cardiomyocytes Form Grafts With Enhanced Structure and Function in Injured Hearts. Circulation 2022, 145 (18), 1412–1426. [DOI] [PubMed] [Google Scholar]
- (156).Blakney AK; Swartzlander MD; Bryant SJ The effects of substrate stiffness on the in vitro activation of macrophages and in vivo host response to poly(ethylene glycol)-based hydrogels. J. Biomed. Mater. Res. A 2012, 100 (6), 1375–1386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (157).Traister A; Li M; Aafaqi S; Lu M; Arab S; Radisic M; Gross G; Guido F; Sherret J; Verma S Integrin-linked kinase mediates force transduction in cardiomyocytes by modulating SERCA2a/PLN function. Nat. Commun 2014, 5 (1), 4533. [DOI] [PubMed] [Google Scholar]
- (158).Nakayama KH; Hou L; Huang NF Role of extracellular matrix signaling cues in modulating cell fate commitment for cardiovascular tissue engineering. Adv. Healthcare Mater 2014, 3 (5), 628–641. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (159).Thavandiran N; Nunes SS; Xiao Y; Radisic M Topological and electrical control of cardiac differentiation and assembly. Stem Cell. Res. Ther 2013, 4 (1), 1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (160).Weir NA; Buchanan FJ; Orr JF; Farrar DF; Boyd A Processing, annealing and sterilisation of poly-L-lactide. Biomaterials 2004, 25 (18), 3939–3949. [DOI] [PubMed] [Google Scholar]
- (161).Gentile P; Chiono V; Carmagnola I; Hatton PV An overview of poly(lactic-co-glycolic) acid (PLGA)-based biomaterials for bone tissue engineering. Int. J. Mol. Sci 2014, 15 (3), 3640–3659. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (162).Vert M. Lactic acid-based bioresorbable polymers: properties and traps. Med. J. Malaysia 2004, 59 Suppl B, 73–74. [PubMed] [Google Scholar]
- (163).Mikos AG; Bao Y; Cima LG; Ingber DE; Vacanti JP; Langer R Preparation of poly(glycolic acid) bonded fiber structures for cell attachment and transplantation. J Biomed Mater Res 1993, 27 (2), 183–189. [DOI] [PubMed] [Google Scholar]
- (164).Rho JY; Ashman RB; Turner CH Young's modulus of trabecular and cortical bone material: ultrasonic and microtensile measurements. J. Biomech 1993, 26 (2), 111–119. [DOI] [PubMed] [Google Scholar]
- (165).Athanasiou KA; Agarwal A; Dzida FJ Comparative-Study of the Intrinsic Mechanical-Properties of the Human Acetabular and Femoral-Head Cartilage. J. Orthop. Res 1994, 12 (3), 340–349. [DOI] [PubMed] [Google Scholar]
- (166).Tonge TK; Atlan LS; Voo LM; Nguyen TD Full-field bulge test for planar anisotropic tissues: Part I – Experimental methods applied to human skin tissue. Acta Biomater. 2013, 9 (4), 5913–5925. [DOI] [PubMed] [Google Scholar]
- (167).Wuyts FL; Vanhuyse VJ; Langewouters GJ; Decraemer WF; Raman ER; Buyle S Elastic properties of human aortas in relation to age and atherosclerosis: a structural model. Phys. Med. Biol 1995, 40 (10), 1577. [DOI] [PubMed] [Google Scholar]
- (168).Ichihara K; Taguchi T; Sakuramoto I; Kawano S; Kawai S Mechanism of the spinal cord injury and the cervical spondylotic myelopathy: new approach based on the mechanical features of the spinal cord white and gray matter. J. Neurosurg 2003, 99 (3), 278–285. [DOI] [PubMed] [Google Scholar]
- (169).Nagueh SF; Shah G; Wu Y; Torre-Amione G; King NM; Lahmers S; Witt CC; Becker K; Labeit S; Granzier HL Altered titin expression, myocardial stiffness, and left ventricular function in patients with dilated cardiomyopathy. Circulation 2004, 110 (2), 155–162. [DOI] [PubMed] [Google Scholar]
- (170).Weis SM; Emery JL; Becker KD; McBride DJ Jr.; Omens JH; McCulloch AD Myocardial mechanics and collagen structure in the osteogenesis imperfecta murine (oim). Circ. Res 2000, 87 (8), 663–669. [DOI] [PubMed] [Google Scholar]
- (171).Coirault C; Samuel JL; Chemla D; Pourny JC; Lambert F; Marotte F; Lecarpentier Y Increased compliance in diaphragm muscle of the cardiomyopathic Syrian hamster. J. Appl. Physiol 1998, 85 (5), 1762–1769. [DOI] [PubMed] [Google Scholar]
- (172).Omens JH Stress and strain as regulators of myocardial growth. Prog. Biophys. Mol. Biol 1998, 69 (2), 559–572. [DOI] [PubMed] [Google Scholar]
- (173).Wang Y; Ameer GA; Sheppard BJ; Langer R A tough biodegradable elastomer. Nat Biotechnol 2002, 20 (6), 602–606. [DOI] [PubMed] [Google Scholar]
- (174).Langer R; Vacanti JP Tissue engineering. Science 1993, 260 (5110), 920–926. [DOI] [PubMed] [Google Scholar]
- (175).Freier T; Kunze C; Nischan C; Kramer S; Sternberg K; Saß M; Hopt UT; Schmitz K-P In vitro and in vivo degradation studies for development of a biodegradable patch based on poly (3-hydroxybutyrate). Biomaterials 2002, 23 (13), 2649–2657. [DOI] [PubMed] [Google Scholar]
- (176).Cha Y; Pitt C The acceleration of degradation-controlled drug delivery from polyester microspheres. J. Controlled Release 1989, 8 (3), 259–265. [Google Scholar]
- (177).Valverde C; Lligadas G; Ronda JC; Galia M; Cadiz V Hydrolytic and enzymatic degradation studies of aliphatic 10-undecenoic acid-based polyesters. Polym. Degrad. Stab 2018, 155, 84–94. [Google Scholar]
- (178).Göpferich A. Mechanisms of polymer degradation and erosion. The biomaterials: silver jubilee compendium 1996, 117–128. [DOI] [PubMed] [Google Scholar]
- (179).Siepmann J; Elkharraz K; Siepmann F; Klose D How autocatalysis accelerates drug release from PLGA-based microparticles: a quantitative treatment. Biomacromolecules 2005, 6 (4), 2312–2319. [DOI] [PubMed] [Google Scholar]
- (180).Lam CX; Savalani MM; Teoh S-H; Hutmacher DW Dynamics of in vitro polymer degradation of polycaprolactone-based scaffolds: accelerated versus simulated physiological conditions. Biomed. Mater 2008, 3 (3), 034108. [DOI] [PubMed] [Google Scholar]
- (181).Li Y; Chu Z; Li X; Ding X; Guo M; Zhao H; Yao J; Wang L; Cai Q; Fan Y The effect of mechanical loads on the degradation of aliphatic biodegradable polyesters. Regener. Biomater 2017, 4 (3), 179–190. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (182).Badylak SF; Gilbert TW Immune response to biologic scaffold materials. In Semin. Immunol, 2008; Elsevier: Vol. 20, pp 109–116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (183).Yang F; Liu Y-H; Yang X-P; Xu J; Kapke A; Carretero OA Myocardial infarction and cardiac remodelling in mice. Exp. Physiol 2002, 87 (5), 547–555. [DOI] [PubMed] [Google Scholar]
- (184).Minatoguchi S; Takemura G; Chen X-H; Wang N; Uno Y; Koda M; Arai M; Misao Y; Lu C; Suzuki K Acceleration of the healing process and myocardial regeneration may be important as a mechanism of improvement of cardiac function and remodeling by postinfarction granulocyte colony–stimulating factor treatment. Circulation 2004, 109 (21), 2572–2580. [DOI] [PubMed] [Google Scholar]
- (185).Bodner GM Metabolism Part II: The tricarboxylic acid (TCA), citric acid, or Krebs cycle. J. Chem. Educ 1986, 63 (8), 673. [Google Scholar]
- (186).Saborano R; Wongpinyochit T; Totten JD; Johnston BF; Seib FP; Duarte IF Metabolic reprogramming of macrophages exposed to silk, poly (lactic-co-glycolic acid), and silica nanoparticles. Adv. Healthcare Mater 2017, 6 (14), 1601240. [DOI] [PubMed] [Google Scholar]
- (187).Pomerantseva I; Krebs N; Hart A; Neville CM; Huang AY; Sundback CA Degradation behavior of poly (glycerol sebacate). Journal of Biomedical Materials Research Part A: An Official Journal of The Society for Biomaterials, The Japanese Society for Biomaterials, and The Australian Society for Biomaterials and the Korean Society for Biomaterials 2009, 91 (4), 1038–1047. [DOI] [PubMed] [Google Scholar]
- (188).Oh Y-R; Jang Y-A; Song JK; Eom GT Efficient enzymatic depolymerization of polycaprolactone into 6-hydroxyhexanoic acid by optimizing reaction conditions and microbial conversion of 6-hydroxyhexanoic acid into adipic acid for eco-friendly upcycling of polycaprolactone. Biochem. Eng. J 2022, 185, 108504. [Google Scholar]
- (189).Yadav A; Saini V; Arora S MCP-1: chemoattractant with a role beyond immunity: a review. Clin. Chim. Acta 2010, 411 (21-22), 1570–1579. [DOI] [PubMed] [Google Scholar]
- (190).Martinez FO; Gordon S The M1 and M2 paradigm of macrophage activation: time for reassessment. F1000prime reports 2014, 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (191).Gonzalez A. C. d. O.; Costa TF; Andrade Z. d. A.; Medrado ARAP Wound healing-A literature review. An. Bras. Dermatol 2016, 91, 614–620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (192).Spinale FG Myocardial matrix remodeling and the matrix metalloproteinases: influence on cardiac form and function. Physiol. Rev 2007, 87 (4), 1285–1342. [DOI] [PubMed] [Google Scholar]
- (193).Davenport Huyer L; Pascual-Gil S; Wang Y; Mandla S; Yee B; Radisic M Advanced strategies for modulation of the material–macrophage interface. Adv. Funct. Mater 2020, 30 (44), 1909331. [Google Scholar]
- (194).Park A; Wu B; Griffith LG Integration of surface modification and 3D fabrication techniques to prepare patterned poly (L-lactide) substrates allowing regionally selective cell adhesion. J. Biomater. Sci. Polym. Ed 1998, 9 (2), 89–110. [DOI] [PubMed] [Google Scholar]
- (195).Luan HH; Medzhitov R Food fight: role of itaconate and other metabolites in antimicrobial defense. Cell Metab 2016, 24 (3), 379–387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (196).Amici SA; Dong J; Guerau-de-Arellano M Molecular Mechanisms Modulating the Phenotype of Macrophages and Microglia. Front. Immunol 2017, 8, 1520, From NLM. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (197).Cordes T; Wallace M; Michelucci A; Divakaruni AS; Sapcariu SC; Sousa C; Koseki H; Cabrales P; Murphy AN; Hiller K Immunoresponsive gene 1 and itaconate inhibit succinate dehydrogenase to modulate intracellular succinate levels. J. Biol. Chem 2016, 291 (27), 14274–14284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (198).Ma C; Tian X; Kim JP; Xie D; Ao X; Shan D; Lin Q; Hudock MR; Bai X; Yang J Citrate-based materials fuel human stem cells by metabonegenic regulation. PNAS 2018, 115 (50), E11741–E11750. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (199).Karahaliloğlu Z. Cell-compatible PHB/silk fibroin composite nanofibermat for tissue engineering applications. Turk. J. Biol 2017, 41 (3), 503–513. [Google Scholar]
- (200).Sadeghi D; Karbasi S; Razavi S; Mohammadi S; Shokrgozar MA; Bonakdar S Electrospun poly (hydroxybutyrate)/chitosan blend fibrous scaffolds for cartilage tissue engineering. J. Appl. Polym. Sci 2016, 133 (47). [Google Scholar]
- (201).Tai H-Y; Fu E; Cheng L-P; Don T-M Fabrication of asymmetric membranes from polyhydroxybutyrate and biphasic calcium phosphate/chitosan for guided bone regeneration. J. Polym. Res 2014, 21, 1–11. [Google Scholar]
- (202).Mohammadalizadeh Z; Karbasi S; Arasteh S Physical, mechanical and biological evaluation of poly (3-hydroxybutyrate)-chitosan/MWNTs as a novel electrospun scaffold for cartilage tissue engineering applications. Polym.-Plast. Technol. Mater 2020, 59 (4), 417–429. [Google Scholar]
- (203).Karimi A; Karbasi S; Razavi S; Zargar EN Poly (hydroxybutyrate)/chitosan aligned electrospun scaffold as a novel substrate for nerve tissue engineering. Adv. Biomed. Res 2018, 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (204).Iron R; Mehdikhani M; Naghashzargar E; Karbasi S; Semnani D Effects of nano-bioactive glass on structural, mechanical and bioactivity properties of Poly (3-hydroxybutyrate) electrospun scaffold for bone tissue engineering applications. Materials Technology 2019, 34 (9), 540–548. [Google Scholar]
- (205).Zhuikov VA; Akoulina EA; Chesnokova DV; Wenhao Y; Makhina TK; Demyanova IV; Zhuikova YV; Voinova VV; Belishev NV; Surmenev RA The growth of 3T3 fibroblasts on PHB, PLA and PHB/PLA blend films at different stages of their biodegradation in vitro. Polymers 2020, 13 (1), 108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (206).Van de Velde K; Kiekens P Biopolymers: overview of several properties and consequences on their applications. Polym. Test 2002, 21 (4), 433–442. [Google Scholar]
- (207).Ma H; Sun Y; Tang Y; Shen Y; Kan Z; Li Q; Fang S; Lu Y; Zhou X; Li Z Robust Electrospun Nanofibers from Chemosynthetic Poly (4-hydroxybutyrate) as Artificial Dural Substitute. Macromol. Biosci 2021, 21 (7), 2100134. [DOI] [PubMed] [Google Scholar]
- (208).Li Y; Yao S; Han C; Yu Y; Xiao L Ternary blends from biological poly(3-hydroxybutyrate-co-4-hydroxyvalerate), poly(L-lactic acid), and poly(vinyl acetate) with balanced properties. Int. J. Biol. Macromol 2021, 181, 60–71. [DOI] [PubMed] [Google Scholar]
- (209).Fu N; Xie J; Li G; Shao X; Shi S; Lin S; Deng S; Sun K; Lin Y P34HB film promotes cell adhesion, in vitro proliferation, and in vivo cartilage repair. Rsc Advances 2015, 5 (28), 21572–21579. [Google Scholar]
- (210).Lei C; Zhu H; Li J; Li J; Feng X; Chen J Preparation and characterization of polyhydroxybutyrate-co-hydroxyvalerate/silk fibroin nanofibrous scaffolds for skin tissue engineering. Polym. Eng. Sci 2015, 55 (4), 907–916. [Google Scholar]
- (211).Prabhakaran MP; Vatankhah E; Ramakrishna S Electrospun aligned PHBV/collagen nanofibers as substrates for nerve tissue engineering. Biotechnol. Bioeng 2013, 110 (10), 2775–2784. [DOI] [PubMed] [Google Scholar]
- (212).Sadat-Shojai M; Khorasani M-T; Jamshidi A A new strategy for fabrication of bone scaffolds using electrospun nano-HAp/PHB fibers and protein hydrogels. Chem. Eng. J 2016, 289, 38–47. [Google Scholar]
- (213).Sultana N; Khan TH In vitro degradation of PHBV scaffolds and nHA/PHBV composite scaffolds containing hydroxyapatite nanoparticles for bone tissue engineering. J. Nanomater 2012, 2012, 1–12. [Google Scholar]
- (214).Matzinos P; Tserki V; Kontoyiannis A; Panayiotou C Processing and characterization of starch/polycaprolactone products. Polym. Degrad. Stab 2002, 77 (1), 17–24. [Google Scholar]
- (215).Domingos M; Chiellini F; Cometa S; De Giglio E; Grillo-Fernandes E; Bártolo P; Chiellini E Evaluation of in vitro degradation of PCL scaffolds fabricated via BioExtrusion. Part 1: Influence of the degradation environment. Virtual and Physical Prototyping 2010, 5 (2), 65–73. [Google Scholar]
- (216).Lee J; Tae G; Kim YH; Park IS; Kim S-H; Kim SH The effect of gelatin incorporation into electrospun poly(l-lactide-co-ε-caprolactone) fibers on mechanical properties and cytocompatibility. Biomaterials 2008, 29 (12), 1872–1879. [DOI] [PubMed] [Google Scholar]
- (217).Haim Zada M; Kumar A; Elmalak O; Markovitz E; Icekson R; Domb AJ In vitro and in vivo degradation behavior and the long-term performance of biodegradable PLCL balloon implants. Int. J. Pharm 2020, 574, 118870. [DOI] [PubMed] [Google Scholar]
- (218).Jee MH; Park JH; Choi SY; Baik DH Effects of Introduction of L-Lactide on Microstructures, Thermal Properties and In vitro Degradation of Poly(glycolide-co-ε-caprolactone) Block Copolymer. Fibers Polym.2022, 23 (10), 2683–2691. [Google Scholar]
- (219).Low YJ; Kittur MI; Andriyana A; Ang BC; Zainal Abidin NI A novel approach to evaluate the mechanical responses of elastin-like bioresorbable poly(glycolide-co-caprolactone) (PGCL) suture. J. Mech. Behav. Biomed. Mater 2023, 140, 105723. [DOI] [PubMed] [Google Scholar]
- (220).Wang Y; Sheppard BJ; Langer R Poly (glycerol sebacate)—a novel biodegradable elastomer for tissue engineering. MRS Online Proc. Libr 2002, 724. [Google Scholar]
- (221).Rai R; Tallawi M; Grigore A; Boccaccini AR Synthesis, properties and biomedical applications of poly (glycerol sebacate)(PGS): A review. Prog. Polym. Sci 2012, 37 (8), 1051–1078. [Google Scholar]
- (222).Shou Y; Campbell SB; Lam A; Lausch AJ; Santerre JP; Radisic M; Davenport Huyer L Toward renewable and functional biomedical polymers with tunable degradation rates based on itaconic acid and 1, 8-octanediol. ACS Appl. Polym. Mater 2021, 3 (4), 1943–1955. [Google Scholar]
- (223).Ghista DN; Vayo WH; Sandler H Elastic modulus of the human intact left ventricle—determination and physiological interpretation. Med. Biol. Eng 1975, 13, 151–161. [DOI] [PubMed] [Google Scholar]
- (224).Lee JM; Boughner DR Mechanical properties of human pericardium. Differences in viscoelastic response when compared with canine pericardium. Circulation Res. 1985, 57 (3), 475–481. [DOI] [PubMed] [Google Scholar]
- (225).Girton T; Oegema T; Grassl E; Isenberg B; Tranquillo R Mechanisms of stiffening and strengthening in media-equivalents fabricated using glycation. J. Biomech. Eng 2000, 122 (3), 216–223. [DOI] [PubMed] [Google Scholar]
- (226).Matsuda M; Nosaka T; Sato M; Iijima J; Ohshima N; Fukushima H Effects of exercise training on biochemical and biomechanical properties of rat aorta. Angiology 1989, 40 (1), 51–58. [DOI] [PubMed] [Google Scholar]
- (227).Lai BFL; Huyer LD; Lu RXZ; Drecun S; Radisic M; Zhang B InVADE: integrated vasculature for assessing dynamic events. Adv. Funct. Mater 2017, 27 (46), 1703524. [Google Scholar]
- (228).Boutry CM; Kaizawa Y; Schroeder BC; Chortos A; Legrand A; Wang Z; Chang J; Fox P; Bao Z A stretchable and biodegradable strain and pressure sensor for orthopaedic application. Nat. Electron 2018, 1 (5), 314–321. [Google Scholar]
- (229).Sundback CA; Shyu JY; Wang Y; Faquin WC; Langer RS; Vacanti JP; Hadlock TA Biocompatibility analysis of poly (glycerol sebacate) as a nerve guide material. Biomaterials 2005, 26 (27), 5454–5464. [DOI] [PubMed] [Google Scholar]
- (230).Kolewe ME; Park H; Gray C; Ye X; Langer R; Freed LE 3D structural patterns in scalable, elastomeric scaffolds guide engineered tissue architecture. Adv. Mater 2013, 25 (32), 4459–4465. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (231).Pollet AMAO; Toonder J. M. J. d. Recapitulating the Vasculature Using Organ-On-Chip Technology. Bioengineering 2020, 7, 17–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (232).Liu C; Campbell SB; Li J; Bannerman D; Pascual-Gil S; Kieda J; Wu Q; Herman PR; Radisic M High Throughput Omnidirectional Printing of Tubular Microstructures from Elastomeric Polymers. Adv. Healthcare Mater 2022, 11 (23), 2201346. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (233).Moradkhannejhad L; Abdouss M; Nikfarjam N; Shahriari MH; Heidary V The effect of molecular weight and content of PEG on in vitro drug release of electrospun curcumin loaded PLA/PEG nanofibers. J. Drug Delivery Sci. Technol 2020, 56, 101554–101554. [Google Scholar]
- (234).Fan T; Daniels R Preparation and Characterization of Electrospun Polylactic Acid (PLA) Fiber Loaded with Birch Bark Triterpene Extract for Wound Dressing. AAPS PharmSciTech 2021, 22 (6), 205–205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (235).Wu J.-h.; Hu T.-g.; Wang H; Zong M.-h.; Wu H; Wen P Electrospinning of PLA Nanofibers: Recent Advances and Its Potential Application for Food Packaging. Journal of Agricultural and Food Chemistrygricultural and food chemistry 2022, 70, 8207–8221-8207–8221. [DOI] [PubMed] [Google Scholar]
- (236).Krynauw H; Omar R; Koehne J; Limbert G; Davies NH; Bezuidenhout D; Franz T Electrospun polyester-urethane scaffold preserves mechanical properties and exhibits strain stiffening during in situ tissue ingrowth and degradation. SN Appl. Sci 2020, 2 (5), 953–953. [Google Scholar]
- (237).Hu J.-j.; Liu C.-c.; Lin C.-h.; Tuan-Mu H-Y Synthesis, Characterization, and Electrospinning of a Functionalizable, Polycaprolactone-Based Polyurethane for Soft Tissue Engineering. Polymers 2021, 13, 1527–1527. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (238).Yi F; Lavan DA Poly (glycerol sebacate) Nanofiber Scaffolds by Core/Shell Electrospinning. Macromol. Biosci 2008, 8, 803–806. [DOI] [PubMed] [Google Scholar]
- (239).Masoumi N; Larson BL; Annabi N; Kharaziha M; Zamanian B; Shapero KS; Cubberley AT; Camci-unal G; Manning KB; Jr JEM; et al. Electrospun PGS:PCL Microfibers Align Human Valvular Interstitial Cells and Provide Tunable Scaffold Anisotropy. Adv. Healthcare Mater 2014, 3, 929–939. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (240).Salehi S; Bahners T; Gutmann JS; Fuchsluger TA Characterization of structural, mechanical and nano-mechanical properties of electrospun PGS/PCL fiber. RSC Advances 2014, 4, 16951–16957. [Google Scholar]
- (241).Hu J; Kai D; Ye H; Tian L; Ding X; Ramakrishna S; Jun X Electrospinning of poly (glycerol sebacate)-based nano fibers for nerve tissue engineering. Mater. Sci. Eng., C 2017, 70, 1089–1094. [DOI] [PubMed] [Google Scholar]
- (242).Xu B; Li Y; Zhu C; Cook WD; Forsythe J; Chen Q Fabrication, mechanical properties and cytocompatibility of elastomeric nanofibrous mats of poly (glycerol sebacate). Eur. Polym. J 2015, 64, 79–92. [Google Scholar]
- (243).Ravi-Kumar S; Lies B; Zhang X; Lyu H; Qin H Laser ablation of polymers: A review. Polym. Int 2019, 68 (8), 1391–1401. [Google Scholar]
- (244).Osuchowska PN; Ostrowski R; Sarzyński A; Strzelec M; Mierczyk Z; Trafny EA Results in Physics Microstructured polyethylene terephthalate (PET) for microsieving of cancer cells. Results Phys. 2019, 15 (August), 102612–102612. [Google Scholar]
- (245).Mezera M; Alamri S; Hendriks WAPM; Hertwig A; Elert AM; Bonse J; Kunze T; Lasagni AF; Römer G.-w. R. B. E. Hierarchical Micro-/Nano-Structures on Polycarbonate via UV Pulsed Laser Processing. 2020; Vol. 10, pp 1184–1184. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (246).Shakeri A; Khan S; Jarad NA; Didar TF The Fabrication and Bonding of Thermoplastic Microfluidics: A Review. Materials 2022, 15, 6478–6478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (247).Park H; Larson BL; Guillemette MD; Jain SR; Hua C; Engelmayr GC Jr; Freed LE The significance of pore microarchitecture in a multi-layered elastomeric scaffold for contractile cardiac muscle constructs. Biomaterials 2011, 32 (7), 1856–1864. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (248).Zhang B; Lai BFL; Xie R; Davenport Huyer L; Montgomery M; Radisic M; Huyer LD; Montgomery M; Radisic M; Davenport Huyer L; et al. Microfabrication of AngioChip, a biodegradable polymer scaffold with microfluidic vasculature. Nat. Protoc 2018, 13 (8), 1793–1813. [DOI] [PubMed] [Google Scholar]
- (249).Lu RXZ; Lai BFL; Benge T; Wang EY; Davenport Huyer L; Rafatian N; Radisic M; Huyer LD; Rafatian N; Radisic M Heart-on-a-Chip Platform for Assessing Toxicity of Air Pollution Related Nanoparticles. Adv. Mater. Technol 2021, 6 (2), 1–15. [Google Scholar]
- (250).Lu RXZ; Lai BFL; Rafatian N; Gustafson D; Campbell SB; Banerjee A; Kozak R; Mossman K; Mubareka S; Howe KL; et al. Vasculature-on-a-chip platform with innate immunity enables identification of angiopoietin-1 derived peptide as a therapeutic for SARS-CoV-2 induced inflammation. Lab Chip 2022, 22 (6), 1171–1186. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (251).Lai BFL; Davenport Huyer L; Lu RXZ; Drecun S; Radisic M; Zhang B; Huyer LD; Lu RXZ; Drecun S; Radisic M; et al. InVADE: Integrated Vasculature for Assessing Dynamic Events. Adv. Funct. Mater 2017, 27 (46), 1703524–1703524. [Google Scholar]
- (252).Lai BFL; Lu RXZ; Hu Y; Davenport Huyer L; Dou W; Wang EY; Radulovich N; Tsao MS; Sun Y; Radisic M; et al. Recapitulating Pancreatic Tumor Microenvironment through Synergistic Use of Patient Organoids and Organ-on-a-Chip Vasculature. Adv. Funct. Mater 2020, 30 (48), 2000545–2000545. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (253).Hinton TJ; Jallerat Q; Palchesko RN; Park JH; Grodzicki MS; Shue H-J; Ramadan MH; Hudson AR; Feinberg AW Three-dimensional printing of complex biological structures by freeform reversible embedding of suspended hydrogels. Sci. Adv 2015, 1 (9), e1500758–e1500758. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (254).Savoji H; Davenport Huyer L; Mohammadi MH; Lun Lai BF; Rafatian N; Bannerman D; Shoaib M; Bobicki ER; Ramachandran A; Radisic M 3D Printing of Vascular Tubes Using Bioelastomer Prepolymers by Freeform Reversible Embedding. ACS Biomater. Sci. Eng 2020, 6 (3), 1333–1343. [DOI] [PubMed] [Google Scholar]
- (255).Melchiorri AJ; Hibino N; Best CA; Yi T; Lee YU; Kraynak CA; Kimerer LK; Krieger A; Kim P; Breuer CK; et al. 3D-Printed Biodegradable Polymeric Vascular Grafts. Adv. Healthcare Mater 2016, 5, 319–325. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (256).Elomaa L; Teixeira S; Hakala R; Korhonen H; Grijpma DW; Seppälä JV Preparation of poly(ε-caprolactone)-based tissue engineering scaffolds by stereolithography. Acta Biomater. 2011, 7 (11), 3850–3856. [DOI] [PubMed] [Google Scholar]
- (257).Elomaa L; Kokkari A; Närhi T; Seppälä JV Porous 3D modeled scaffolds of bioactive glass and photocrosslinkable poly(ε-caprolactone) by stereolithography. Compos. Sci. Technol 2013, 74, 99–106. [Google Scholar]
- (258).Wales DJ; Keshavarz M; Howe C; Yeatman E 3D Printability Assessment of Poly (octamethylene maleate (anhydride) citrate) and Poly (ethylene glycol) Diacrylate Copolymers for Biomedical Applications. ACS Appl. Polym. Mater 2022, 4 (8), 5457–5470. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (259).Iyer RK; Chiu LL; Reis LA; Radisic M Engineered cardiac tissues. Curr. Opin. Biotechnol 2011, 22, 706–714. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (260).Shakeri A; Rahmani S; Imani SM; Osborne M; Yousefi H; Didar TF Biofunctional interfaces for cell culture in microfluidic devices. Pal K, Kraatz H-B, Khasnobish A, Bag S, Banerjee I, Kuruganti U Eds.; Woodhead Publishing, 2019; pp 635–699. [Google Scholar]
- (261).Shakeri A; Jarad NA; Leung A; Soleymani L; Didar TF Biofunctionalization of Glass- and Paper-Based Microfluidic Devices: A Review. Adv. Mater. Interfaces 2019, 6 (19), 1900940. [Google Scholar]
- (262).Shakeri A; Khan S; Didar TF Conventional and emerging strategies for the fabrication and functionalization of PDMS-based microfluidic devices. Lab Chip 2021, 21 (16), 3053–3075. [DOI] [PubMed] [Google Scholar]
- (263).Shakeri A; Jarad NA; Khan S; F Didar T Bio-functionalization of microfluidic platforms made of thermoplastic materials: A review. Anal. Chim. Acta 2022, 1209, 339283–339283. [DOI] [PubMed] [Google Scholar]
- (264).Zhao Y; Rafatian N; Feric NT; Cox BJ; Aschar-Sobbi R; Wang EY; Aggarwal P; Zhang B; Conant G; Ronaldson-Bouchard K A platform for generation of chamber-specific cardiac tissues and disease modeling. Cell 2019, 176 (4), 913–927. e918. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (265).Wang EY; Zhao Y; Okhovatian S; Smith JB; Radisic M Intersection of stem cell biology and engineering towards next generation in vitro models of human fibrosis. Frontiers in Bioengineering and Biotechnology 2022, 10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (266).Lai BFL; Lu RXZ; Hu Y; Davenport Huyer L; Dou W; Wang EY; Radulovich N; Tsao MS; Sun Y; Radisic M Recapitulating Pancreatic Tumor Microenvironment through Synergistic Use of Patient Organoids and Organ-on-a-Chip Vasculature. Advanced functional materials 2020, 30 (48), 2000545. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (267).Mandla S; Davenport Huyer L; Wang Y; Radisic M Macrophage polarization with angiopoietin-1 peptide QHREDGS. ACS Biomater. Sci. Eng 2019, 5 (9), 4542–4550. [DOI] [PubMed] [Google Scholar]
- (268).Ruther F; Roether JA; Boccaccini AR 3D Printing of Mechanically Resistant Poly (Glycerol Sebacate) (PGS)-Zein Scaffolds for Potential Cardiac Tissue Engineering Applications. Adv. Eng. Mater 2022, 24 (9), 2101768–2101768. [Google Scholar]
- (269).Hermanová S; Pumera M Biodegradable polyester platform for extrusion-based bioprinting. Bioprinting 2022, 26, e00198–e00198. [Google Scholar]
- (270).Li M; Mondrinos MJ; Chen X; Gandhi MR; Ko FK; Lelkes PI Elastin Blends for Tissue Engineering Scaffolds. J. Biomed. Mater. Res., Part A 2006, 79 (4), 963–973. [DOI] [PubMed] [Google Scholar]
- (271).Chai F; Maton M; Degoutin S; Vermet G; Simon N; Rousseaux C; Martel B; Blanchemain N In vivo evaluation of post-operative pain reduction on rat model after implantation of intraperitoneal PET meshes functionalised with cyclodextrins and loaded with ropivacaine. Biomaterials 2019, 192, 260–270. [DOI] [PubMed] [Google Scholar]
- (272).Ivanoska-Dacikj A; Bogoeva-Gaceva G; Krumme A; Tarasova E; Scalera C; Stojkovski V; Gjorgoski I; Ristoski T Biodegradable polyurethane/graphene oxide scaffolds for soft tissue engineering: in vivo behavior assessment. Int. J. Polym. Mater. Polym. Biomater 2020, 69 (17), 1101–1111. [Google Scholar]
- (273).Budak K; Sogut O; Aydemir Sezer U A review on synthesis and biomedical applications of polyglycolic acid. J. Polym. Res 2020, 27 (8), 208–208. [Google Scholar]
- (274).Xia Q; Ye W; Zhang Q; Liu D; Gao S; Fan Z; Liu Q Structure, properties, and in vitro degradation behavior of biodegradable poly(L-lactic acid)-trimethylene carbonate-glycolide terpolymer. J. Appl. Polym. Sci 2022, 139 (40), e52968–e52968. [Google Scholar]
- (275).Weis SM; Emery JL; Becker KD; McBride DJ Jr; Omens JH; McCulloch AD Myocardial mechanics and collagen structure in the osteogenesis imperfecta murine (oim). Circ. Res 2000, 87 (8), 663–669. [DOI] [PubMed] [Google Scholar]
- (276).Nagueh SF; Shah G; Wu Y; Torre-Amione G; King NM; Lahmers S; Witt CC; Becker K; Labeit S; Granzier HL Altered titin expression, myocardial stiffness, and left ventricular function in patients with dilated cardiomyopathy. Circulation 2004, 110 (2), 155–162. [DOI] [PubMed] [Google Scholar]
- (277).Bannerman D; Pascual-Gil S; Radisic M An optimal gel patch for the injured heart. Nat. Biomed. Eng 2019, 3 (8), 592–593. [DOI] [PubMed] [Google Scholar]
- (278).Dvir T; Kedem A; Ruvinov E; Levy O; Freeman I; Landa N; Holbova R; Feinberg MS; Dror S; Etzion Y Prevascularization of cardiac patch on the omentum improves its therapeutic outcome. PNAS 2009, 106 (35), 14990–14995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (279).Song H; Yoon C; Kattman SJ; Dengler J; Massé S; Thavaratnam T; Gewarges M; Nanthakumar K; Rubart M; Keller GM Interrogating functional integration between injected pluripotent stem cell-derived cells and surrogate cardiac tissue. PNAS 2010, 107 (8), 3329–3334. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (280).Reis LA; Chiu LL; Liang Y; Hyunh K; Momen A; Radisic M A peptide-modified chitosan-collagen hydrogel for cardiac cell culture and delivery. Acta Biomater 2011. [DOI] [PubMed] [Google Scholar]
- (281).Ott HC; Matthiesen TS; Goh SK; Black LD; Kren SM; Netoff TI; Taylor DA Perfusion-decellularized matrix: using nature's platform to engineer a bioartificial heart. Nat. Med 2008, 14 (2), 213–221. [DOI] [PubMed] [Google Scholar]
- (282).Taylor DA From stem cells and cadaveric matrix to engineered organs. Curr. Opin. Biotechnol 2009, 20 (5), 598–605. [DOI] [PubMed] [Google Scholar]
- (283).DeQuach JA; Mezzano V; Miglani A; Lange S; Keller GM; Sheikh F; Christman KL Simple and high yielding method for preparing tissue specific extracellular matrix coatings for cell culture. PLoS One 2010, 5 (9), e13039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (284).Zimmermann W-H; Schneiderbanger K; Schubert P; Didie M; Munzel F; Heubach J; Kostin S; Neuhuber W; Eschenhagen T Tissue engineering of a differentiated cardiac muscle construct. Circulation Res. 2002, 90 (2), 223–230. [DOI] [PubMed] [Google Scholar]
- (285).Leor J; Tuvia S; Guetta V; Manczur F; Castel D; Willenz U; Petnehazy O; Landa N; Feinberg MS; Konen E; et al. Intracoronary injection of in situ forming alginate hydrogel reverses left ventricular remodeling after myocardial infarction in Swine. J. Am. Coll. Cardiol 2009, 54 (11), 1014–1023. [DOI] [PubMed] [Google Scholar]
- (286).Zimmermann W-H; Melnychenko I; Wasmeier G; Didié M; Naito H; Nixdorff U; Hess A; Budinsky L; Brune K; Michaelis B Engineered heart tissue grafts improve systolic and diastolic function in infarcted rat hearts. Nat. Med 2006, 12 (4), 452–458. [DOI] [PubMed] [Google Scholar]
- (287).Montgomery M; Ahadian S; Davenport Huyer L; Lo Rito M; Civitarese RA; Vanderlaan RD; Wu J; Reis LA; Momen A; Akbari S; et al. Flexible shape-memory scaffold for minimally invasive delivery of functional tissues. Nat. Mater 2017, 16 (10), 1038–1046. [DOI] [PubMed] [Google Scholar]
- (288).Nuckols TK; Keeler E; Morton SC; Anderson L; Doyle B; Booth M; Shanman R; Grein J; Shekelle P Economic Evaluation of Quality Improvement Interventions for Bloodstream Infections Related to Central Catheters: A Systematic Review. JAMA Intern Med 2016, 176 (12), 1843–1854. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (289).Peace CG; O’Neill LA The role of itaconate in host defense and inflammation. The Journal of Clinical Investigation 2022, 132 (2). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (290).Kuzmanov U; Wang EY; Vanderlaan R; Kim DH; Lee S-H; Hadipour-Lakmehsari S; Guo H; Zhao Y; McFadden M; Sharma P Mapping signalling perturbations in myocardial fibrosis via the integrative phosphoproteomic profiling of tissue from diverse sources. Nat. Biomed. Eng 2020, 4 (9), 889–900. [DOI] [PubMed] [Google Scholar]
- (291).Zhu C; Zhang Z; Liu Q; Wang Z; Jin J Synthesis and biodegradation of aliphatic polyesters from dicarboxylic acids and diols. Journal of applied polymer science 2003, 90 (4), 982–990. [Google Scholar]
- (292).Reed GW; Rossi JE; Cannon CP Acute myocardial infarction. The Lancet 2017, 389 (10065), 197–210. [DOI] [PubMed] [Google Scholar]
- (293).Ahadian S; Civitarese R; Bannerman D; Mohammadi MH; Lu R; Wang E; Davenport-Huyer L; Lai B; Zhang B; Zhao Y Organ-on-a-chip platforms: a convergence of advanced materials, cells, and microscale technologies. Adv. Healthcare Mater 2018, 7 (2), 1700506. [DOI] [PubMed] [Google Scholar]
