Next Article in Journal
Biomineralization in Three-Dimensional Scaffolds Based on Bacterial Nanocellulose for Bone Tissue Engineering: Feature Characterization and Stem Cell Differentiation
Next Article in Special Issue
Properties of Resorbable Conduits Based on Poly(L-Lactide) Nanofibers and Chitosan Fibers for Peripheral Nerve Regeneration
Previous Article in Journal
Three-Dimensional Stress Fields in Thick Orthotropic Plates with Sharply Curved Notches under In-Plane and Out-of-Plane Shear
Previous Article in Special Issue
Synthetic Heparan Sulfate Mimetic Polymer Enhances Corneal Nerve Regeneration and Wound Healing after Experimental Laser Ablation Injury in Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Current Strategies for Engineered Vascular Grafts and Vascularized Tissue Engineering

1
Department of Organ Transplantation, Zhujiang Hospital, Southern Medical University, Guangzhou 510280, China
2
Center for Chemical Biology and Drug Discovery, Laboratory of Computational Biomedicine, Guangzhou Institute of Biomedicine and Health, Chinese Academy of Sciences, Guangzhou 510530, China
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(9), 2015; https://doi.org/10.3390/polym15092015
Submission received: 16 December 2022 / Revised: 21 April 2023 / Accepted: 21 April 2023 / Published: 24 April 2023
(This article belongs to the Special Issue Biopolymer-Based Biomimetic Scaffolds)

Abstract

:
Blood vessels not only transport oxygen and nutrients to each organ, but also play an important role in the regulation of tissue regeneration. Impaired or occluded vessels can result in ischemia, tissue necrosis, or even life-threatening events. Bioengineered vascular grafts have become a promising alternative treatment for damaged or occlusive vessels. Large-scale tubular grafts, which can match arteries, arterioles, and venules, as well as meso- and microscale vasculature to alleviate ischemia or prevascularized engineered tissues, have been developed. In this review, materials and techniques for engineering tubular scaffolds and vasculature at all levels are discussed. Examples of vascularized tissue engineering in bone, peripheral nerves, and the heart are also provided. Finally, the current challenges are discussed and the perspectives on future developments in biofunctional engineered vessels are delineated.

1. Introduction

Biological processes, such as tissue regeneration, stem cell maintenance, and cancer progression, can be regulated by vasculature and its surrounding microenvironment through distinct biophysical and biochemical cues, as well as the interference of cell behavior and cell–cell communications [1,2,3,4,5]. Vasculature exhibit tissue-specific phenotypes, varying from molecular signatures to architectures and functionalities in different organs [6]. Vessel networks not only play an important role in the transportation of oxygen and nutrients, but also regulate the deposition and remodeling of the peripheral ECM through signaling between vascular and nonvascular cells [7,8].
In general, three distinct processes are responsible for vascularization: vasculogenesis, angiogenesis, and arteriogenesis [9,10]. Vasculogenesis is a process including neovessel formation associated with embryogenesis, while angiogenesis describes the formation of vasculature from existing and mature vessels; finally, arteriogenesis indicates the consolidation and remodeling of existing collateral vessels [9]. Cell signature, biochemical, and biophysical cues that regulate vascularization in these three mechanisms have been studied to guide the in vitro formation of vascular grafts of various diameters, as well as interconnected vasculature. However, the complexity of the vascularization process and technical challenges currently limit most of these efforts in the preclinical stage.
Blood vessels can be categorized into three types: arteries, veins, and capillaries (Figure 1). Each type of blood vessel has its own unique functionality and structure. Arteries and veins are responsible for the transportation of oxygenated or deoxygenated blood from or to the heart, respectively. They are composed of a tunica internal, which supports the vascular endothelium surrounded by one or more layered smooth muscle cells (SMCs), tunica media, and collagen fibril tunica externa. Arteries usually possess thicker tunica media compared to veins to withstand high pressure. Capillaries connect arteries and veins and are responsible for the exchange of oxygen, nutrients, and waste in each organ. The walls of capillaries are only one cell in thickness, and are surrounded by a thin vasal lamina and scattered pericytes to facilitate solute transportation. For different purposes, multiple types of engineered vascular grafts have been developed. Tubular grafts of varied diameters were fabricated as substitutes for blocked arteries or veins, while capillary networks and hierarchical vasculature were developed for prevascularized tissue engineering or the regeneration of ischemic tissues. A comprehensive understanding of vascular biology in different tissues facilitates the development of vascularized tissue engineering. Recapitulating the architecture of native vascular networks may pave a new path for engineering large-scale tissues and even organs.
In this review, we briefly introduce the categories of engineered vascular grafts, and then summarize the materials and techniques used for their fabrication. Then, we discuss the relationship between neovascularization and tissue regeneration in peripheral nerves, bone, and the heart, and summarize the current vascularization strategies in these engineered tissues. Finally, we discuss the current challenges and delineate our perspective on the future development of biofunctional engineered vessels.

2. Tissue-Engineered Vascular Grafts

Cardiovascular diseases, including, but not limited to, coronary heart disease (CHD) and peripheral arterial disease (PAD), are the principle cause of death worldwide, especially in developed countries, where cardiovascular diseases affect more people than all types of cancer combined [11,12]. Common disorders such as myocardial infarction (MI) and critical limb ischemia happen as a result of the occlusion or stenosis of blood vessels; thus, the prompt restoration of local blood perfusion is required to prevent heart failure or to aid in the repair of ischemia [13,14,15]. A large diameter (>8 mm) and medium to large diameters (6–8 mm) are usually needed for aortic, iliac, and femoral artery repairs, while small-diameter (1–6 mm) vascular grafting is required for coronary artery replacement caused by atherosclerosis or embolism [6].
Most tissues in the body are nourished by hierarchical tree-like vascular structures with complex and diverse branching configurations. Cells embedded in the biological environment cannot obtain vital nutrients and oxygen, or dispose of metabolic waste without a proper vascular network; e.g., occlusion in mesovasculature (50 μm–1 mm) or microvasculature (<50 μm) can lead to cell death and subsequent tissue necrosis, or even complete tissue dysfunction [16,17]. For tissue engineering thicker than the diffusion limitation distance (100–200 μm), a prevascularized network is required to facilitate the transportation of nutrients and waste in cell-dense constructs [18,19,20]. Hierarchically connected arteriole-sized and capillary-sized built-in vasculature can mimic natural vasculature in tissues, and anastomosis between the engineered constructs and the host is required for the rapid functionalization of engineered tissues after implantation [21,22,23,24].

2.1. Middle- and Large-Diameter Artificial Grafts

Large-diameter artificial grafts were firstly developed due to their fabrication being less complex [25]. Synthetic nonbiodegradable polyethylene terephthalate (PET, Dacron®) was used for the first artificial vascular graft, and was successfully implanted in a human in 1954 by Barkey [26,27]. Since then, a variety of materials, including PET, expanded polytetrafluoroethylene (ePTFE), and polyurethane (PU), have been clinically approved and/or used in clinical trials for commercial vascular grafts as substitutes to autologous vascular grafts [28,29,30]. Due to the high blood flow rates in large vessels, the risk for occlusion is relatively low, and satisfactory results with a patency of 85–90% at 5 years and 75–80% at 10 years were obtained after long-term implantation [25,31]. Medium- to large-diameter vascular grafts have demonstrated more than a 50% patency within 15 years of implantation [30,32,33,34]. However, commercial synthetic grafts are much stiffer and highly hydrophobic compared to native vessels, leading to poor interactions with vascular cells and subsequent reduced biocompatibility [35,36,37]. In addition, polymeric materials are susceptible to bacterial infection, which can lead to graft rejection due to an inflammatory response being evoked [38].
Accessorial modifications to luminal surfaces have been developed to improve biocompatibility and clinical success rates. Bioactive molecules, including heparin [39,40], gelatin [41,42], dopamine [41,43,44], collagen [45,46,47], RGD peptides [48], etc., were immobilized to enhance cell adhesion, proliferation, and reendothelialization by mimicking the natural extracellular matrix (ECM), achieving improved patency rates, reduced bleeding, and the prevention of graft collapse [49,50].

2.2. Small-Diameter Artificial Grafts

Even though synthetic vascular grafts have demonstrated satisfactory outcomes in large- and middle-diameter vascular repairs, polymeric small-diameter grafting can only maintain 39% patency at 5 years, since they fail to recapitulate the key elements of the native vasculature [51,52]. The rough and hydrophobic luminal surface of commercial synthetic grafts fails to allow endothelial cells (ECs) to attach, proliferate, and form a confluent endothelium, which can resist thrombus formation by preventing platelets and red blood cells from adhering to endothelial cells via a barrier composed of glycocalyx [53,54]. Instead, increased interactions with blood cells and protein adsorption take place, accelerating clot formation, and SMCs migration to the intima, inducing intimal hyperplasia, which is the major cause of mid-to-late graft failure [55,56].
Although plenty of studies aiming at capturing the structural and functional properties of native arteries have been carried out, there have been no small-diameter synthetic grafts approved yet for clinical use at this time. In addition, current small-diameter revascularization exclusively relies on autologous grafts [57,58]. Tissue-engineered vascular grafts need to replicate the major features of native blood vessels for long-term functionality and avoid possible complications, including thrombosis, graft failure, and aneurysms [59]. First of all, the mechanical properties of grafts need to combine high strength with elasticity to withstand high-pressure flow without rapture or deformation, and to retain integrity after anastomotic sutures [60]. Secondly, essential biocompatibility is required for EC attachment, proliferation, and endothelium formation to resist thrombosis and intimal hyperplasia [54,61,62]. Thirdly, implanted grafts should maintain a long-term matching to surrounding blood vessels [63]. For example, for nondegradable vascular grafts, the mechanical properties should be kept as the value immediately after fabrication to avoid dilation, delamination, and graft rupture, while the compensation of newly formed ECMs should match the biomaterial digestion for biodegradable tissue-engineered grafts with no aneurysms or intimal hyperplasia development during tissue remodeling [64]. Finally, the production and sterilization process must be facilely expanded from laboratory to industry scale, providing a versatile off-the-shelf small-diameter vascular graft product [56].

2.3. Engineered Vasculature

A dense network of capillaries within 100 μm from one another provides optimal conditions for the transportation of gasses, nutrients, metabolites, and circulating cells to tissues in the body [6,65]. Previous works have demonstrated that prematured microvasculature can improve overall in vivo tissue regeneration, organization, and functionality through the construct–host microvascular integration upon implantation [21,66,67,68]. Various strategies for appropriate revascularized networks in thick engineered constructs have been explored, including optimizing the micro–nanostructure, morphology, porosity, and roughness of the extracellular environment, and introducing biochemical and/or mechanical stimuli [69,70,71,72,73]. However, engineered microvasculature formed through self-assembling are hard to surgically integrate to a host’s vascular network [74]. The ideal vasculature for thick engineered tissue should contain seamlessly connected hierarchical meso- and microvasculature, of which the diameter should follow Murray’s law [75]. The perfusion of vasculature should be achieved as soon as possible, either through surgically or spontaneously anastomosing the mesoscale vessels with the host’s blood vessels [21,22,76].

3. Material Selection

Both synthetic and natural polymers have been used for the production of vascular scaffolds. Synthetic polymers, including polyesters, polyethers, and poloxamers, have been widely used for vascular scaffolds due to their adjustable mechanical strength and degradation rates, thermal stability, and well-established fabrication techniques [77]. However, the lack of cell binding sites and high hydrophobicity of synthetic polymers can impair endothelial cell adhesion and proliferation, as well as causing platelet aggregation and intimal hyperplasia [21,78,79,80].
Natural polymers, such as proteins and polysaccharides, have also been utilized for the development of engineered vessels. Natural polymers generally exhibit more similar properties to native ECMs by providing cell recognition and adhesion domains. Despite having bioactive properties superior to synthetic materials, unmodified natural polymers crosslink mostly via physical interactions, and exhibit poor mechanical properties and rapid degradation [81]. Moreover, a limited supply, batch-to-batch variations, and potential cross-contamination hinder them from reaching off-the-shelf use [82,83]. In order to improve the properties of fabricated vasculature, advances in the isolation, purification, and manufacturing processes of natural polymers are required, and modifications with synthetic functional groups could enhance the mechanical strength to match the mechanical properties of native vessels [56,84].

3.1. Synthetic Polymers

Synthetic materials used for engineered vascular scaffolds can be classified into two categories: nondegradable and biodegradable polymers. Large-, medium-, and small-diameter tissue-engineered vascular grafts composed of each category have been reported since the 1950s [85].

3.1.1. Nondegradable Polymers

Nondegradable polymers, such as ePTFE and PET, have mostly been used in the creation of medium- and large-diameter artificial blood vessel grafts, which have been widely applied in human treatments since the 1970s [86]. The electronegativity and biostability of ePTFE and PET can minimize thrombosis and restenosis in the case of high rates of blood flow and large flow areas in large- and medium-diameter vessels. However, a few of them have been used as vascular scaffolds of <6 mm. Unsatisfactory results, such as cell hyperplasia and the eventual failure of these vascular grafts, were obtained because of their high hydrophobicity, mismatches between the materials and native vessels, and the induced local chronic foreign body response, resulting in thrombi formation [87,88,89]. Alterations have been determined to increase the hydrophilicity to in turn enhance cell adhesion and proliferation [90]. Chen et al. embedded ECM and CD34 in ePTFE to facilitate endothelization and decrease platelet adhesion [91].

3.1.2. Biodegradable Polymers

Compared to nondegradable polymers, biodegradable polymers exhibit better biocompatibility with suitable degradation rates.

PGA

Polyglycolic acid (PGA) is a highly flexible polyester, with a degradation rate of approximately 6–8 weeks, inducing little inflammatory response [92,93]. It has been reported that PGA vascular grafts were able to remain patent up to 4 weeks after implantation in vivo [94]. The in vivo degradation rate is too fast to match the regeneration rate for clinical applications. Other polymers, such as polylactic acid (PLA), poly (lactide-co-caprolactone) (PLCL), and poly (DL-caprolactone-co-lactic acid) (PCLLA), were added to slow the in vivo degradation rate, introducing satisfactory remodeling observed up to 6 months (Figure 2) [95,96,97]. Niklason et al. seeded vascular smooth muscle cells (VSMCs) onto PGA conduits and cultured them with a pulsatile flow for 8 weeks in vitro. The full maturation of the vascular grafts was achieved by replacing the PGA scaffolds with VSMC-secreted collagen matrix deposition [94,98,99]. After decellularization, the engineered blood vessels were tested in multiple small and large animal models, showing over an 88% patency in 3.5 mm sized grafts up to 6 months [100,101]. The engineered human acellular vessels were named as Humacyte (Humacyte Inc., NC, USA), with clinical trials being underway since 2012 [74,102,103].

PLA

PLA is a nonimmunogenic polyester approved by the FDA. The degradation rate of PLA is relatively slow (6–8 months). However, the highly hydrophobic surface of PLA impairs cell development, and results in the unsatisfactory performance of vascular scaffolds composed of only PLA [104]. A copolymer of PLA and polycaprolactone (PCL) was developed to adjust the biodegradation rates, biocompatibility, and elasticity for facilitating endothelial cell adhesion and proliferation in vitro, as well as endothelization and collagen fiber deposition in vivo [105,106]. Combining PLA with bioactive molecules such as heparin, collagen, and chitosan through electrospinning can greatly increase the infiltration and attachment of endothelial cells, thus, improving the hemocompatibility of the vascular grafts [104,107]. Hashi et al. reported that by seeding bone-marrow-derived stem cells (BMDSCs) onto PLA vascular scaffolds, the patency rates could remain at nearly 100% after 60 days postoperative in a rat model [108].

PCL

PCL serves as the most commonly used polymer in vascular scaffolds due to its excellent biocompatibility, suitable in vivo degradation rates, low cost, mature fabrication techniques, and off-the-shelf properties [109]. Nevertheless, the performance of PCL grafts is similar to ePTFE grafts in patency rates, in spite of their superior biocompatibility, faster endothelization, and less-consequential stenosis-free rate [110]. By increasing the porosity or reducing the hydrophobicity by blending natural polymers, the mechanical strength of PCL scaffolds can be adjusted to those of the native artery (1–2 MPa), and reendothelialization can be promoted without blood coagulation (Figure 3) [111,112,113].

PGS

Poly (glycerol-co-sebacate) (PGS) is a highly biocompatible but rapidly bioabsorbable elastomer. The degradation rate of PGS is similar to PGA, while the hemocompatibility of PGS is superior, as it can not only facilitate the proliferation of ECs and the infiltration of VSMCs, but also prevent platelet adhesion and inflammation [114,115,116,117]. Khosravi et al. used PGS as the inner layer with a PCL outer sheath, which exhibited excellent long-term patency, a mild inflammatory response, and the regeneration of organized ECs and SMC constructs [118].

PEG

Polyethylene glycol (PEG) is a nontoxic biodegradable polymer approved by the FDA due to having nonimmunogenic and nonantigenic performance in biological applications [119]. Diacrylate-derived PEG (PEGDA) maintains the nonthrombogenic and hyperplasia reducing properties of PEG and is able to gel rapidly in the presence of photoinitiators and light irradiation [120,121]. Hahn et al. developed a tissue-engineered vascular scaffold by enveloping smooth muscle progenitor cells within a PEGDA hydrogel decorated using adhesive ligands and collagenase degradable sequences [122]. After culturing in vitro for 8 weeks, the elastin and collagen deposition reached the native arterial favorable range [122]. Hou et al. developed amphiphilic and fatigue-resistant small-diameter vascular grafts by combining PEGDA with polyhydroxyalkanoates (PHAs) [123]. High patency rates were achieved 3 months postoperation, and cell infiltration, as well as remarkable tissue regeneration, was observed (Figure 4) [123].

3.2. Natural Polymers

Natural components of the ECM have attracted increasing attention as alternatives for less bioactive synthetic grafts. Proteins and polysaccharides, such as collagen, gelatin, elastin, fibrin, chitosan, alginate, and cellulose, possess better biocompatibility and bioactivity due to the remaining biomimetic binding sites for cell development that promote the recruitment, colonization, and proliferation of intrinsic cells [124,125,126,127,128].

3.2.1. Collagen and Gelatin

Collagen and its derivative gelatin are the most used natural polymers in vascular tissue engineering. Collagen is the main component of the ECM in most tissues, including blood vessels [129]. However, the in vivo degradation rate of collagen is approximately 5 h, which is too short for clinical applications [130]. Crosslinked collagen conduits after dehydration can match a mammalian vein value with a burst pressure of ~1300 mmHg and a compliance of 1.7%/100 mmHg, but a low strength at anastomosis interfaces [131]. Circular electronspun meshes of collagen were developed by Zhang et al., possessing a similar dynamic compliance and increased rupture force to saphenous veins [132]. Collagen and gelatin were also used in combination with synthetic polymers for vascular scaffold fabrication through electrospinning [42,133]. By altering the concentration of natural proteins, the hemocompatibility and mechanical strength can be balanced for vascular cell growth and a reduction in the local inflammatory response [109]. In order to increase the mechanical strength of the grafts, gelatin was modified with methacrylate groups to form a semisynthetic gelatin methacryloyl (GelMA), which could maintain the bioactivity of gelatin with adjustable mechanical properties [134,135]. The cell-attached peptide motifs of GelMA hydrogels mimic the essential properties of native ECMs for cell proliferation and spread, while the metalloproteinase-responsive peptide motifs lead to dilation and a decrease in mechanical strength after in vivo implantation [136]. Liang et al. developed a series of diameter-tunable microtubes with enhanced mechanical strength by blending a H-bonding monomer N-acryloyl glycinamide nanoclay with GelMA (Figure 5) [137]. The coaxial printed vascular grafts exhibited a Young’s modulus of ~21 MPa, a burst pressure of ~2500 mmHg, a suture retention strength of ~280 gf, and an antifatigue performance of ~200 cycles [137].

3.2.2. Elastin

Elastin is another predominant component of the ECM in arterial tissues. It is able to promote SMCs’ contractility and introduce aligned collagen fiber deposition that mimics the construction of the native ECM [138,139]. Elastin was used in association with PCL by Wise et al. to enhance its patency and mechanical properties. However, the durability was just like the human internal mammary artery tested in a rabbit model [140]. Elastin has been proved to play a significant role in aortic morphogenesis and the prevention of intimal hyperplasia in native tissue, while only a little in vivo research has been carried out due to its limited resources, high expense, and batch-to-batch variations [141]. Lee et al. modified human-recombinant elastin with methacrylic groups and fabricated it into vascular scaffolds via extrusion printing with GelMA (Figure 6) [142]. The endothelium barrier was observed in the printed vascular grafts to have a minimal inflammatory response and efficient biodegradation in vivo.

3.2.3. Silk Fibroin

Silk fibroin (SF) proteins are natural fibers characterized by mechanical hardness, processing malleability, self-assembly, and biocompatibility [143]. Different from collagen and elastin, SF can be produced into nanofiber tubes alone or in combination with bioactive macromolecules as support [144,145,146,147]. Marelli et al. developed a SF tubular scaffold coated with collagen that exhibited a physiologically relevant burst pressure (894 ± 24.91 mmHg) and compliance (3.24 ± 0.58%/80–120 mmHg) [147]. Liu et al. blended heparin into the inner layer of SF tubes; the bilayered grafts showed successful SMCs and fibroblast infiltration in the outer layer as well as low thrombogenicity on the luminal surface [148]. Cattaneo et al. engineered SF scaffolds through electrospinning and in vivo tests, showing that elastic lamina and vasa vasorum could be formed within 7 days in a rat abdominal aorta [149]. Filipe et al. found near-complete endothelialization by 6 weeks postsurgery in a rat aortic interposition grafting model [35], and Enomoto et al. reported that small-diameter SF-based grafts (1.5 mm diameter) retained an 85% patency at 1 year after implantation [150].

3.2.4. Chitosan

Chitosan is a mucoadhesive and biodegradable polysaccharide that is the second most used natural polymer next to collagen in vascular tissue engineering [58,151]. Chitosan has a similar structure to glycosaminoglycans (GAGs) in the ECM and exhibits biodegradability, anticoagulation, anti-inflammatory, and antimicrobial properties [152]. Chitosan has been used for the surface modification of several polymer-based vascular scaffolds, with its cationic property enhancing cell adhesion, morphology, and growth [153,154,155]. Chitosan-based tubular scaffolds have also been fabricated, with their porosity structure providing a favorable microenvironment for cell infiltration and tissue integration [156,157]. Aussel et al. engineered a vascular scaffold with chitosan through sequential NaOH-induced gelation and gaseous NH3-induced gelation [158]. By modulating the chitosan concentration, the mechanical properties of the tubes could reach a suture retention strength of 1.1 ± 0.5 N, a compliance of 11.50 ± 3.90%/100 mmHg, and a burst pressure of 1313 ± 196 mmHg [158]. Yan et al. fabricated a vessel-like-structured chitosan hydrogel via a templated electrodeposition, of which the diameter could be lowered to 400 μm in an ambient environment [159].

3.2.5. Alginate

Alginate is an anionic polysaccharide composed of guluronic acid and mannuronic acid [160]. Alginate can go through ionotropic gelation in the presence of divalent ions such as Ca2+. However, there are no cell-adhesive ligands on it before chemical modification. Cell-adhesive motifs, such as RGD peptides, gelatin, heparin, etc., were used to modify alginate to optimize the microenvironment for cell development [161]. Gao et al. incorporated decellularized ECMs into an alginate hydrogel and developed a series of vascular tubes through the coaxial printing technique [162,163,164]. It was reported that the engineered EC-laden grafts exhibited not only an endothelium maturation, but also selective permeability and antithrombogenicity [164]. Moreover, when combined with an atorvastatin delivery microvehicle, endothelial progenitor cells (EPCs) could quickly eliminate local ischemia in a mouse model (Figure 7) [163].

4. Fabrication Methods for Tissue-Engineered Vascular Grafts

4.1. Biobased Techniques

4.1.1. Decellularization

To obtain decellularized scaffolds that can maintain the native architecture and composition of blood vessels, cellular components were removed through the use of detergents, chelators, enzymes, freeze/thawing, agitation, or their combinations [6,165,166,167]. A decellularized ECM (DECM) possesses biochemical and mechanical cues that can promote cell adhesion, proliferation, differentiation, and tissue organization (Figure 8) [64,168]. Allogenic human DECM vessels have demonstrated successful clinical outcomes in clinical trials [169]. However, the lack of sources and the potential pathogenicity prevent it from off-the-shelf applications. The xenogenic DECM has a prolific source and exhibits similar or increased compliance and burst pressure compared to human-native vessels [25,170,171], while the incomplete removal of cellular components and immunological ECM components can elicit a severe inflammatory response, which can lead to acute thrombosis and calcification [169,172]. Surface modifications were used to improve the endothelialization of the DECM scaffolds. Dimitrievska et al. deposited heparin as a continuous shielding layer to the underlying thrombogenic ECM, reducing platelet adhesion and improving blood compatibility [173]. Jiang et al. decorated DECM grafts with an antioxidant poly (1, 8-octamethylene-citrate-co-cysteine) (POCC), achieving reduced oxidative tissue damage and calcification [174]. Pre-endothelium through recellularization with ECs, bone-marrow-derived cells, and genetic knockout cells can greatly reduce thrombosis formation and increase the long-term patency up to 60% [175,176,177].
Decellularized engineered scaffolds have also been fabricated. There are two strategies mostly used to produce cell-laden engineered grafts before decellularization: one is culturing cells on native or synthetic polymer scaffolds to produce an adequate ECM [100,178], and the other is to embed polymeric rods in vivo to form an autologous fibrocellular tissue capsule [88,179]. Syedain et al. cultured fibroblasts in a fibrin tube and disposed of the cellular components to acquire the ECM-deposited grafts [178]. The implanted scaffolds showed fibroblast and SMC infiltration in vivo with collagen and elastin deposition. In addition, the diameter and volume increased with the growth of the host without rupture, demonstrating successful in vivo remodeling. Geelhoed et al. subcutaneously implanted copolymer poly(ethylene oxide terephthalate)-poly (butylene terephthalate) (PEOT/PBT) rods in goats and obtained fibrotic tissue tubes after one month [180]. After decellularization, the DECM grafts possessed sufficient mechanical strength with a bursting pressure of 2382 ± 129 mmHg and a suture retention strength of 1.97 ± 0.49 N. In addition, the endothelialization and abundant expression of collagen were observed at 8 weeks after grafting.

4.1.2. Self-Assembly

In this approach, vascular cells are placed in a 3D environment and organize themselves to form the biomimetic vascular structure with/without biochemical and/or mechanical stimuli. Three techniques have been used to facilitate vascular self-assembly up till now, including cell-sheet assembly, microtissue aggregation, and cell printing [181]. For the cell-sheet assembly strategy, engineered cell sheets were rolled over a mandrel to form tubular structure, in which the cell junction and produced ECMs were maintained to promote cell communications [182,183]. The deposition of ECM proteins, such as collagen, laminin, fibronectin, and elastin, was facilitated via in vitro culturing, as well as the remolding of grafts into native-like constructs with equivalent mechanical properties [183]. L’Heureux et al. engineered SMCs and dermal fibroblast monolayer sheets and rolled them around a mandrel for the successful generation of a bilayered vascular graft via self-assembly [184]. After dynamic conditioning for 8 weeks in vitro and seeding ECs within the lumen to form an endothelium, the tubular grafts showed a bursting pressure of 2600 mmHg, while the grafts exhibited bleeding and failure in vivo at 7 days postimplantation [184]. To optimize this technique, L’Heureux et al. used human fibroblasts and increased the in vitro dynamic conditioning to 28 weeks [47]. The scaffold-free engineered grafts were named Cytografts (Cytograft Tissue Engineering, Inc., Novato, CA, USA), and were used in a clinical trial with ten patients. The grafts showed promising results in the clinical trial; in spite of one patient having died for a cause not related to the graft and three failures due to dilatation, thrombosis, or aneurysms, the grafts had a 78% patency at 1 month and 60% patency at 6 months [185]. Cell sheets combined with synthetic polymer sheets were used to enhance the mechanical properties and handling of the grafts [186,187], and the micropatterning of the polymer sheets was carried out to obtain cell sheets with a predesigned cell alignment [188,189,190].
For the microtissue aggregation strategy, cells were encapsulated into an ECM mimicking a 3D environment with high density. Vascular-like constructs were formed via self-organization and the secretion of the ECM [191]. Kelm et al. fabricated a tubular graft through the aggregation of human-artery-derived fibroblasts and human umbilical vein endothelial cells (HUVECs); the accumulation of vessel-like tissue occurred within 14 days with an enhanced ECM expression and maturation [192]. The bioprinting strategy also embeds cells into an ECM mimicking ink, but deposits via the 3D-printing technique. Mironov et al. engineered a tubular graft through the fusion of printed cell spheroids [193], and Norette et al. developed a complex vascular tree with connected branches of accurate diameter and wall thickness [194]. Andrique et al. produced mature functional blood vessels by engineering hollow alginate hydrogel tubes internally coated with ECMs (Figure 9) [195]. The biomimetic configuration of lumens was achieved through the self-assembly of ECs and SMCs and perfusability, with contractility in response to vasoconstrictor agents being observed after the vesseloids reached homeostasis. Nevertheless, most bioprinted self-assembled vascular grafts are proof-of-concept for now, with the lack of mechanical properties and long-term stability hindering them from clinical applications.

4.2. Engineering-Based Techniques

4.2.1. Electrospinning

In electrospinning, polymer solutions are dispersed through a strong electrical field onto a grounded or oppositely charged metal collector, creating continuous nano-to-microscale fibers for the formation of a thin film [196,197]. The diameter, morphology, alignment, and thickness of the electrospun fibers can be adjusted through the parameters, including the flow rate, solvent system, surface tension, supplied voltage, tip-to-collector distance, humidity, and temperature [198,199,200,201,202]. For the fabrication of engineered vascular graft, fibrous scaffolds are collected either on a rotating mandrel or, subsequently, rolled into a tubular structure to mimic the structure of native blood vessels [203]. Synthetic polymers, such as PCL, PLA, PGA, PLCL, poly (lactic-co-glycolic acid) (PLGA), poly (carbonate urethane) (PCU), and polyurethane (PU), and natural polymers, such as collagen, gelatin, chitosan, elastin, silk fibroin, and lecithin, have been used for vascular scaffold fabrication through electrospinning [35,144,204,205,206,207,208]. The electrospun grafts of natural polymers possess better biocompatibility, but inferior mechanical properties, while scaffolds fabricated with synthetic polymers exhibit comparable mechanical strength to native blood vessels, but a lack of bioactivity for cell adhesion, proliferation, and, development [35,204,209]. The electrospinning of blended natural and synthetic polymers has been used to fabricate vascular grafts with both superior biocompatibility and mechanical strength. By optimizing each component’s ratio in hybrid polymers, the presence of synthetic polymers, such as PCL, PLGA, or PLCL, helped retain the superior mechanical properties without interfering with the cell viability (Figure 10) [208,210].
Advanced electrospinning technologies, such as coelectrospinning, simultaneous electrospinning/electrospraying, sequential electrospinning, coaxial electrospinning, and sacrificial electrospinning, have been used to fabricate grafts with more complicated structures and superior properties [211]. Pan et al. tailored the degradation rates to better match in vivo cell infiltration and ECM deposition by adding a slow-degradation PCL into fast-degradation polydioxanone fibers via coelectrospinning [212]. Hyaluronic acid–gelatin nanoparticles were deposited simultaneously during PCL/collagen nanofiber electrospinning by Ekaputra et al., resulting in the enzymatical degradation of the nanoparticles creating pockets that improved cell infiltration [213]. Wang et al. developed an asymmetric vascular graft, of which the inner layer exhibited anticoagulant properties, and the outer layer showed antibacterial properties due to the sequential electrospinning of an inner layer of PCL/carboxymethyl chitosan and an outer layer of PCL/chitosan [214]. A gelatin shell was fabricated out of a synthetic polymer core using the coaxial electrospinning technique to enhance the proliferation of ECs and SMCs [113]. Moreover, bioactive polymers could be embedded in the core structure, performing a sustained release after implantation [215,216]. Sacrificial electrospinning has been used to create micro-to-mesovasculature networks. Water soluble fibers, such as polysaccharide pullulan and poly(N-isopropylacrylamide) (PNIPAM) fibers, were electrospun and embedded in bulk ECM-mimicking scaffolds, and after the dissolution of the fibers, perfusable channels with diameters from 1 to 55 μm could be obtained [217,218,219].

4.2.2. Molding

Both tubular-shaped grafts and vasculature with interconnected branching structures can be developed through molding. In this approach, a polymer solution is cast into a hollowed-out mold and removed after solidification in the predesigned architecture [25,131]. A customized structure can be facilely produced by altering the geometry of the mold [220]. Tubular scaffolds can be fabricated with a simple device being set up, including an annular mold with an internal rod. The luminal diameter and thickness of the wall can be altered through the combination of differently sized annular molds and internal rods. However, the mechanical strength of the molded vascular grafts is relatively low compared to native blood vessels [221]. Strategies such as applying in vitro dynamic conditioning to cell-laden scaffolds, compacting the grafts through centrifugal forces, and the incorporation of porous meshes have all been used to strengthen molded tubular structures, especially those composed of natural polymers [222,223]. Associated with pore-generating techniques, the porosity of the scaffolds can be tuned to enhance cell infiltration. Lee et al. fabricated a PGS graft with a 75–80% porosity through salt leaching, which allowed for increased cell infiltration and ECM remodeling [224]. When combined with an inserted mesh or electrospun coating, this can reinforce the wall of the molded vascular grafts [225], while cell adhesion and orientation can be altered through surface patterning [226].
Sacrificial molding has been used to develop mesovasculature. The mold is firstly fabricated by using sacrificial polymers and then embedded within an ECM-like mesh [64]. After removing the mold through physical separation or dissolution, the ECM-like scaffold is left with interconnected mesochannels [64]. Chen et al. cast a 3D-reconstructed PLA filament in polydimethylsiloxane (PDMS) and, subsequently, dissolved it, producing vessel channels with endothelium through EC seeding [227]. In this work, blood vessel networks from different tissues, from mice to humans, have been successfully transcribed to physical microfluidic devices. Nevertheless, vasculature fabricated via this approach usually remain the bulk of ECM-like scaffolds, limiting their application in the theoretical study of the transportation of blood vessels, although not suitable for clinical applications.

4.2.3. 3D Printing

Three-dimensional printing is another technology used in tissue engineering that has drawn increasing attention over the past decades. It is an additive manufacturing method in which inks are deposited layer-by-layer until the formation of predesigned computer-aided design (CAD) models. Extrusion, inkjet, and stereolithography are the three most used strategies for vascular graft development [6].
Extrusion-based printing is the most popular bioprinting technique that uses a fluid dispensed through mechanical or pneumatic actuation [228,229]. The technique can be applied to form large-scale 3D constructs with inks of varying viscosities, while the resolution is restricted in the range of 200–1000 μm, with tradeoffs present between printability and biocompatibility [230]. Vascular scaffolds can be fabricated with thermoplastics via fused deposition modeling; however, the high temperature required can deactivate bioactive molecules and damage cells [231]. As a result, acellular grafts have been fabricated with subsequent surface modifications to improve their hemocompatibility [232,233]. Instead, nature-derived polymers, such as alginate, gelatin, and fibrin, have been used for cell-laden vascular graft printing [234]. Tubular structures can be fabricated using layer-by-layer deposition through the repeated circular movement of the printer nozzle or direct extrusion in tubular architecture via coaxial needles. Two or more layered concentric tubes can be developed in a one-step process by applying coaxial printing. For example, alginate tubes can be fabricated via the coaxial printing of a CaCl2 core surrounded by a sodium alginate shell, while a GelMA tube can be obtained by extruding a gelatin core surrounded by a GelMA shell and removing gelatin by increasing the temperature after photocrosslinking [235,236,237].
Inkjet-based bioprinting is a noncontact printing technique, in which cell aggregates are firstly dispensed and then deposited onto a substrate, forming a vascular structure through self-assembly [238,239]. Even though inkjet-based bioprinting usually possesses a better resolution and faster printing speed compared with extrusion-based techniques, the required limited cell density and use of low-viscosity bioink, to avoid nozzle blockage, often result in a relatively weak mechanical strength [240]. Laser-assisted bioprinting and deposition into a reservoir with crosslinkers have been used to strengthen printed scaffolds, and complex structures (such as zig-zag tubes and vasculature with both horizontal and vertical bifurcations) and vascular scaffolds with increased cell density can be produced, respectively [241,242,243].
Stereolithography-based bioprinting uses digital micromirror arrays to precisely control the light irradiation of the printed area in which photo-sensitive bioinks accumulate layer-by-layer through photocrosslinking [244]. Both acellular and cellular modes work for this approach. Stereolithography has its advantages in the throughput of vascular grafts with complex geometries in high resolution, while long exposure to UV light during the curing process may potentially damage cell viability [245,246]. Melchiorri et al. transcribed a recipient’s specific curvature into a customized vascular graft through stereolithography-based printing, which sustained the patency and functionality up to 6 months after implantation (Figure 11) [245].

4.2.4. Laser Degradation

Laser energy has been used to selectively degrade or ablate predetermined regions within hydrogels with a two-photon absorption profile, including natural proteins, such as collagen, elastin, silk, high-protein-content hybrid materials (e.g., PEG–fibrinogen, PEG–albumin, and PEG–gelatin), PEG, and PEGDA [247,248,249,250,251,252]. High-resolution patterned microvasculature with mm scale z-depth penetration can be fabricated using laser degradation with/without cell encapsulation [6]. Brandenberg et al. demonstrated it by fabricating microchannels in a collagen hydrogel with a 50 μm diameter [251]. Heintz et al. combined a photodegradation hydrogel and image-guided laser control to generate a complex microvasculature that accurately recapitulated the in vivo vascular architecture and possessed a diameter size down to 3 μm (Figure 12) [247]. Intraluminal topography can also be produced through the use of the laser degradation method, and a complete endothelium can be acquired to promoted cell adhesion, migration, and proliferation [253].

5. Vascularized Tissue Engineering

The vasculature plays a fundamental role in the supply of blood, oxygen, and nutrients, and removing metabolic wastes for the maintenance of the function of all tissues, except the corneas, cartilage, and skin epithelium. A physiologically relevant tissue comprises of at least 1 × 108 cells/cm3, which requires a high level of oxygen and nutrient supply [254]. The prevascularization of tissue-engineered grafts is preferred, especially for large-scale scaffolds. Peripheral nerves, bone, and the heart are tissues of which normal functions highly rely on blood supply. For example, a lack of blood supply in bone may lead to osteonecrosis, while the occlusion of blood vessels in the heart may lead to myocardial infarction. For this reason, vascularized tissue engineering scaffolds were designed for the regeneration of peripheral nerve, bone, and heart tissue. In this section, strategies for vascularized tissue engineering developments in peripheral nerve, bone, and heart tissue were summarized (Table 1).

5.1. Peripheral Nerve Tissue Engineering

The vascular system of peripheral nerves, also named vasa nervorum, can be categorized into two groups: the extrinsic system, which comprises a series of arteries and veins around the surface of peripheral nerves and supplies the epineurial and perineurial regions, and the intrinsic system, which is composed of small arteries and supplies the inner endoneurial nerve compartment [255,256]. While extrinsic vasculature can autoregulate in response to external physiological changes, intrinsic vessels lack this ability, and are susceptible to ischemia when the systemic blood pressure drops [257,258,259]. To minimize the interference induced by fluctuations in systemic blood pressure, peripheral nerve tissues are prone to low oxygen conditions with larger capillary distances in the vasa nervorum compared to other tissues [260]. Both extrinsic and intrinsic vasculature are important to nerve regeneration after injury. Blood vessels can proceed with Schwann cell migration and proliferation in the injured region, facilitate the formation of the Bungner bands, and guide axons growing from the proximal stump to the distal end. In addition, endothelial cells are able to secrete bioactive molecules, such as vitronectin, heparin sulphate proteoglycans, and the brain-derived neurotrophic factor (BDNF), to promote neurogenesis and nerve regeneration [261,262,263].
The first vascularized nerve graft was an ulnar nerve harvested by Strange et al. together with its blood supply [264]. Enhanced nerve regeneration in both length and size were obtained with vascularized nerve grafts, and Kanaya et al. demonstrated an improved sciatic function index (SFI) compared to the nonvascularized group in a rat sciatic nerve model [265,266]. However, autologous vascularized nerve grafts have limited resources and usually deteriorate the donor-site morbidity and scarring due to the removal of the local vascular supply, especially for large nerves. To overcome this disadvantage, nerve grafts were preimplanted between a host artery and vein for the generation of a microcirculation network. Ozcan et al. produced a vascularized amnion tube by embedding it between the femoral artery and vein for 3 weeks, subsequently using it to bridge a 1 cm long femoral nerve gap [267]. The vascularized amnion conduits exhibited superior performance in nerve regeneration compared to nonvascularized grafts in terms of the axon diameter, myelin thickness, and the number of myelinated axons. Iijima et al. demonstrated that PGA nerve conduits prevascularized with host superficial inferior epigastric vessels promoted remyelination and regeneration in a rat sciatic nerve injury model [268]. Nevertheless, a two-stage surgery was required in this approach, which prolonged the waiting time and could increase the risk of complications. Blood-vessel-including tabulation is another strategy for the fabrication of vascularized nerve grafts. In this approach, native blood vessels are wrapped with the nerve conduit. Kakinoki et al. bridged a sciatic nerve gap of 25 mm with a silicone tube encapsulating sural vessels in a rat model, demonstrating regeneration across the nerve gap and the reinnervation of the tibialis anterior muscle [269]. With the development of vascular tissue engineering, autologous blood vessels used in vascularized nerve grafts can be replaced with engineered small-diameter vascular grafts or micro- and mesoscale vasculature, avoiding the sacrifice of donor nerves and their blood supply, as well as the need for a two-stage surgery.

5.2. Bone Tissue Engineering

Blood vessels serve as a structural template in bone regeneration. Bone development takes place around them, and subsequent bone homeostasis is induced by them in the osteogenic environment [270]. The coupling of angiogenesis and osteogenesis occurs from skeletal development to postnatal bone repairing. The vascular endothelial growth factor (VEGF) is the master regulator for the effective coupling of angiogenesis and osteogenesis. During bone repair, the VEGF is produced by bone progenitor cells and osteoblasts to promote the migration and proliferation of ECs, while the ECs secrete osteogenic factors such as bone morphogenetic protein-2 (BMP-2) and insulin-like growth factor (IGF-1) to support osteoblast differentiation [271,272,273].
Several strategies have been used to develop synergistic vascularization and osteogenesis in engineered bone scaffolds, including ion doping, adding growth factors, altering topography, and the coculturing of cells [69,70,274,275,276]. Bioactive ions, such as Mg2+, Ca2+, Co2+, Cu2+, and Se2+, have been used as chemical inducers of the HIF-1α/VEGF signaling pathway to promote angiogenesis [277,278,279]. Various ion-doped materials have been fabricated, including bioglasses, metal–organic frameworks (MOF), metal scaffolds, and hybrid scaffolds, with synthetic polymers [280,281,282]. Improved vascularized bone formation has been demonstrated by Ma et al. on a Mg-doped tantalum scaffold in a rat femur condyle bone defect model [278]. The VEGF and its derived angiogenic peptides have also been introduced into bone grafts to enhance neovascularization. Li et al. loaded an angiogenic peptide QK into a GelMA-based injectable hydrogel, which exhibited synergistic osteogenic and angiogenic effects with a 58.2 ± 3.5% BV/TV ratio [283]. Furthermore, a mesoporous structure (2–50 nm) can not only promote the proliferation and differentiation of osteoblasts and ECs, but also increase the bone–matrix interface strength and stimulate mineralization [284]. The coculture of angiogenic cells with bone progenitor cells has been shown to have positive effects both on neovascularization and osteogenesis [285,286]. Zhou et al. reported enhanced capillary and bone formation in a BMSC/EC coculture system [287], and Wei et al. demonstrated cell–cell communication between human mesenchymal stem cells (hMSCs) and HUVECs taking place at a distance of ≤200 μm via paracrine secretion [288]. Despite current strategies for vascularized tissue engineering bone, prevascularized scaffolds for large defect repairs remain challenging, as the required microenvironment for osteogenesis and angiogenesis differs from mechanical strength to biochemical cues. Three-dimensional printing techniques can serve as potential solutions to this dilemma, although bioinks with adjustable mechanical strengths are required.

5.3. Heart Tissue Engineering

Cardiovascular diseases remain the leading cause of death, with heart failure potentially occurring due to the subsequent functional impairment of vasculature [289,290]. The coronary vasculature is crucial for cardiac development, homeostasis, and regeneration, as the heart is a highly vascularized organ where every cardiomyocyte is located close to a capillary [7,291]. ECs can regulate cardiac remodeling via angiocrine signals. For example, the cardioactive growth factor neuregulin-1 (NRG-1) released by ECs can promote cardiomyocyte proliferation and growth, as well as being capable of regulating the myocardial response to infarction and cardiomyocyte hypertrophy [292,293,294,295]; endothelin-1 secreted by ECs can regulate cardiomyocyte contractility and induce tissue remodeling [296].
The injection of stem cells and the implantation of cell-loaded or acellular scaffolds are the most used strategies for heart regeneration; however, the low cell survival rates and apoptosis due to a lacking oxygen and nutrient supply limit their clinical applications. Prevascularization is desired in cardiac tissue engineering, and connection between the engineered vasculature to the host is optimal for quickly supplying the implanted grafts. Coculture with vessel-forming cells is the most used prevascularization method at this time [297]. In this approach, endothelial cells and associated mural cells are cultured with cardiomyocytes and/or stem cells to promote vasculature formation [298,299]. Cocultured spheroids have been fabricated to provide suitable 3D structured for cell–cell and cell–matrix interactions, and interconnections to surrounding host blood vessels with blood perfusion were obtained [300]. Angiogenic growth factors, such as the VEGF, platelet-derived growth factor (PDGF), and fibroblast growth factor (FGF), have also been incorporated with synthetic, natural, or hybrid grafts to enhance the vascularization [301,302,303]. Moreover, microvascular tubes fabricated with cell sheet technology, as well as microvasculature developed via 3D printing or microfabrication, have been used for the prevascularization of engineered cardiac tissues. Well-perfused microchannels have been fabricated by coculturing ECs with cardiac cell sheets in a collagen hydrogel [304,305]. One strategy for prevascularization using 3D printing involves developing constructs with interconnected microchannel networks first, and then seeding with ECs, SMCs, and/or fibroblasts. Integration to the host’s vasculature can be achieved with this approach [306,307]. Another strategy is simultaneously bioprinting vasculature with surrounding cardiac tissue. Szklanny et al. engineered a hierarchical vasculature with millimetric tubular grafts and 3D-bioprinted vascularized tissue interconnections, and demonstrated their successful support of the in vitro functionality of cardiomyocytes (Figure 13) [20].

6. Conclusions and Perspectives

Engineered middle- and large-diameter vessels composed of nondegradable synthetic materials have been used for decades, while some biodegradable ones have reached the early stages of clinical translation. A comparable or better patency in long-term performance has been achieved, compared to autologous grafts. For small-diameter vascular grafts, acellular ECM-deposited products, such as Humacyte, have shown more potency in clinical applications and are currently under clinical trials. Nevertheless, most other works stay in the preclinical stage. Advanced biofabrication technologies have brought increasing vasculature of both micro- and mesoscale in recent years. However, limited results for the associated prevascularized engineered tissues have been reported. The lack of a comprehensive understanding of the specific vascular biology in each tissue, along with limited fabrication strategies for high-resolution large-scale scaffolds, are some of the road-blocks preventing the incorporation of established vasculature with engineered tissues. However, we believe that further development in materials science, stem cell biology, and biofabrication technologies can advance the engineering of functionally and hierarchically prevascularized tissues to off-the-shelf applications.

Author Contributions

Conceptualization, L.-P.W.; writing—original draft preparation, J.C.; writing—review and editing, D.Z.; supervision, M.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key R&D Program of China (No. 2019YFA0110500), the Guangzhou Science and Technology Program (SL2022B03J00487), the Key Science and Technology Project of Guangzhou City (No. 2023B03J1231), the Guangdong Pearl River Talents Program (No. 2017GC010411), the National Natural Science Foundation of China (32201111), and the Chinese Postdoctoral Science Foundation (2021M691464).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Acknowledgments

The authors thank the National Key R&D Program of China, the Guangzhou Science and Technology Program, the Key Science and Technology Project of Guangzhou City, the Guangdong Pearl River Talents Program, the National Natural Science Foundation of China, and the Chinese Postdoctoral Science Foundation for the financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ngo, M.T.; Harley, B.A. Angiogenic biomaterials to promote therapeutic regeneration and investigate disease progression. Biomaterials 2020, 255, 120207. [Google Scholar] [CrossRef]
  2. Mohyeldin, A.; Garzón-Muvdi, T.; Quiñones-Hinojosa, A. Oxygen in stem cell biology: A critical component of the stem cell niche. Cell Stem Cell 2010, 7, 150–161. [Google Scholar] [CrossRef] [PubMed]
  3. Yin, H.; Price, F.; Rudnicki, M.A. Satellite cells and the muscle stem cell niche. Physiol. Rev. 2013, 93, 23–67. [Google Scholar] [CrossRef] [PubMed]
  4. Lane, S.W.; Williams, D.A.; Watt, F.M. Modulating the stem cell niche for tissue regeneration. Nat. Biotechnol. 2014, 32, 795–803. [Google Scholar] [PubMed]
  5. Putnam, A.J. The instructive role of the vasculature in stem cell niches. Biomater. Sci. 2014, 2, 1562–1573. [Google Scholar] [CrossRef]
  6. Fleischer, S.; Tavakol, D.N.; Vunjak-Novakovic, G. From arteries to capillaries: Approaches to engineering human vasculature. Adv. Funct. Mater. 2020, 30, 1910811. [Google Scholar] [CrossRef]
  7. Rafii, S.; Butler, J.M.; Ding, B.-S. Angiocrine functions of organ-specific endothelial cells. Nature 2016, 529, 316–325. [Google Scholar] [CrossRef]
  8. Chaudhuri, O.; Koshy, S.T.; Branco da Cunha, C.; Shin, J.-W.; Verbeke, C.S.; Allison, K.H.; Mooney, D.J. Extracellular matrix stiffness and composition jointly regulate the induction of malignant phenotypes in mammary epithelium. Nat. Mater. 2014, 13, 970–978. [Google Scholar] [CrossRef]
  9. Crosby, C.O.; Zoldan, J. Mimicking the physical cues of the ECM in angiogenic biomaterials. Regen. Biomater. 2019, 6, 61–73. [Google Scholar] [CrossRef]
  10. Qadura, M.; Terenzi, D.C.; Verma, S.; Al-Omran, M.; Hess, D.A. Concise review: Cell therapy for critical limb ischemia: An integrated review of preclinical and clinical studies. Stem Cells 2018, 36, 161–171. [Google Scholar] [CrossRef]
  11. Virani, S.S.; Alonso, A.; Aparicio, H.J.; Benjamin, E.J.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Cheng, S.; Delling, F.N. Heart disease and stroke statistics—2021 update: A report from the American Heart Association. Circulation 2021, 143, e254–e743. [Google Scholar]
  12. Virani, S.S.; Alonso, A.; Benjamin, E.J.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Chang, A.R.; Cheng, S.; Delling, F.N. Heart disease and stroke statistics—2020 update: A report from the American Heart Association. Circulation 2020, 141, e139–e596. [Google Scholar]
  13. Martins, B.; Ferreira, D.; Neto, C.; Abelha, A.; Machado, J. Data mining for cardiovascular disease prediction. J. Med. Syst. 2021, 45, 1–8. [Google Scholar] [CrossRef] [PubMed]
  14. Timmis, A.; Vardas, P.; Townsend, N.; Torbica, A.; Katus, H.; De Smedt, D.; Gale, C.P.; Maggioni, A.P.; Petersen, S.E.; Huculeci, R. European Society of Cardiology: Cardiovascular disease statistics 2021. Eur. Heart J. 2022, 43, 716–799. [Google Scholar] [PubMed]
  15. Zhao, D. Epidemiological features of cardiovascular disease in Asia. JACC Asia 2021, 1, 1–13. [Google Scholar] [CrossRef] [PubMed]
  16. Perets, A.; Baruch, Y.; Weisbuch, F.; Shoshany, G.; Neufeld, G.; Cohen, S. Enhancing the vascularization of three-dimensional porous alginate scaffolds by incorporating controlled release basic fibroblast growth factor microspheres. J. Biomed. Mater. Res. Part A: Off. J. Soc. Biomater. Jpn. Soc. Biomater. Aust. Soc. Biomater. Korean Soc. Biomater. 2003, 65, 489–497. [Google Scholar] [CrossRef] [PubMed]
  17. Levick, J.R. An Introduction to Cardiovascular Physiology; Butterworth-Heinemann: Oxford, UK, 2013. [Google Scholar]
  18. Traore, M.A.; George, S.C. Tissue engineering the vascular tree. Tissue Eng. Part B Rev. 2017, 23, 505–514. [Google Scholar] [CrossRef]
  19. Song, H.-H.G.; Rumma, R.T.; Ozaki, C.K.; Edelman, E.R.; Chen, C.S. Vascular tissue engineering: Progress, challenges, and clinical promise. Cell Stem Cell 2018, 22, 340–354. [Google Scholar]
  20. Szklanny, A.A.; Machour, M.; Redenski, I.; Chochola, V.; Goldfracht, I.; Kaplan, B.; Epshtein, M.; Simaan Yameen, H.; Merdler, U.; Feinberg, A. 3D Bioprinting of Engineered Tissue Flaps with Hierarchical Vessel Networks (VesselNet) for Direct Host-To-Implant Perfusion. Adv. Mater. 2021, 33, 2102661. [Google Scholar] [CrossRef]
  21. Ben-Shaul, S.; Landau, S.; Merdler, U.; Levenberg, S. Mature vessel networks in engineered tissue promote graft–host anastomosis and prevent graft thrombosis. Proc. Natl. Acad. Sci. USA 2019, 116, 2955–2960. [Google Scholar] [CrossRef]
  22. Rouwkema, J.; Khademhosseini, A. Vascularization and angiogenesis in tissue engineering: Beyond creating static networks. Trends Biotechnol. 2016, 34, 733–745. [Google Scholar] [CrossRef]
  23. Mandrycky, C.J.; Howard, C.C.; Rayner, S.G.; Shin, Y.J.; Zheng, Y. Organ-on-a-chip systems for vascular biology. J. Mol. Cell. Cardiol. 2021, 159, 1–13. [Google Scholar] [CrossRef] [PubMed]
  24. Levenberg, S.; Rouwkema, J.; Macdonald, M.; Garfein, E.S.; Kohane, D.S.; Darland, D.C.; Mar, R.; Van Blitterswijk, C.A.; Mulligan, R.C.; D’Amore, P.A.; et al. Engineering vascularized skeletal muscle tissue. Nat. Biotechnol. 2005, 23, 879. [Google Scholar] [CrossRef]
  25. Pashneh-Tala, S.; MacNeil, S.; Claeyssens, F. The tissue-engineered vascular graft—Past, present, and future. Tissue Eng. Part B Rev. 2016, 22, 68–100. [Google Scholar] [CrossRef] [PubMed]
  26. Bakey, M.; De Bakey, M. Successful resection of aneurysm of distal aortic arch and replacement by graft. J. Am. Med. Assoc. 1954, 155, 1398. [Google Scholar] [CrossRef] [PubMed]
  27. Walpoth, B.H.; Bowlin, G.L. The daunting quest for a small diameter vascular graft. Expert Rev. Med. Devices 2005, 2, 647–651. [Google Scholar] [CrossRef]
  28. Zdrahala, R.J. Small caliber vascular grafts. Part I: State of the art. J. Biomater. Appl. 1996, 10, 309–329. [Google Scholar] [CrossRef]
  29. Zdrahala, R.J. Small caliber vascular grafts. Part II: Polyurethanes revisited. J. Biomater. Appl. 1996, 11, 37–61. [Google Scholar] [CrossRef]
  30. Chlupáč, J.; Filova, E.; Bačáková, L. Blood vessel replacement: 50 years of development and tissue engineering paradigms in vascular surgery. Physiol. Res. 2009, 58, S119–S139. [Google Scholar] [CrossRef]
  31. Nevelsteen, A.; Wouters, L.; Suy, R. Aortofemoral Dacron reconstruction for aorto-iliac occlusive disease: A 25-year survey. Eur. J. Vasc. Surg. 1991, 5, 179–186. [Google Scholar] [CrossRef]
  32. Goldman, S.; Zadina, K.; Moritz, T.; Ovitt, T.; Sethi, G.; Copeland, J.G.; Thottapurathu, L.; Krasnicka, B.; Ellis, N.; Anderson, R.J. Long-term patency of saphenous vein and left internal mammary artery grafts after coronary artery bypass surgery: Results from a Department of Veterans Affairs Cooperative Study. J. Am. Coll. Cardiol. 2004, 44, 2149–2156. [Google Scholar] [CrossRef]
  33. Polo, J.R.; Ligero, J.M.; Diaz-Cartelle, J.; Garcia-Pajares, R.; Cervera, T.; Reparaz, L. Randomized comparison of 6-mm straight grafts versus 6-to 8-mm tapered grafts for brachial-axillary dialysis access. J. Vasc. Surg. 2004, 40, 319–324. [Google Scholar] [CrossRef] [PubMed]
  34. García-Pajares, R.; Polo, J.R.; Flores, n.; Gonzalez-Tabares, E.; Solís, J.V. Upper arm polytetrafluoroethylene grafts for dialysis access. Analysis of two different graft sizes: 6 mm and 6–8 mm. Vasc. Endovasc. Surg. 2003, 37, 335–343. [Google Scholar] [CrossRef] [PubMed]
  35. Filipe, E.C.; Santos, M.; Hung, J.; Lee, B.S.; Yang, N.; Chan, A.H.; Ng, M.K.; Rnjak-Kovacina, J.; Wise, S.G. Rapid endothelialization of off-the-shelf small diameter silk vascular grafts. JACC Basic Transl. Sci. 2018, 3, 38–53. [Google Scholar] [CrossRef] [PubMed]
  36. Bao, L.; Tang, J.; Hong, F.F.; Lu, X.; Chen, L. Physicochemical properties and in vitro biocompatibility of three bacterial nanocellulose conduits for blood vessel applications. Carbohydr. Polym. 2020, 239, 116246. [Google Scholar] [CrossRef] [PubMed]
  37. Dekker, A.; Reitsma, K.; Beugeling, T.; Bantjes, A.; Feijen, J.; Van Aken, W. Adhesion of endothelial cells and adsorption of serum proteins on gas plasma-treated polytetrafluoroethylene. Biomaterials 1991, 12, 130–138. [Google Scholar] [CrossRef]
  38. Gharamti, A.; Kanafani, Z.A. Vascular graft infections: An update. Infect. Dis. Clin. 2018, 32, 789–809. [Google Scholar] [CrossRef]
  39. Mi, H.-Y.; Jing, X.; Thomsom, J.A.; Turng, L.-S. Promoting endothelial cell affinity and antithrombogenicity of polytetrafluoroethylene (PTFE) by mussel-inspired modification and RGD/heparin grafting. J. Mater. Chem. B 2018, 6, 3475–3485. [Google Scholar] [CrossRef]
  40. Zhu, A.; Ming, Z.; Jian, S. Blood compatibility of chitosan/heparin complex surface modified ePTFE vascular graft. Appl. Surf. Sci. 2005, 241, 485–492. [Google Scholar] [CrossRef]
  41. Ku, S.H.; Park, C.B. Human endothelial cell growth on mussel-inspired nanofiber scaffold for vascular tissue engineering. Biomaterials 2010, 31, 9431–9437. [Google Scholar] [CrossRef]
  42. Jiang, Y.-C.; Wang, X.-F.; Xu, Y.-Y.; Qiao, Y.-H.; Guo, X.; Wang, D.-F.; Li, Q.; Turng, L.-S. Polycaprolactone nanofibers containing vascular endothelial growth factor-encapsulated gelatin particles enhance mesenchymal stem cell differentiation and angiogenesis of endothelial cells. Biomacromolecules 2018, 19, 3747–3753. [Google Scholar] [CrossRef] [PubMed]
  43. Lee, H.; Dellatore, S.M.; Miller, W.M.; Messersmith, P.B. Mussel-inspired surface chemistry for multifunctional coatings. Science 2007, 318, 426–430. [Google Scholar] [CrossRef] [PubMed]
  44. Li, L.-Y.; Cui, L.-Y.; Zeng, R.-C.; Li, S.-Q.; Chen, X.-B.; Zheng, Y.; Kannan, M.B. Advances in functionalized polymer coatings on biodegradable magnesium alloys—A review. Acta Biomater. 2018, 79, 23–36. [Google Scholar] [CrossRef] [PubMed]
  45. Hu, X.; Shen, H.; Yang, F.; Bei, J.; Wang, S. Preparation and cell affinity of microtubular orientation-structured PLGA (70/30) blood vessel scaffold. Biomaterials 2008, 29, 3128–3136. [Google Scholar] [CrossRef]
  46. Sell, S.A.; Wolfe, P.S.; Garg, K.; McCool, J.M.; Rodriguez, I.A.; Bowlin, G.L. The use of natural polymers in tissue engineering: A focus on electrospun extracellular matrix analogues. Polymers 2010, 2, 522–553. [Google Scholar] [CrossRef]
  47. L’Heureux, N.; Dusserre, N.; Konig, G.; Victor, B.; Keire, P.; Wight, T.N.; Chronos, N.A.; Kyles, A.E.; Gregory, C.R.; Hoyt, G. Human tissue-engineered blood vessels for adult arterial revascularization. Nat. Med. 2006, 12, 361–365. [Google Scholar] [CrossRef]
  48. Lee, Y.B.; Shin, Y.M.; Lee, J.-h.; Jun, I.; Kang, J.K.; Park, J.-C.; Shin, H. Polydopamine-mediated immobilization of multiple bioactive molecules for the development of functional vascular graft materials. Biomaterials 2012, 33, 8343–8352. [Google Scholar] [CrossRef]
  49. Freeman, J.; Chen, A.; Weinberg, R.J.; Okada, T.; Chen, C.; Lin, P.H. Sustained thromboresistant bioactivity with reduced intimal hyperplasia of heparin-bonded polytetrafluoroethylene propaten graft in a chronic canine femoral artery bypass model. Ann. Vasc. Surg. 2018, 49, 295–303. [Google Scholar] [CrossRef]
  50. Lumsden, A.B.; Morrissey, N.J.; Staffa, R.; Lindner, J.; Janousek, L.; Treska, V.; Stadler, P.; Moursi, M.; Storck, M.; Johansen, K. Randomized controlled trial comparing the safety and efficacy between the FUSION BIOLINE heparin-coated vascular graft and the standard expanded polytetrafluoroethylene graft for femoropopliteal bypass. J. Vasc. Surg. 2015, 61, 703–712.e701. [Google Scholar] [CrossRef]
  51. Aalders, G.J.; van Vroonhoven, T.J. Polytetrafluoroethylene versus human umbilical vein in above-knee femoropopliteal bypass: Six-year results of a randomized clinical trial. J. Vasc. Surg. 1992, 16, 816–824. [Google Scholar] [CrossRef]
  52. Rosenthal, D.; Evans, R.; McKinsey, J.; Seagraves, M.; Lamis, P.; Clark, M.; Daniel, W. Prosthetic above-knee femoropopliteal bypass for intermittent claudication. J. Cardiovasc. Surg. 1990, 31, 462–468. [Google Scholar]
  53. Moore, K.H.; Murphy, H.A.; George, E.M. The glycocalyx: A central regulator of vascular function. Am. J. Physiol. -Regul. Integr. Comp. Physiol. 2021, 320, R508–R518. [Google Scholar] [CrossRef] [PubMed]
  54. van Hinsbergh, V.W. The endothelium: Vascular control of haemostasis. Eur. J. Obstet. Gynecol. Reprod. Biol. 2001, 95, 198–201. [Google Scholar] [CrossRef] [PubMed]
  55. Jaffer, I.; Fredenburgh, J.; Hirsh, J.; Weitz, J. Medical device-induced thrombosis: What causes it and how can we prevent it? J. Thromb. Haemost. 2015, 13, S72–S81. [Google Scholar] [CrossRef] [PubMed]
  56. Moore, M.J.; Tan, R.P.; Yang, N.; Rnjak-Kovacina, J.; Wise, S.G. Bioengineering artificial blood vessels from natural materials. Trends Biotechnol. 2021, 40, 693–707. [Google Scholar] [CrossRef]
  57. Veith, F.J.; Gupta, S.K.; Ascer, E.; White-Flores, S.; Samson, R.H.; Scher, L.A.; Towne, J.B.; Bernhard, V.M.; Bonier, P.; Flinn, W.R. Six-year prospective multicenter randomized comparison of autologous saphenous vein and expanded polytetrafluoroethylene grafts in infrainguinal arterial reconstructions. J. Vasc. Surg. 1986, 3, 104–114. [Google Scholar] [CrossRef] [PubMed]
  58. Leal, B.B.; Wakabayashi, N.; Oyama, K.; Kamiya, H.; Braghirolli, D.I.; Pranke, P. Vascular tissue engineering: Polymers and methodologies for small caliber vascular grafts. Front. Cardiovasc. Med. 2021, 7, 592361. [Google Scholar] [CrossRef]
  59. Best, C.A.; Szafron, J.M.; Rocco, K.A.; Zbinden, J.; Dean, E.W.; Maxfield, M.W.; Kurobe, H.; Tara, S.; Bagi, P.S.; Udelsman, B.V. Differential outcomes of venous and arterial tissue engineered vascular grafts highlight the importance of coupling long-term implantation studies with computational modeling. Acta Biomater. 2019, 94, 183–194. [Google Scholar] [CrossRef]
  60. Seifu, D.G.; Purnama, A.; Mequanint, K.; Mantovani, D. Small-diameter vascular tissue engineering. Nat. Rev. Cardiol. 2013, 10, 410–421. [Google Scholar] [CrossRef]
  61. Qiu, Y.; Brown, A.C.; Myers, D.R.; Sakurai, Y.; Mannino, R.G.; Tran, R.; Ahn, B.; Hardy, E.T.; Kee, M.F.; Kumar, S. Platelet mechanosensing of substrate stiffness during clot formation mediates adhesion, spreading, and activation. Proc. Natl. Acad. Sci. USA 2014, 111, 14430–14435. [Google Scholar] [CrossRef]
  62. Milleret, V.; Hefti, T.; Hall, H.; Vogel, V.; Eberli, D. Influence of the fiber diameter and surface roughness of electrospun vascular grafts on blood activation. Acta Biomater. 2012, 8, 4349–4356. [Google Scholar] [CrossRef] [PubMed]
  63. Radke, D.; Jia, W.; Sharma, D.; Fena, K.; Wang, G.; Goldman, J.; Zhao, F. Tissue engineering at the blood-contacting surface: A review of challenges and strategies in vascular graft development. Adv. Healthc. Mater. 2018, 7, 1701461. [Google Scholar] [CrossRef] [PubMed]
  64. Wang, Z.; Mithieux, S.M.; Weiss, A.S. Fabrication techniques for vascular and vascularized tissue engineering. Adv. Healthc. Mater. 2019, 8, 1900742. [Google Scholar] [CrossRef]
  65. Herbert, S.P.; Stainier, D.Y. Molecular control of endothelial cell behaviour during blood vessel morphogenesis. Nat. Rev. Mol. Cell Biol. 2011, 12, 551–564. [Google Scholar] [CrossRef]
  66. Chatelain, L.; Louyriac, E.; Douair, I.; Lu, E.; Tuna, F.; Wooles, A.J.; Gardner, B.M.; Maron, L.; Liddle, S.T. Terminal uranium (V)-nitride hydrogenations involving direct addition or Frustrated Lewis Pair mechanisms. Nat. Commun. 2020, 11, 337. [Google Scholar] [CrossRef]
  67. Palikuqi, B.; Nguyen, D.-H.T.; Li, G.; Schreiner, R.; Pellegata, A.F.; Liu, Y.; Redmond, D.; Geng, F.; Lin, Y.; Gómez-Salinero, J.M. Adaptable haemodynamic endothelial cells for organogenesis and tumorigenesis. Nature 2020, 585, 426–432. [Google Scholar] [CrossRef]
  68. Landau, S.; Szklanny, A.A.; Yeo, G.C.; Shandalov, Y.; Kosobrodova, E.; Weiss, A.S.; Levenberg, S. Tropoelastin coated PLLA-PLGA scaffolds promote vascular network formation. Biomaterials 2017, 122, 72–82. [Google Scholar] [CrossRef] [PubMed]
  69. Gu, Y.; Zhang, J.; Zhang, X.; Liang, G.; Xu, T.; Niu, W. Three-dimensional printed Mg-doped β-TCP bone tissue engineering scaffolds: Effects of magnesium ion concentration on osteogenesis and angiogenesis in vitro. Tissue Eng. Regen. Med. 2019, 16, 415–429. [Google Scholar] [CrossRef]
  70. Wang, C.; Lai, J.; Li, K.; Zhu, S.; Lu, B.; Liu, J.; Tang, Y.; Wei, Y. Cryogenic 3D printing of dual-delivery scaffolds for improved bone regeneration with enhanced vascularization. Bioact. Mater. 2021, 6, 137–145. [Google Scholar] [CrossRef]
  71. Nguyen, D.-H.T.; Stapleton, S.C.; Yang, M.T.; Cha, S.S.; Choi, C.K.; Galie, P.A.; Chen, C.S. Biomimetic model to reconstitute angiogenic sprouting morphogenesis in vitro. Proc. Natl. Acad. Sci. USA 2013, 110, 6712–6717. [Google Scholar] [CrossRef]
  72. Landau, S.; Ben-Shaul, S.; Levenberg, S. Oscillatory strain promotes vessel stabilization and alignment through fibroblast YAP-mediated mechanosensitivity. Adv. Sci. 2018, 5, 1800506. [Google Scholar] [CrossRef]
  73. Kim, S.; Chung, M.; Ahn, J.; Lee, S.; Jeon, N.L. Interstitial flow regulates the angiogenic response and phenotype of endothelial cells in a 3D culture model. Lab Chip 2016, 16, 4189–4199. [Google Scholar] [CrossRef] [PubMed]
  74. Gui, L.; Niklason, L.E. Vascular tissue engineering: Building perfusable vasculature for implantation. Curr. Opin. Chem. Eng. 2014, 3, 68–74. [Google Scholar] [CrossRef]
  75. Emerson, D.R.; Cieślicki, K.; Gu, X.; Barber, R.W. Biomimetic design of microfluidic manifolds based on a generalised Murray’s law. Lab Chip 2006, 6, 447–454. [Google Scholar] [CrossRef]
  76. Czajka, C.A.; Calder, B.W.; Yost, M.J.; Drake, C.J. Implanted scaffold-free prevascularized constructs promote tissue repair. Ann. Plast. Surg. 2015, 74, 371. [Google Scholar] [CrossRef]
  77. Ercolani, E.; Del Gaudio, C.; Bianco, A. Vascular tissue engineering of small-diameter blood vessels: Reviewing the electrospinning approach. J. Tissue Eng. Regen. Med. 2015, 9, 861–888. [Google Scholar] [CrossRef]
  78. Braghirolli, D.; Helfer, V.; Chagastelles, P.; Dalberto, T.; Gamba, D.; Pranke, P. Electrospun scaffolds functionalized with heparin and vascular endothelial growth factor increase the proliferation of endothelial progenitor cells. Biomed. Mater. 2017, 12, 025003. [Google Scholar] [CrossRef]
  79. Mohan, T.; Nagaraj, C.; Nagy, B.M.; Bračicč, M.; Maver, U.; Olschewski, A.; Stana Kleinschek, K.; Kargl, R. Nano-and micropatterned polycaprolactone cellulose composite surfaces with tunable protein adsorption, fibrin clot formation, and endothelial cellular response. Biomacromolecules 2019, 20, 2327–2337. [Google Scholar] [CrossRef] [PubMed]
  80. Wang, N.; Zheng, W.; Cheng, S.; Zhang, W.; Liu, S.; Jiang, X. In vitro evaluation of essential mechanical properties and cell behaviors of a novel polylactic-co-glycolic acid (PLGA)-based tubular scaffold for small-diameter vascular tissue engineering. Polymers 2017, 9, 318. [Google Scholar] [CrossRef] [PubMed]
  81. Qiu, Y.L.; Chen, X.; Hou, Y.L.; Hou, Y.J.; Tian, S.B.; Chen, Y.H.; Yu, L.; Nie, M.H.; Liu, X.Q. Characterization of different biodegradable scaffolds in tissue engineering. Mol. Med. Rep. 2019, 19, 4043–4056. [Google Scholar] [CrossRef] [PubMed]
  82. Wang, D.; Xu, Y.; Li, Q.; Turng, L.-S. Artificial small-diameter blood vessels: Materials, fabrication, surface modification, mechanical properties, and bioactive functionalities. J. Mater. Chem. B 2020, 8, 1801–1822. [Google Scholar] [CrossRef] [PubMed]
  83. Hutmacher, D.W. Scaffold design and fabrication technologies for engineering tissues—State of the art and future perspectives. J. Biomater. Sci. Polym. Ed. 2001, 12, 107–124. [Google Scholar] [CrossRef] [PubMed]
  84. Joyce, K.; Fabra, G.T.; Bozkurt, Y.; Pandit, A. Bioactive potential of natural biomaterials: Identification, retention and assessment of biological properties. Signal Transduct. Target. Ther. 2021, 6, 1–28. [Google Scholar]
  85. Hess, F. History of (MICRO) vascular surgery and the development of small-caliber blood vessel prostheses (with some notes on patency rates and re-endothelialization). Microsurgery 1985, 6, 59–69. [Google Scholar] [CrossRef] [PubMed]
  86. Sullivan, S.J.; Maki, T.; Borland, K.M.; Mahoney, M.D.; Solomon, B.A.; Muller, T.E.; Monaco, A.P.; Chick, W.L. Biohybrid artificial pancreas: Long-term implantation studies in diabetic, pancreatectomized dogs. Science 1991, 252, 718–721. [Google Scholar] [CrossRef] [PubMed]
  87. Catto, V.; Farè, S.; Freddi, G.; Tanzi, M.C. Vascular tissue engineering: Recent advances in small diameter blood vessel regeneration. Int. Sch. Res. Not. 2014, 2014, 923030. [Google Scholar] [CrossRef]
  88. Geelhoed, W.J.; Moroni, L.; Rotmans, J.I. Utilizing the foreign body response to grow tissue engineered blood vessels in vivo. J. Cardiovasc. Transl. Res. 2017, 10, 167–179. [Google Scholar] [CrossRef]
  89. Kumar, V.A.; Brewster, L.P.; Caves, J.M.; Chaikof, E.L. Tissue engineering of blood vessels: Functional requirements, progress, and future challenges. Cardiovasc. Eng. Technol. 2011, 2, 137–148. [Google Scholar] [CrossRef]
  90. Gabriel, M.; Niederer, K.; Becker, M.; Raynaud, C.M.; Vahl, C.-F.; Frey, H. Tailoring novel PTFE surface properties: Promoting cell adhesion and antifouling properties via a wet chemical approach. Bioconjugate Chem. 2016, 27, 1216–1221. [Google Scholar] [CrossRef]
  91. Chen, L.; He, H.; Wang, M.; Li, X.; Yin, H. Surface coating of polytetrafluoroethylene with extracellular matrix and anti-CD34 antibodies facilitates endothelialization and inhibits platelet adhesion under sheer stress. Tissue Eng. Regen. Med. 2017, 14, 359–370. [Google Scholar] [CrossRef]
  92. Maurmann, N.; Sperling, L.-E.; Pranke, P. Electrospun and electrosprayed scaffolds for tissue engineering. Cut.-Edge Enabling Technol. Regen. Med. 2018, 1078, 79–100. [Google Scholar]
  93. Tysoe, O.C.; Justin, A.W.; Brevini, T.; Chen, S.E.; Mahbubani, K.T.; Frank, A.K.; Zedira, H.; Melum, E.; Saeb-Parsy, K.; Markaki, A.E. Isolation and propagation of primary human cholangiocyte organoids for the generation of bioengineered biliary tissue. Nat. Protoc. 2019, 14, 1884–1925. [Google Scholar] [CrossRef]
  94. Niklason, L.; Gao, J.; Abbott, W.; Hirschi, K.; Houser, S.; Marini, R.; Langer, R. Functional arteries grown in vitro. Science 1999, 284, 489–493. [Google Scholar] [CrossRef] [PubMed]
  95. O’brien, F.J. Biomaterials & scaffolds for tissue engineering. Mater. Today 2011, 14, 88–95. [Google Scholar]
  96. Fukunishi, T.; Best, C.A.; Sugiura, T.; Opfermann, J.; Ong, C.S.; Shinoka, T.; Breuer, C.K.; Krieger, A.; Johnson, J.; Hibino, N. Preclinical study of patient-specific cell-free nanofiber tissue-engineered vascular grafts using 3-dimensional printing in a sheep model. J. Thorac. Cardiovasc. Surg. 2017, 153, 924–932. [Google Scholar] [CrossRef]
  97. Melchiorri, A.J.; Hibino, N.; Brandes, Z.R.; Jonas, R.A.; Fisher, J.P. Development and assessment of a biodegradable solvent cast polyester fabric small-diameter vascular graft. J. Biomed. Mater. Res. Part A 2014, 102, 1972–1981. [Google Scholar] [CrossRef]
  98. Niklason, L.E.; Langer, R.S. Advances in tissue engineering of blood vessels and other tissues. Transpl. Immunol. 1997, 5, 303–306. [Google Scholar] [CrossRef]
  99. Niklason, L.E.; Abbott, W.; Gao, J.; Klagges, B.; Hirschi, K.K.; Ulubayram, K.; Conroy, N.; Jones, R.; Vasanawala, A.; Sanzgiri, S. Morphologic and mechanical characteristics of engineered bovine arteries. J. Vasc. Surg. 2001, 33, 628–638. [Google Scholar] [CrossRef]
  100. Dahl, S.L.; Kypson, A.P.; Lawson, J.H.; Blum, J.L.; Strader, J.T.; Li, Y.; Manson, R.J.; Tente, W.E.; DiBernardo, L.; Hensley, M.T. Readily available tissue-engineered vascular grafts. Sci. Transl. Med. 2011, 3, 68ra9. [Google Scholar] [CrossRef]
  101. Dahl, L.; Lawson, J.; Prichard, H.; Manson, R.; Tente, W.; Kypson, A. Abstracts from the emerging science series, April 24, 2013. Circulation 2013, 127, 2071–2072. [Google Scholar]
  102. Lawson, J.H.; Glickman, M.H.; Ilzecki, M.; Jakimowicz, T.; Jaroszynski, A.; Peden, E.K.; Pilgrim, A.J.; Prichard, H.L.; Guziewicz, M.; Przywara, S. Bioengineered human acellular vessels for dialysis access in patients with end-stage renal disease: Two phase 2 single-arm trials. Lancet 2016, 387, 2026–2034. [Google Scholar] [CrossRef]
  103. Lawson, J.H.; Glickman, M.H.; Jakimowicz, T.; Peden, E.K.; Szmidt, J.; Witkiewicz, W.; Zubilewicz, T.; Niklason, L.E. Bioengineered Human Acellular Vessels for Dialysis Access: Completed Phase 2 Studies. J. Am. Coll. Surg. 2017, 225, e52. [Google Scholar] [CrossRef]
  104. Lin, C.; Liu, C.; Zhang, L.; Huang, Z.; Zhao, P.; Chen, R.; Pang, M.; Chen, Z.; He, L.; Luo, C. Interaction of iPSC-derived neural stem cells on poly (L-lactic acid) nanofibrous scaffolds for possible use in neural tissue engineering. Int. J. Mol. Med. 2018, 41, 697–708. [Google Scholar] [CrossRef] [PubMed]
  105. Laurent, C.P.; Vaquette, C.; Liu, X.; Schmitt, J.-F.; Rahouadj, R. Suitability of a PLCL fibrous scaffold for soft tissue engineering applications: A combined biological and mechanical characterisation. J. Biomater. Appl. 2018, 32, 1276–1288. [Google Scholar] [CrossRef] [PubMed]
  106. Jang, B.S.; Cheon, J.Y.; Kim, S.H.; Park, W.H. Small diameter vascular graft with fibroblast cells and electrospun poly (L-lactide-co-ε-caprolactone) scaffolds: Cell Matrix Engineering. J. Biomater. Sci. Polym. Ed. 2018, 29, 942–959. [Google Scholar] [CrossRef] [PubMed]
  107. Caracciolo, P.C.; Rial-Hermida, M.I.; Montini-Ballarin, F.; Abraham, G.A.; Concheiro, A.; Alvarez-Lorenzo, C. Surface-modified bioresorbable electrospun scaffolds for improving hemocompatibility of vascular grafts. Mater. Sci. Eng. C 2017, 75, 1115–1127. [Google Scholar] [CrossRef]
  108. Hashi, C.K.; Zhu, Y.; Yang, G.-Y.; Young, W.L.; Hsiao, B.S.; Wang, K.; Chu, B.; Li, S. Antithrombogenic property of bone marrow mesenchymal stem cells in nanofibrous vascular grafts. Proc. Natl. Acad. Sci. USA 2007, 104, 11915–11920. [Google Scholar] [CrossRef]
  109. Bertram, U.; Steiner, D.; Poppitz, B.; Dippold, D.; Köhn, K.; Beier, J.P.; Detsch, R.; Boccaccini, A.R.; Schubert, D.W.; Horch, R.E. Vascular tissue engineering: Effects of integrating collagen into a PCL based nanofiber material. BioMed Res. Int. 2017, 2017, 9616939. [Google Scholar] [CrossRef]
  110. Pektok, E.; Nottelet, B.; Tille, J.-C.; Gurny, R.; Kalangos, A.; Moeller, M.; Walpoth, B.H. Degradation and healing characteristics of small-diameter poly (ε-caprolactone) vascular grafts in the rat systemic arterial circulation. Circulation 2008, 118, 2563–2570. [Google Scholar] [CrossRef]
  111. Gao, J.; Jiang, L.; Liang, Q.; Shi, J.; Hou, D.; Tang, D.; Chen, S.; Kong, D.; Wang, S. The grafts modified by heparinization and catalytic nitric oxide generation used for vascular implantation in rats. Regen. Biomater. 2018, 5, 105–114. [Google Scholar] [CrossRef]
  112. Koosehgol, S.; Ebrahimian-Hosseinabadi, M.; Alizadeh, M.; Zamanian, A. Preparation and characterization of in situ chitosan/polyethylene glycol fumarate/thymol hydrogel as an effective wound dressing. Mater. Sci. Eng. C 2017, 79, 66–75. [Google Scholar] [CrossRef] [PubMed]
  113. Nagiah, N.; Johnson, R.; Anderson, R.; Elliott, W.; Tan, W. Highly compliant vascular grafts with gelatin-sheathed coaxially structured nanofibers. Langmuir 2015, 31, 12993–13002. [Google Scholar] [CrossRef] [PubMed]
  114. Wang, Y.; Ameer, G.A.; Sheppard, B.J.; Langer, R. A tough biodegradable elastomer. Nat. Biotechnol. 2002, 20, 602–606. [Google Scholar] [CrossRef] [PubMed]
  115. Wu, W.; Allen, R.A.; Wang, Y. Fast-degrading elastomer enables rapid remodeling of a cell-free synthetic graft into a neoartery. Nat. Med. 2012, 18, 1148–1153. [Google Scholar] [CrossRef] [PubMed]
  116. Rai, R.; Tallawi, M.; Grigore, A.; Boccaccini, A.R. Synthesis, properties and biomedical applications of poly (glycerol sebacate) (PGS): A review. Prog. Polym. Sci. 2012, 37, 1051–1078. [Google Scholar] [CrossRef]
  117. Motlagh, D.; Yang, J.; Lui, K.Y.; Webb, A.R.; Ameer, G.A. Hemocompatibility evaluation of poly (glycerol-sebacate) in vitro for vascular tissue engineering. Biomaterials 2006, 27, 4315–4324. [Google Scholar] [CrossRef]
  118. Khosravi, R.; Best, C.A.; Allen, R.A.; Stowell, C.E.; Onwuka, E.; Zhuang, J.J.; Lee, Y.-U.; Yi, T.; Bersi, M.R.; Shinoka, T. Long-term functional efficacy of a novel electrospun poly (glycerol sebacate)-based arterial graft in mice. Ann. Biomed. Eng. 2016, 44, 2402–2416. [Google Scholar] [CrossRef]
  119. Alconcel, S.N.; Baas, A.S.; Maynard, H.D. FDA-approved poly (ethylene glycol)–protein conjugate drugs. Polym. Chem. 2011, 2, 1442–1448. [Google Scholar] [CrossRef]
  120. Hill-West, J.L.; Chowdhury, S.M.; Slepian, M.J.; Hubbell, J.A. Inhibition of thrombosis and intimal thickening by in situ photopolymerization of thin hydrogel barriers. Proc. Natl. Acad. Sci. USA 1994, 91, 5967–5971. [Google Scholar] [CrossRef]
  121. West, J.L.; Hubbell, J.A. Photopolymerized hydrogel materials for drug delivery applications. React. Polym. 1995, 25, 139–147. [Google Scholar] [CrossRef]
  122. Hahn, M.S.; McHale, M.K.; Wang, E.; Schmedlen, R.H.; West, J.L. Physiologic pulsatile flow bioreactor conditioning of poly (ethylene glycol)-based tissue engineered vascular grafts. Ann. Biomed. Eng. 2007, 35, 190–200. [Google Scholar] [CrossRef] [PubMed]
  123. Hou, J.; Zhang, X.; Wu, Y.; Jie, J.; Wang, Z.; Chen, G.-Q.; Sun, J.; Wu, L.-P. Amphiphilic and fatigue-resistant organohydrogels for small-diameter vascular grafts. Sci. Adv. 2022, 8, eabn5360. [Google Scholar] [CrossRef] [PubMed]
  124. Zou, Y.; Zhang, L.; Yang, L.; Zhu, F.; Ding, M.; Lin, F.; Wang, Z.; Li, Y. “Click” chemistry in polymeric scaffolds: Bioactive materials for tissue engineering. J. Control. Release 2018, 273, 160–179. [Google Scholar] [CrossRef] [PubMed]
  125. Devalliere, J.; Chen, Y.; Dooley, K.; Yarmush, M.L.; Uygun, B.E. Improving functional re-endothelialization of acellular liver scaffold using REDV cell-binding domain. Acta Biomater. 2018, 78, 151–164. [Google Scholar] [CrossRef]
  126. Liao, S.; He, Q.; Yang, L.; Liu, S.; Zhang, Z.; Guidoin, R.; Fu, Q.; Xie, X. Toward endothelialization via vascular endothelial growth factor immobilization on cell-repelling functional polyurethanes. J. Biomed. Mater. Res. Part B Appl. Biomater. 2019, 107, 965–977. [Google Scholar] [CrossRef]
  127. Zhang, H.; Jia, X.; Han, F.; Zhao, J.; Zhao, Y.; Fan, Y.; Yuan, X. Dual-delivery of VEGF and PDGF by double-layered electrospun membranes for blood vessel regeneration. Biomaterials 2013, 34, 2202–2212. [Google Scholar] [CrossRef]
  128. Li, Y.; Jiang, K.; Feng, J.; Liu, J.; Huang, R.; Chen, Z.; Yang, J.; Dai, Z.; Chen, Y.; Wang, N. Construction of Small-Diameter Vascular Graft by Shape-Memory and Self-Rolling Bacterial Cellulose Membrane. Adv. Healthc. Mater. 2017, 6, 1601343. [Google Scholar] [CrossRef]
  129. Huynh, T.; Abraham, G.; Murray, J.; Brockbank, K.; Hagen, P.-O.; Sullivan, S. Remodeling of an acellular collagen graft into a physiologically responsive neovessel. Nat. Biotechnol. 1999, 17, 1083–1086. [Google Scholar] [CrossRef]
  130. Damink, L.O.; Dijkstra, P.; Van Luyn, M.; Van Wachem, P.; Nieuwenhuis, P.; Feijen, J. In vitro degradation of dermal sheep collagen cross-linked using a water-soluble carbodiimide. Biomaterials 1996, 17, 679–684. [Google Scholar] [CrossRef]
  131. Li, X.; Xu, J.; Nicolescu, C.T.; Marinelli, J.T.; Tien, J. Generation, endothelialization, and microsurgical suture anastomosis of strong 1-mm-diameter collagen tubes. Tissue Eng. Part A 2017, 23, 335–344. [Google Scholar] [CrossRef]
  132. Zhang, F.; Xie, Y.; Celik, H.; Akkus, O.; Bernacki, S.H.; King, M.W. Engineering small-caliber vascular grafts from collagen filaments and nanofibers with comparable mechanical properties to native vessels. Biofabrication 2019, 11, 035020. [Google Scholar] [CrossRef] [PubMed]
  133. Elsayed, Y.; Lekakou, C.; Labeed, F.; Tomlins, P. Fabrication and characterisation of biomimetic, electrospun gelatin fibre scaffolds for tunica media-equivalent, tissue engineered vascular grafts. Mater. Sci. Eng. C 2016, 61, 473–483. [Google Scholar] [CrossRef] [PubMed]
  134. Zhao, Q.; Wang, J.; Cui, H.; Chen, H.; Wang, Y.; Du, X. Programmed shape-morphing scaffolds enabling facile 3D endothelialization. Adv. Funct. Mater. 2018, 28, 1801027. [Google Scholar] [CrossRef]
  135. Ying, G.; Jiang, N.; Yu, C.; Zhang, Y.S. Three-dimensional bioprinting of gelatin methacryloyl (GelMA). Bio-Des. Manuf. 2018, 1, 215–224. [Google Scholar] [CrossRef]
  136. Yue, K.; Trujillo-de Santiago, G.; Alvarez, M.M.; Tamayol, A.; Annabi, N.; Khademhosseini, A. Synthesis, properties, and biomedical applications of gelatin methacryloyl (GelMA) hydrogels. Biomaterials 2015, 73, 254–271. [Google Scholar] [CrossRef]
  137. Liang, Q.; Gao, F.; Zeng, Z.; Yang, J.; Wu, M.; Gao, C.; Cheng, D.; Pan, H.; Liu, W.; Ruan, C. Coaxial scale-up printing of diameter-tunable biohybrid hydrogel microtubes with high strength, perfusability, and endothelialization. Adv. Funct. Mater. 2020, 30, 2001485. [Google Scholar] [CrossRef]
  138. Gong, C.; Wang, H.; Liu, P.; Guo, T. Impact of intraoperative vascular occlusion during liver surgery on long-term outcomes: A systematic review and meta-analysis. Int. J. Surg. 2017, 44, 110–116. [Google Scholar] [CrossRef]
  139. Nguyen, T.-U.; Shojaee, M.; Bashur, C.A.; Kishore, V. Electrochemical fabrication of a biomimetic elastin-containing bi-layered scaffold for vascular tissue engineering. Biofabrication 2018, 11, 015007. [Google Scholar] [CrossRef]
  140. XIONG, D.-s. Synthesis and properties of physically crosslinked poly (vinyl alcohol) hydrogels. J. China Univ. Min. Technol. 2008, 18, 271–274. [Google Scholar]
  141. Li, D.Y.; Brooke, B.; Davis, E.C.; Mecham, R.P.; Sorensen, L.K.; Boak, B.B.; Eichwald, E.; Keating, M.T. Elastin is an essential determinant of arterial morphogenesis. Nature 1998, 393, 276–280. [Google Scholar] [CrossRef]
  142. Lee, S.; Sani, E.S.; Spencer, A.R.; Guan, Y.; Weiss, A.S.; Annabi, N. Human-recombinant-Elastin-based bioinks for 3D bioprinting of vascularized soft tissues. Adv. Mater. 2020, 32, 2003915. [Google Scholar] [CrossRef]
  143. Golińska, P. Biopolymer-based nanofilms: Utility and toxicity. In Biopolymer-Based Nano Films; Elsevier: Amsterdam, The Netherlands, 2021; pp. 353–385. [Google Scholar]
  144. Lovett, M.; Eng, G.; Kluge, J.; Cannizzaro, C.; Vunjak-Novakovic, G.; Kaplan, D.L. Tubular silk scaffolds for small diameter vascular grafts. Organogenesis 2010, 6, 217–224. [Google Scholar] [CrossRef]
  145. Catto, V.; Farè, S.; Cattaneo, I.; Figliuzzi, M.; Alessandrino, A.; Freddi, G.; Remuzzi, A.; Tanzi, M.C. Small diameter electrospun silk fibroin vascular grafts: Mechanical properties, in vitro biodegradability, and in vivo biocompatibility. Mater. Sci. Eng. C 2015, 54, 101–111. [Google Scholar] [CrossRef]
  146. Marcolin, C.; Draghi, L.; Tanzi, M.; Faré, S. Electrospun silk fibroin–gelatin composite tubular matrices as scaffolds for small diameter blood vessel regeneration. J. Mater. Sci. Mater. Med. 2017, 28, 1–12. [Google Scholar] [CrossRef] [PubMed]
  147. Marelli, B.; Achilli, M.; Alessandrino, A.; Freddi, G.; Tanzi, M.C.; Farè, S.; Mantovani, D. Collagen-reinforced electrospun silk fibroin tubular construct as small calibre vascular graft. Macromol. Biosci. 2012, 12, 1566–1574. [Google Scholar] [CrossRef]
  148. Liu, S.; Dong, C.; Lu, G.; Lu, Q.; Li, Z.; Kaplan, D.L.; Zhu, H. Bilayered vascular grafts based on silk proteins. Acta Biomater. 2013, 9, 8991–9003. [Google Scholar] [CrossRef]
  149. Cattaneo, I.; Figliuzzi, M.; Azzollini, N.; Catto, V.; Farè, S.; Tanzi, M.C.; Alessandrino, A.; Freddi, G.; Remuzzi, A. In vivo regeneration of elastic lamina on fibroin biodegradable vascular scaffold. Int. J. Artif. Organs 2013, 36, 166–174. [Google Scholar] [CrossRef]
  150. Enomoto, S.; Sumi, M.; Kajimoto, K.; Nakazawa, Y.; Takahashi, R.; Takabayashi, C.; Asakura, T.; Sata, M. Long-term patency of small-diameter vascular graft made from fibroin, a silk-based biodegradable material. J. Vasc. Surg. 2010, 51, 155–164. [Google Scholar] [CrossRef] [PubMed]
  151. Fowler, G.C. Pfenninger and Fowler’s Procedures for Primary Care E-Book; Elsevier Health Sciences: Amsterdam, The Netherlands, 2019. [Google Scholar]
  152. Winn, H.R. Youmans and Winn Neurological Surgery; Elsevier Health Sciences: Amsterdam, The Netherlands, 2022. [Google Scholar]
  153. Zhu, A.P.; Zhao, F.; Fang, N. Regulation of vascular smooth muscle cells on poly (ethylene terephthalate) film by O-carboxymethylchitosan surface immobilization. J. Biomed. Mater. Res. Part A: Off. J. Soc. Biomater. Jpn. Soc. Biomater. Aust. Soc. Biomater. Korean Soc. Biomater. 2008, 86, 467–476. [Google Scholar] [CrossRef] [PubMed]
  154. Vrana, N.E.; Liu, Y.; McGuinness, G.B.; Cahill, P.A. Characterization of poly (vinyl alcohol)/chitosan hydrogels as vascular tissue engineering scaffolds. In Macromolecular Symposia; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2008; pp. 106–110. [Google Scholar]
  155. Chupa, J.M.; Foster, A.M.; Sumner, S.R.; Madihally, S.V.; Matthew, H.W. Vascular cell responses to polysaccharide materials: In vitro and in vivo evaluations. Biomaterials 2000, 21, 2315–2322. [Google Scholar] [CrossRef]
  156. Kim, I.-Y.; Seo, S.-J.; Moon, H.-S.; Yoo, M.-K.; Park, I.-Y.; Kim, B.-C.; Cho, C.-S. Chitosan and its derivatives for tissue engineering applications. Biotechnol. Adv. 2008, 26, 1–21. [Google Scholar] [CrossRef]
  157. Zhu, C.; Fan, D.; Duan, Z.; Xue, W.; Shang, L.; Chen, F.; Luo, Y. Initial investigation of novel human-like collagen/chitosan scaffold for vascular tissue engineering. J. Biomed. Mater. Res. Part A Off. J. Soc. Biomater. Jpn. Soc. Biomater. Aust. Soc. Biomater. Korean Soc. Biomater. 2009, 89, 829–840. [Google Scholar] [CrossRef] [PubMed]
  158. Aussel, A.; Montembault, A.; Malaise, S.; Foulc, M.P.; Faure, W.; Cornet, S.; Aid, R.; Chaouat, M.; Delair, T.; Letourneur, D. In vitro mechanical property evaluation of chitosan-based hydrogels intended for vascular graft development. J. Cardiovasc. Transl. Res. 2017, 10, 480–488. [Google Scholar] [CrossRef] [PubMed]
  159. Yan, K.; Yang, C.; Zhong, W.; Lu, Z.; Li, X.; Shi, X.; Wang, D. Wire templated electrodeposition of vessel-like structured chitosan hydrogel by using a pulsed electrical signal. Soft Matter 2020, 16, 9471–9478. [Google Scholar] [CrossRef] [PubMed]
  160. Angra, V.; Sehgal, R.; Kaur, M.; Gupta, R. Commercialization of bionanocomposites. In Bionanocomposites in Tissue Engineering and Regenerative Medicine; Elsevier: Amsterdam, The Netherlands, 2021; pp. 587–610. [Google Scholar]
  161. Tahmasebifar, A.; Yilmaz, B.; Baran, E.T. Polysaccharide-based 3D bioprinter inks for tissue engineering. In Polysaccharide-Based Nanocomposites for Gene Delivery and Tissue Engineering; Elsevier: Amsterdam, The Netherlands, 2021; pp. 207–242. [Google Scholar]
  162. Gao, G.; Kim, H.; Kim, B.S.; Kong, J.S.; Lee, J.Y.; Park, B.W.; Chae, S.; Kim, J.; Ban, K.; Jang, J. Tissue-engineering of vascular grafts containing endothelium and smooth-muscle using triple-coaxial cell printing. Appl. Phys. Rev. 2019, 6, 041402. [Google Scholar] [CrossRef]
  163. Gao, G.; Lee, J.H.; Jang, J.; Lee, D.H.; Kong, J.S.; Kim, B.S.; Choi, Y.J.; Jang, W.B.; Hong, Y.J.; Kwon, S.M. Tissue engineered bio-blood-vessels constructed using a tissue-specific bioink and 3D coaxial cell printing technique: A novel therapy for ischemic disease. Adv. Funct. Mater. 2017, 27, 1700798. [Google Scholar] [CrossRef]
  164. Gao, G.; Park, J.Y.; Kim, B.S.; Jang, J.; Cho, D.W. Coaxial cell printing of freestanding, perfusable, and functional in vitro vascular models for recapitulation of native vascular endothelium pathophysiology. Adv. Healthc. Mater. 2018, 7, 1801102. [Google Scholar] [CrossRef]
  165. O’Neill, J.D.; Anfang, R.; Anandappa, A.; Costa, J.; Javidfar, J.; Wobma, H.M.; Singh, G.; Freytes, D.O.; Bacchetta, M.D.; Sonett, J.R. Decellularization of human and porcine lung tissues for pulmonary tissue engineering. Ann. Thorac. Surg. 2013, 96, 1046–1056. [Google Scholar] [CrossRef]
  166. Soffer-Tsur, N.; Shevach, M.; Shapira, A.; Peer, D.; Dvir, T. Optimizing the biofabrication process of omentum-based scaffolds for engineering autologous tissues. Biofabrication 2014, 6, 035023. [Google Scholar] [CrossRef]
  167. DeQuach, J.A.; Yuan, S.H.; Goldstein, L.S.; Christman, K.L. Decellularized porcine brain matrix for cell culture and tissue engineering scaffolds. Tissue Eng. Part A 2011, 17, 2583–2592. [Google Scholar] [CrossRef]
  168. Badylak, S.F.; Taylor, D.; Uygun, K. Whole-organ tissue engineering: Decellularization and recellularization of three-dimensional matrix scaffolds. Annu. Rev. Biomed. Eng. 2011, 13, 27–53. [Google Scholar] [CrossRef] [PubMed]
  169. Ayala, P.; Dai, E.; Hawes, M.; Liu, L.; Chaudhuri, O.; Haller, C.A.; Mooney, D.J.; Chaikof, E.L. Evaluation of a bioengineered construct for tissue engineering applications. J. Biomed. Mater. Res. Part B Appl. Biomater. 2018, 106, 2345–2354. [Google Scholar] [CrossRef] [PubMed]
  170. Conklin, B.; Richter, E.; Kreutziger, K.; Zhong, D.-S.; Chen, C. Development and evaluation of a novel decellularized vascular xenograft. Med. Eng. Phys. 2002, 24, 173–183. [Google Scholar] [CrossRef]
  171. Liu, R.H.; Ong, C.S.; Fukunishi, T.; Ong, K.; Hibino, N. Review of vascular graft studies in large animal models. Tissue Eng. Part B Rev. 2018, 24, 133–143. [Google Scholar] [CrossRef] [PubMed]
  172. Gui, L.; Muto, A.; Chan, S.A.; Breuer, C.K.; Niklason, L.E. Development of decellularized human umbilical arteries as small-diameter vascular grafts. Tissue Eng. Part A 2009, 15, 2665–2676. [Google Scholar] [CrossRef] [PubMed]
  173. Dimitrievska, S.; Cai, C.; Weyers, A.; Balestrini, J.L.; Lin, T.; Sundaram, S.; Hatachi, G.; Spiegel, D.A.; Kyriakides, T.R.; Miao, J. Click-coated, heparinized, decellularized vascular grafts. Acta Biomater. 2015, 13, 177–187. [Google Scholar] [CrossRef]
  174. Jiang, B.; Suen, R.; Wang, J.-J.; Zhang, Z.J.; Wertheim, J.A.; Ameer, G.A. Vascular scaffolds with enhanced antioxidant activity inhibit graft calcification. Biomaterials 2017, 144, 166–175. [Google Scholar] [CrossRef]
  175. Borschel, G.H.; Huang, Y.-C.; Calve, S.; Arruda, E.M.; Lynch, J.B.; Dow, D.E.; Kuzon, W.M.; Dennis, R.G.; Brown, D.L. Tissue engineering of recellularized small-diameter vascular grafts. Tissue Eng. 2005, 11, 778–786. [Google Scholar] [CrossRef] [PubMed]
  176. Tillman, B.W.; Yazdani, S.K.; Neff, L.P.; Corriere, M.A.; Christ, G.J.; Soker, S.; Atala, A.; Geary, R.L.; Yoo, J.J. Bioengineered vascular access maintains structural integrity in response to arteriovenous flow and repeated needle puncture. J. Vasc. Surg. 2012, 56, 783–793. [Google Scholar] [CrossRef]
  177. Kristofik, N.J.; Qin, L.; Calabro, N.E.; Dimitrievska, S.; Li, G.; Tellides, G.; Niklason, L.E.; Kyriakides, T.R. Improving in vivo outcomes of decellularized vascular grafts via incorporation of a novel extracellular matrix. Biomaterials 2017, 141, 63–73. [Google Scholar] [CrossRef]
  178. Syedain, Z.; Reimer, J.; Lahti, M.; Berry, J.; Johnson, S.; Bianco, R.; Tranquillo, R.T. Tissue engineering of acellular vascular grafts capable of somatic growth in young lambs. Nat. Commun. 2016, 7, 12951. [Google Scholar] [CrossRef]
  179. Sparks, C.H. Die-grown reinforced arterial grafts: Observations on long-term animal grafts and clinical experience. Ann. Surg. 1970, 172, 787. [Google Scholar] [CrossRef] [PubMed]
  180. Geelhoed, W.; van der Bogt, K.; Rothuizen, T.; Damanik, F.; Hamming, J.; Mota, C.; van Agen, M.; de Boer, H.; Restrepo, M.T.; Hinz, B. A novel method for engineering autologous non-thrombogenic in situ tissue-engineered blood vessels for arteriovenous grafting. Biomaterials 2020, 229, 119577. [Google Scholar] [CrossRef] [PubMed]
  181. Carrabba, M.; Madeddu, P. Current strategies for the manufacture of small size tissue engineering vascular grafts. Front. Bioeng. Biotechnol. 2018, 6, 41. [Google Scholar] [CrossRef] [PubMed]
  182. Moschouris, K.; Firoozi, N.; Kang, Y. The application of cell sheet engineering in the vascularization of tissue regeneration. Regen. Med. 2016, 11, 559–570. [Google Scholar] [CrossRef] [PubMed]
  183. Iwasaki, K.; Kojima, K.; Kodama, S.; Paz, A.C.; Chambers, M.; Umezu, M.; Vacanti, C.A. Bioengineered three-layered robust and elastic artery using hemodynamically-equivalent pulsatile bioreactor. Circulation 2008, 118, S52–S57. [Google Scholar] [CrossRef] [PubMed]
  184. L’heureux, N.; Pâquet, S.; Labbé, R.; Germain, L.; Auger, F.A. A completely biological tissue-engineered human blood vessel. FASEB J. 1998, 12, 47–56. [Google Scholar]
  185. McAllister, T.N.; Maruszewski, M.; Garrido, S.A.; Wystrychowski, W.; Dusserre, N.; Marini, A.; Zagalski, K.; Fiorillo, A.; Avila, H.; Manglano, X. Effectiveness of haemodialysis access with an autologous tissue-engineered vascular graft: A multicentre cohort study. Lancet 2009, 373, 1440–1446. [Google Scholar] [CrossRef]
  186. Yuan, B.; Jin, Y.; Sun, Y.; Wang, D.; Sun, J.; Wang, Z.; Zhang, W.; Jiang, X. A strategy for depositing different types of cells in three dimensions to mimic tubular structures in tissues. Adv. Mater. 2012, 24, 890–896. [Google Scholar] [CrossRef]
  187. Rayatpisheh, S.; Heath, D.E.; Shakouri, A.; Rujitanaroj, P.-O.; Chew, S.Y.; Chan-Park, M.B. Combining cell sheet technology and electrospun scaffolding for engineered tubular, aligned, and contractile blood vessels. Biomaterials 2014, 35, 2713–2719. [Google Scholar] [CrossRef]
  188. Xing, Q.; Qian, Z.; Tahtinen, M.; Yap, A.H.; Yates, K.; Zhao, F. Aligned nanofibrous cell-derived extracellular matrix for anisotropic vascular graft construction. Adv. Healthc. Mater. 2017, 6, 1601333. [Google Scholar] [CrossRef] [PubMed]
  189. Deepthi, S.; Sundaram, M.N.; Vijayan, P.; Nair, S.V.; Jayakumar, R. Engineering poly (hydroxy butyrate-co-hydroxy valerate) based vascular scaffolds to mimic native artery. Int. J. Biol. Macromol. 2018, 109, 85–98. [Google Scholar] [CrossRef] [PubMed]
  190. Kumar, V.A.; Caves, J.M.; Haller, C.A.; Dai, E.; Liu, L.; Grainger, S.; Chaikof, E.L. Acellular vascular grafts generated from collagen and elastin analogs. Acta Biomater. 2013, 9, 8067–8074. [Google Scholar] [CrossRef] [PubMed]
  191. Marga, F.; Jakab, K.; Khatiwala, C.; Shepherd, B.; Dorfman, S.; Hubbard, B.; Colbert, S.; Forgacs, G. Toward engineering functional organ modules by additive manufacturing. Biofabrication 2012, 4, 022001. [Google Scholar] [CrossRef]
  192. Kelm, J.M.; Lorber, V.; Snedeker, J.G.; Schmidt, D.; Broggini-Tenzer, A.; Weisstanner, M.; Odermatt, B.; Mol, A.; Zünd, G.; Hoerstrup, S.P. A novel concept for scaffold-free vessel tissue engineering: Self-assembly of microtissue building blocks. J. Biotechnol. 2010, 148, 46–55. [Google Scholar] [CrossRef]
  193. Mironov, V.; Visconti, R.P.; Kasyanov, V.; Forgacs, G.; Drake, C.J.; Markwald, R.R. Organ printing: Tissue spheroids as building blocks. Biomaterials 2009, 30, 2164–2174. [Google Scholar] [CrossRef]
  194. Norotte, C.; Marga, F.S.; Niklason, L.E.; Forgacs, G. Scaffold-free vascular tissue engineering using bioprinting. Biomaterials 2009, 30, 5910–5917. [Google Scholar] [CrossRef]
  195. Andrique, L.; Recher, G.; Alessandri, K.; Pujol, N.; Feyeux, M.; Bon, P.; Cognet, L.; Nassoy, P.; Bikfalvi, A. A model of guided cell self-organization for rapid and spontaneous formation of functional vessels. Sci. Adv. 2019, 5, eaau6562. [Google Scholar] [CrossRef]
  196. Pham, Q.P.; Sharma, U.; Mikos, A.G. Electrospinning of polymeric nanofibers for tissue engineering applications: A review. Tissue Eng. 2006, 12, 1197–1211. [Google Scholar] [CrossRef]
  197. Doshi, J.; Reneker, D.H. Electrospinning process and applications of electrospun fibers. J. Electrost. 1995, 35, 151–160. [Google Scholar] [CrossRef]
  198. Zargham, S.; Bazgir, S.; Tavakoli, A.; Rashidi, A.S.; Damerchely, R. The effect of flow rate on morphology and deposition area of electrospun nylon 6 nanofiber. J. Eng. Fibers Fabr. 2012, 7, 155892501200700414. [Google Scholar] [CrossRef]
  199. Mokhtari, F.; Latifi, M.; Shamshirsaz, M. Electrospinning/electrospray of polyvinylidene fluoride (PVDF): Piezoelectric nanofibers. J. Text. Inst. 2016, 107, 1037–1055. [Google Scholar] [CrossRef]
  200. Fridrikh, S.V.; Jian, H.Y.; Brenner, M.P.; Rutledge, G.C. Controlling the fiber diameter during electrospinning. Phys. Rev. Lett. 2003, 90, 144502. [Google Scholar] [CrossRef] [PubMed]
  201. Deitzel, J.M.; Kleinmeyer, J.D.; Hirvonen, J.K.; Tan, N.B. Controlled deposition of electrospun poly (ethylene oxide) fibers. Polymer 2001, 42, 8163–8170. [Google Scholar] [CrossRef]
  202. De Vrieze, S.; Van Camp, T.; Nelvig, A.; Hagström, B.; Westbroek, P.; De Clerck, K. The effect of temperature and humidity on electrospinning. J. Mater. Sci. 2009, 44, 1357–1362. [Google Scholar] [CrossRef]
  203. Heydarkhan-Hagvall, S.; Schenke-Layland, K.; Dhanasopon, A.P.; Rofail, F.; Smith, H.; Wu, B.M.; Shemin, R.; Beygui, R.E.; MacLellan, W.R. Three-dimensional electrospun ECM-based hybrid scaffolds for cardiovascular tissue engineering. Biomaterials 2008, 29, 2907–2914. [Google Scholar] [CrossRef]
  204. Wu, Y.; Qin, Y.; Wang, Z.; Wang, J.; Zhang, C.; Li, C.; Kong, D. The regeneration of macro-porous electrospun poly (ε-caprolactone) vascular graft during long-term in situ implantation. J. Biomed. Mater. Res. Part B Appl. Biomater. 2018, 106, 1618–1627. [Google Scholar] [CrossRef]
  205. Fiqrianti, I.A.; Widiyanti, P.; Manaf, M.A.; Savira, C.Y.; Cahyani, N.R.; Bella, F.R. Poly-L-lactic acid (PLLA)-chitosan-collagen electrospun tube for vascular graft application. J. Funct. Biomater. 2018, 9, 32. [Google Scholar] [CrossRef]
  206. Nezarati, R.M.; Eifert, M.B.; Dempsey, D.K.; Cosgriff-Hernandez, E. Electrospun vascular grafts with improved compliance matching to native vessels. J. Biomed. Mater. Res. Part B Appl. Biomater. 2015, 103, 313–323. [Google Scholar] [CrossRef]
  207. Hong, Y.; Ye, S.-H.; Nieponice, A.; Soletti, L.; Vorp, D.A.; Wagner, W.R. A small diameter, fibrous vascular conduit generated from a poly (ester urethane) urea and phospholipid polymer blend. Biomaterials 2009, 30, 2457–2467. [Google Scholar] [CrossRef]
  208. Han, J.; Lazarovici, P.; Pomerantz, C.; Chen, X.; Wei, Y.; Lelkes, P.I. Co-electrospun blends of PLGA, gelatin, and elastin as potential nonthrombogenic scaffolds for vascular tissue engineering. Biomacromolecules 2011, 12, 399–408. [Google Scholar] [CrossRef]
  209. Rocco, K.A.; Maxfield, M.W.; Best, C.A.; Dean, E.W.; Breuer, C.K. In vivo applications of electrospun tissue-engineered vascular grafts: A review. Tissue Eng. Part B Rev. 2014, 20, 628–640. [Google Scholar] [CrossRef] [PubMed]
  210. Lee, S.J.; Yoo, J.J.; Lim, G.J.; Atala, A.; Stitzel, J. In vitro evaluation of electrospun nanofiber scaffolds for vascular graft application. J. Biomed. Mater. Res. Part A Off. J. Soc. Biomater. Jpn. Soc. Biomater. Aust. Soc. Biomater. Korean Soc. Biomater. 2007, 83, 999–1008. [Google Scholar] [CrossRef] [PubMed]
  211. Park, S.; Kim, J.; Lee, M.-K.; Park, C.; Jung, H.-D.; Kim, H.-E.; Jang, T.-S. Fabrication of strong, bioactive vascular grafts with PCL/collagen and PCL/silica bilayers for small-diameter vascular applications. Mater. Des. 2019, 181, 108079. [Google Scholar] [CrossRef]
  212. Pan, Y.; Zhou, X.; Wei, Y.; Zhang, Q.; Wang, T.; Zhu, M.; Li, W.; Huang, R.; Liu, R.; Chen, J. Small-diameter hybrid vascular grafts composed of polycaprolactone and polydioxanone fibers. Sci. Rep. 2017, 7, 3615. [Google Scholar] [CrossRef]
  213. Ekaputra, A.K.; Prestwich, G.D.; Cool, S.M.; Hutmacher, D.W. Combining electrospun scaffolds with electrosprayed hydrogels leads to three-dimensional cellularization of hybrid constructs. Biomacromolecules 2008, 9, 2097–2103. [Google Scholar] [CrossRef]
  214. Wang, Y.; He, C.; Feng, Y.; Yang, Y.; Wei, Z.; Zhao, W.; Zhao, C. A chitosan modified asymmetric small-diameter vascular graft with anti-thrombotic and anti-bacterial functions for vascular tissue engineering. J. Mater. Chem. B 2020, 8, 568–577. [Google Scholar] [CrossRef]
  215. Wu, T.; Jiang, B.; Wang, Y.; Yin, A.; Huang, C.; Wang, S.; Mo, X. Electrospun poly (l-lactide-co-caprolactone)–collagen–chitosan vascular graft in a canine femoral artery model. J. Mater. Chem. B 2015, 3, 5760–5768. [Google Scholar] [CrossRef]
  216. Yin, A.; Zhang, K.; McClure, M.J.; Huang, C.; Wu, J.; Fang, J.; Mo, X.; Bowlin, G.L.; Al-Deyab, S.S.; El-Newehy, M. Electrospinning collagen/chitosan/poly (L-lactic acid-co-ϵ-caprolactone) to form a vascular graft: Mechanical and biological characterization. J. Biomed. Mater. Res. Part A 2013, 101, 1292–1301. [Google Scholar] [CrossRef]
  217. Torre-Muruzabal, A.; Daelemans, L.; Van Assche, G.; De Clerck, K.; Rahier, H. Creation of a nanovascular network by electrospun sacrificial nanofibers for self-healing applications and its effect on the flexural properties of the bulk material. Polym. Test. 2016, 54, 78–83. [Google Scholar] [CrossRef]
  218. Gualandi, C.; Zucchelli, A.; Fernández Osorio, M.; Belcari, J.; Focarete, M.L. Nanovascularization of polymer matrix: Generation of nanochannels and nanotubes by sacrificial electrospun fibers. Nano Lett. 2013, 13, 5385–5390. [Google Scholar] [CrossRef] [PubMed]
  219. Lee, J.B.; Wang, X.; Faley, S.; Baer, B.; Balikov, D.A.; Sung, H.J.; Bellan, L.M. Development of 3D microvascular networks within gelatin hydrogels using thermoresponsive sacrificial microfibers. Adv. Healthc. Mater. 2016, 5, 781–785. [Google Scholar] [CrossRef] [PubMed]
  220. Deng, Y.; Kuiper, J. Functional 3D Tissue Engineering Scaffolds; Elsevier: Amsterdam, The Netherlands, 2017. [Google Scholar]
  221. Karimi, A.; Navidbakhsh, M.; Shojaei, A.; Faghihi, S. Measurement of the uniaxial mechanical properties of healthy and atherosclerotic human coronary arteries. Mater. Sci. Eng. C 2013, 33, 2550–2554. [Google Scholar] [CrossRef] [PubMed]
  222. Seliktar, D.; Black, R.A.; Vito, R.P.; Nerem, R.M. Dynamic mechanical conditioning of collagen-gel blood vessel constructs induces remodeling in vitro. Ann. Biomed. Eng. 2000, 28, 351–362. [Google Scholar] [CrossRef] [PubMed]
  223. Aper, T.; Wilhelmi, M.; Gebhardt, C.; Hoeffler, K.; Benecke, N.; Hilfiker, A.; Haverich, A. Novel method for the generation of tissue-engineered vascular grafts based on a highly compacted fibrin matrix. Acta Biomater. 2016, 29, 21–32. [Google Scholar] [CrossRef] [PubMed]
  224. Lee, K.-W.; Stolz, D.B.; Wang, Y. Substantial expression of mature elastin in arterial constructs. Proc. Natl. Acad. Sci. USA 2011, 108, 2705–2710. [Google Scholar] [CrossRef]
  225. Nieponice, A.; Soletti, L.; Guan, J.; Hong, Y.; Gharaibeh, B.; Maul, T.M.; Huard, J.; Wagner, W.R.; Vorp, D.A. In vivo assessment of a tissue-engineered vascular graft combining a biodegradable elastomeric scaffold and muscle-derived stem cells in a rat model. Tissue Eng. Part A 2010, 16, 1215–1223. [Google Scholar] [CrossRef]
  226. Heath, D.E.; Kang, G.C.; Cao, Y.; Poon, Y.F.; Chan, V.; Chan-Park, M.B. Biomaterials patterned with discontinuous microwalls for vascular smooth muscle cell culture: Biodegradable small diameter vascular grafts and stable cell culture substrates. J. Biomater. Sci. Polym. Ed. 2016, 27, 1477–1494. [Google Scholar] [CrossRef]
  227. Chen, Y.Y.; Kingston, B.R.; Chan, W.C. Transcribing in vivo blood vessel networks into in vitro perfusable microfluidic devices. Adv. Mater. Technol. 2020, 5, 2000103. [Google Scholar] [CrossRef]
  228. Mironov, V.; Boland, T.; Trusk, T.; Forgacs, G.; Markwald, R.R. Organ printing: Computer-aided jet-based 3D tissue engineering. TRENDS Biotechnol. 2003, 21, 157–161. [Google Scholar] [CrossRef]
  229. Ozbolat, I.T.; Hospodiuk, M. Current advances and future perspectives in extrusion-based bioprinting. Biomaterials 2016, 76, 321–343. [Google Scholar] [CrossRef] [PubMed]
  230. Ramesh, S.; Harrysson, O.L.; Rao, P.K.; Tamayol, A.; Cormier, D.R.; Zhang, Y.; Rivero, I.V. Extrusion bioprinting: Recent progress, challenges, and future opportunities. Bioprinting 2021, 21, e00116. [Google Scholar] [CrossRef]
  231. Zein, I.; Hutmacher, D.W.; Tan, K.C.; Teoh, S.H. Fused deposition modeling of novel scaffold architectures for tissue engineering applications. Biomaterials 2002, 23, 1169–1185. [Google Scholar] [CrossRef] [PubMed]
  232. Camacho, P.; Busari, H.; Seims, K.B.; Schwarzenberg, P.; Dailey, H.L.; Chow, L.W. 3D printing with peptide–polymer conjugates for single-step fabrication of spatially functionalized scaffolds. Biomater. Sci. 2019, 7, 4237–4247. [Google Scholar] [CrossRef] [PubMed]
  233. Kabirian, F.; Ditkowski, B.; Zamanian, A.; Hoylaerts, M.F.; Mozafari, M.; Heying, R. Controlled NO-release from 3D-printed small-diameter vascular grafts prevents platelet activation and bacterial infectivity. ACS Biomater. Sci. Eng. 2019, 5, 2284–2296. [Google Scholar] [CrossRef] [PubMed]
  234. Duan, B.; Hockaday, L.A.; Kang, K.H.; Butcher, J.T. 3D bioprinting of heterogeneous aortic valve conduits with alginate/gelatin hydrogels. J. Biomed. Mater. Res. Part A 2013, 101, 1255–1264. [Google Scholar] [CrossRef]
  235. Liu, W.; Zhong, Z.; Hu, N.; Zhou, Y.; Maggio, L.; Miri, A.K.; Fragasso, A.; Jin, X.; Khademhosseini, A.; Zhang, Y.S. Coaxial extrusion bioprinting of 3D microfibrous constructs with cell-favorable gelatin methacryloyl microenvironments. Biofabrication 2018, 10, 024102. [Google Scholar] [CrossRef]
  236. Gao, Q.; He, Y.; Fu, J.-z.; Liu, A.; Ma, L. Coaxial nozzle-assisted 3D bioprinting with built-in microchannels for nutrients delivery. Biomaterials 2015, 61, 203–215. [Google Scholar] [CrossRef]
  237. Xu, L.; Varkey, M.; Jorgensen, A.; Ju, J.; Jin, Q.; Park, J.H.; Fu, Y.; Zhang, G.; Ke, D.; Zhao, W. Bioprinting small diameter blood vessel constructs with an endothelial and smooth muscle cell bilayer in a single step. Biofabrication 2020, 12, 045012. [Google Scholar] [CrossRef]
  238. Jakab, K.; Norotte, C.; Marga, F.; Murphy, K.; Vunjak-Novakovic, G.; Forgacs, G. Tissue engineering by self-assembly and bio-printing of living cells. Biofabrication 2010, 2, 022001. [Google Scholar] [CrossRef]
  239. Rider, P.; Kačarević, Ž.P.; Alkildani, S.; Retnasingh, S.; Barbeck, M. Bioprinting of tissue engineering scaffolds. J. Tissue Eng. 2018, 9, 2041731418802090. [Google Scholar] [CrossRef] [PubMed]
  240. Cui, H.; Nowicki, M.; Fisher, J.P.; Zhang, L.G. 3D bioprinting for organ regeneration. Adv. Healthc. Mater. 2017, 6, 1601118. [Google Scholar] [CrossRef] [PubMed]
  241. Christensen, K.; Xu, C.; Chai, W.; Zhang, Z.; Fu, J.; Huang, Y. Freeform inkjet printing of cellular structures with bifurcations. Biotechnol. Bioeng. 2015, 112, 1047–1055. [Google Scholar] [CrossRef] [PubMed]
  242. Xu, C.; Chai, W.; Huang, Y.; Markwald, R.R. Scaffold-free inkjet printing of three-dimensional zigzag cellular tubes. Biotechnol. Bioeng. 2012, 109, 3152–3160. [Google Scholar] [CrossRef]
  243. Duan, B. State-of-the-art review of 3D bioprinting for cardiovascular tissue engineering. Ann. Biomed. Eng. 2017, 45, 195–209. [Google Scholar] [CrossRef]
  244. Wang, Z.; Abdulla, R.; Parker, B.; Samanipour, R.; Ghosh, S.; Kim, K. A simple and high-resolution stereolithography-based 3D bioprinting system using visible light crosslinkable bioinks. Biofabrication 2015, 7, 045009. [Google Scholar] [CrossRef]
  245. Melchiorri, A.J.; Hibino, N.; Best, C.; Yi, T.; Lee, Y.; Kraynak, C.; Kimerer, L.K.; Krieger, A.; Kim, P.; Breuer, C.K. 3D-printed biodegradable polymeric vascular grafts. Adv. Healthc. Mater. 2016, 5, 319–325. [Google Scholar] [CrossRef] [PubMed]
  246. Derakhshanfar, S.; Mbeleck, R.; Xu, K.; Zhang, X.; Zhong, W.; Xing, M. 3D bioprinting for biomedical devices and tissue engineering: A review of recent trends and advances. Bioact. Mater. 2018, 3, 144–156. [Google Scholar] [CrossRef] [PubMed]
  247. Heintz, K.A.; Bregenzer, M.E.; Mantle, J.L.; Lee, K.H.; West, J.L.; Slater, J.H. Fabrication of 3D biomimetic microfluidic networks in hydrogels. Adv. Healthc. Mater. 2016, 5, 2153–2160. [Google Scholar] [CrossRef]
  248. Oujja, M.; Pérez, S.; Fadeeva, E.; Koch, J.; Chichkov, B.; Castillejo, M. Three dimensional microstructuring of biopolymers by femtosecond laser irradiation. Appl. Phys. Lett. 2009, 95, 263703. [Google Scholar] [CrossRef]
  249. Hribar, K.C.; Meggs, K.; Liu, J.; Zhu, W.; Qu, X.; Chen, S. Three-dimensional direct cell patterning in collagen hydrogels with near-infrared femtosecond laser. Sci. Rep. 2015, 5, 17203. [Google Scholar] [CrossRef] [PubMed]
  250. Applegate, M.B.; Coburn, J.; Partlow, B.P.; Moreau, J.E.; Mondia, J.P.; Marelli, B.; Kaplan, D.L.; Omenetto, F.G. Laser-based three-dimensional multiscale micropatterning of biocompatible hydrogels for customized tissue engineering scaffolds. Proc. Natl. Acad. Sci. USA 2015, 112, 12052–12057. [Google Scholar] [CrossRef] [PubMed]
  251. Brandenberg, N.; Lutolf, M.P. In situ patterning of microfluidic networks in 3D cell-laden hydrogels. Adv. Mater. 2016, 28, 7450–7456. [Google Scholar] [CrossRef] [PubMed]
  252. Berkovitch, Y.; Yelin, D.; Seliktar, D. Photo-patterning PEG-based hydrogels for neuronal engineering. Eur. Polym. J. 2015, 72, 473–483. [Google Scholar] [CrossRef]
  253. Arakawa, C.K.; Badeau, B.A.; Zheng, Y.; DeForest, C.A. Multicellular vascularized engineered tissues through user-programmable biomaterial photodegradation. Adv. Mater. 2017, 29, 1703156. [Google Scholar] [CrossRef] [PubMed]
  254. Giraud, M.-N.; Armbruster, C.; Carrel, T.; Tevaearai, H.T. Current state of the art in myocardial tissue engineering. Tissue Eng. 2007, 13, 1825–1836. [Google Scholar] [CrossRef] [PubMed]
  255. Muangsanit, P.; Shipley, R.J.; Phillips, J.B. Vascularization strategies for peripheral nerve tissue engineering. Anat. Rec. 2018, 301, 1657–1667. [Google Scholar] [CrossRef]
  256. Best, T.J.; Mackinnon, S.E. Peripheral nerve revascularization: A current literature review. J. Reconstr. Microsurg. 1994, 10, 193–204. [Google Scholar] [CrossRef]
  257. Appenzeller, O.; Dhital, K.K.; Cowen, T.; Burnstock, G. The nerves to blood vessels supplying blood to nerves: The innervation of vasa nervorum. Brain Res. 1984, 304, 383–386. [Google Scholar] [CrossRef]
  258. Smith, D.R.; Kobrine, A.I.; Rizzoli, H.V. Absence of autoregulation in peripheral nerve blood flow. J. Neurol. Sci. 1977, 33, 347–352. [Google Scholar] [CrossRef]
  259. Zochodne, D.W. Nerve and ganglion blood flow in diabetes: An appraisal. Int. Rev. Neurobiol. 2002, 50, 161–202. [Google Scholar] [PubMed]
  260. Low, P.A.; Lagerlund, T.D.; McManis, P.G. Nerve blood flow and oxygen delivery in normal, diabetic, and ischemic neuropathy. Int. Rev. Neurobiol. 1989, 31, 355–438. [Google Scholar] [PubMed]
  261. Kokaia, Z.; Lindvall, O. Neurogenesis after ischaemic brain insults. Curr. Opin. Neurobiol. 2003, 13, 127–132. [Google Scholar] [CrossRef] [PubMed]
  262. Leventhal, C.; Rafii, S.; Rafii, D.; Shahar, A.; Goldman, S.A. Endothelial trophic support of neuronal production and recruitment from the adult mammalian subependyma. Mol. Cell. Neurosci. 1999, 13, 450–464. [Google Scholar] [CrossRef]
  263. Isahara, K.; Yamamoto, M. The interaction of vascular endothelial cells and dorsal root ganglion neurites is mediated by vitronectin and heparan sulfate proteoglycans. Dev. Brain Res. 1995, 84, 164–178. [Google Scholar] [CrossRef]
  264. Strange, F.S.C. An operation for nerve pedicle grafting. Preliminary communication. Br. J. Surg. 1947, 34, 423–425. [Google Scholar] [CrossRef]
  265. Koshima, I.; Harii, K. Experimental study of vascularized nerve grafts: Multifactorial analyses of axonal regeneration of nerves transplanted into an acute burn wound. J. Hand Surg. 1985, 10, 64–72. [Google Scholar] [CrossRef]
  266. Kanaya, F.; Firrell, J.; Tsai, T.-M.; Breidenbach, W.C. Functional results of vascularized versus nonvascularized nerve grafting. Plast. Reconstr. Surg. 1992, 89, 924–930. [Google Scholar] [CrossRef]
  267. Ozcan, G.; Shenaq, S.; Spira, M. Vascularized nerve tube: An experimental alternative for vascularized nerve grafts over short gaps. J. Reconstr. Microsurg. 1993, 9, 405–413. [Google Scholar] [CrossRef]
  268. Iijima, Y.; Ajiki, T.; Murayama, A.; Takeshita, K. Effect of artificial nerve conduit vascularization on peripheral nerve in a necrotic bed. Plast. Reconstr. Surg. Glob. Open 2016, 4, e665. [Google Scholar] [CrossRef]
  269. Kakinoki, R.; Nishijima, N.; Ueba, Y.; Oka, M.; Yamamuro, T.; Nakamura, T. Nerve regeneration over a 25 mm gap in rat sciatic nerves using tubes containing blood vessels: The possibility of clinical application. Int. Orthop. 1997, 21, 332–336. [Google Scholar] [CrossRef] [PubMed]
  270. Grosso, A.; Burger, M.G.; Lunger, A.; Schaefer, D.J.; Banfi, A.; Di Maggio, N. It takes two to tango: Coupling of angiogenesis and osteogenesis for bone regeneration. Front. Bioeng. Biotechnol. 2017, 5, 68. [Google Scholar] [CrossRef]
  271. Byambaa, B.; Annabi, N.; Yue, K.; Trujillo-de Santiago, G.; Alvarez, M.M.; Jia, W.; Kazemzadeh-Narbat, M.; Shin, S.R.; Tamayol, A.; Khademhosseini, A. Bioprinted Osteogenic and Vasculogenic Patterns for Engineering 3D Bone Tissue. Adv. Healthc. Mater. 2017, 6, 1700015. [Google Scholar] [CrossRef] [PubMed]
  272. Song, X.; Liu, S.; Qu, X.; Hu, Y.; Zhang, X.; Wang, T.; Wei, F. BMP2 and VEGF promote angiogenesis but retard terminal differentiation of osteoblasts in bone regeneration by up-regulating Id1. Acta Biochim Biophys. Sin. 2011, 43, 796–804. [Google Scholar] [CrossRef] [PubMed]
  273. Mercado-Pagán, Á.E.; Stahl, A.M.; Shanjani, Y.; Yang, Y. Vascularization in bone tissue engineering constructs. Ann. Biomed. Eng. 2015, 43, 718–729. [Google Scholar] [CrossRef]
  274. Casanova, M.R.; Oliveira, C.; Fernandes, E.M.; Reis, R.L.; Silva, T.H.; Martins, A.; Neves, N.M. Spatial immobilization of endogenous growth factors to control vascularization in bone tissue engineering. Biomater. Sci. 2020, 8, 2577–2589. [Google Scholar] [CrossRef]
  275. Divband, B.; Samiei, M.; Davaran, S.; Roshangar, L.; Shahi, S.; Aghazadeh, M. Synthesis and in vitro Evaluation of Thermosensitive PLA-gP (HEM-co-NIPAAM) Hydrogel Used for Delivery of VEGF. Biointerface Res. Appl. Chem. 2020, 11, 8043–8051. [Google Scholar]
  276. Mitra, D.; Whitehead, J.; Yasui, O.W.; Leach, J.K. Bioreactor culture duration of engineered constructs influences bone formation by mesenchymal stem cells. Biomaterials 2017, 146, 29–39. [Google Scholar] [CrossRef]
  277. Xu, Z.; Xu, Y.; Basuthakur, P.; Patra, C.R.; Ramakrishna, S.; Liu, Y.; Thomas, V.; Nanda, H.S. Fibro-porous PLLA/gelatin composite membrane doped with cerium oxide nanoparticles as bioactive scaffolds for future angiogenesis. J. Mater. Chem. B 2020, 8, 9110–9120. [Google Scholar] [CrossRef]
  278. Ma, L.; Cheng, S.; Ji, X.; Zhou, Y.; Zhang, Y.; Li, Q.; Tan, C.; Peng, F.; Zhang, Y.; Huang, W. Immobilizing magnesium ions on 3D printed porous tantalum scaffolds with polydopamine for improved vascularization and osteogenesis. Mater. Sci. Eng. C 2020, 117, 111303. [Google Scholar] [CrossRef]
  279. Stähli, C.; James-Bhasin, M.; Hoppe, A.; Boccaccini, A.R.; Nazhat, S.N. Effect of ion release from Cu-doped 45S5 Bioglass® on 3D endothelial cell morphogenesis. Acta Biomater. 2015, 19, 15–22. [Google Scholar] [CrossRef] [PubMed]
  280. Wu, C.; Zhou, Y.; Fan, W.; Han, P.; Chang, J.; Yuen, J.; Zhang, M.; Xiao, Y. Hypoxia-mimicking mesoporous bioactive glass scaffolds with controllable cobalt ion release for bone tissue engineering. Biomaterials 2012, 33, 2076–2085. [Google Scholar] [CrossRef] [PubMed]
  281. Xue, Y.; Zhu, Z.; Zhang, X.; Chen, J.; Yang, X.; Gao, X.; Zhang, S.; Luo, F.; Wang, J.; Zhao, W. Accelerated bone regeneration by MOF modified multifunctional membranes through enhancement of osteogenic and angiogenic performance. Adv. Healthc. Mater. 2021, 10, 2001369. [Google Scholar] [CrossRef]
  282. Dong, Q.; Zhang, M.; Zhou, X.; Shao, Y.; Li, J.; Wang, L.; Chu, C.; Xue, F.; Yao, Q.; Bai, J. 3D-printed Mg-incorporated PCL-based scaffolds: A promising approach for bone healing. Mater. Sci. Eng. C 2021, 129, 112372. [Google Scholar] [CrossRef]
  283. Li, R.; Zhou, C.; Chen, J.; Luo, H.; Li, R.; Chen, D.; Zou, X.; Wang, W. Synergistic osteogenic and angiogenic effects of KP and QK peptides incorporated with an injectable and self-healing hydrogel for efficient bone regeneration. Bioact. Mater. 2022, 18, 267–283. [Google Scholar] [CrossRef] [PubMed]
  284. Liu, Y.; Yang, S.; Cao, L.; Zhang, X.; Wang, J.; Liu, C. Facilitated vascularization and enhanced bone regeneration by manipulation hierarchical pore structure of scaffolds. Mater. Sci. Eng. C 2020, 110, 110622. [Google Scholar] [CrossRef] [PubMed]
  285. Markou, M.; Kouroupis, D.; Badounas, F.; Katsouras, A.; Kyrkou, A.; Fotsis, T.; Murphy, C.; Bagli, E. Tissue engineering using vascular organoids from human pluripotent stem cell derived mural cell phenotypes. Front. Bioeng. Biotechnol. 2020, 8, 278. [Google Scholar] [CrossRef]
  286. Thrivikraman, G.; Athirasala, A.; Gordon, R.; Zhang, L.; Bergan, R.; Keene, D.R.; Jones, J.M.; Xie, H.; Chen, Z.; Tao, J. Rapid fabrication of vascularized and innervated cell-laden bone models with biomimetic intrafibrillar collagen mineralization. Nat. Commun. 2019, 10, 3520. [Google Scholar] [CrossRef]
  287. Zhou, Y.; Wu, C.; Chang, J. Bioceramics to regulate stem cells and their microenvironment for tissue regeneration. Mater. Today 2019, 24, 41–56. [Google Scholar] [CrossRef]
  288. Piard, C.; Jeyaram, A.; Liu, Y.; Caccamese, J.; Jay, S.M.; Chen, Y.; Fisher, J. 3D printed HUVECs/MSCs cocultures impact cellular interactions and angiogenesis depending on cell-cell distance. Biomaterials 2019, 222, 119423. [Google Scholar] [CrossRef]
  289. Tsao, C.W.; Aday, A.W.; Almarzooq, Z.I.; Alonso, A.; Beaton, A.Z.; Bittencourt, M.S.; Boehme, A.K.; Buxton, A.E.; Carson, A.P.; Commodore-Mensah, Y. Heart disease and stroke statistics—2022 update: A report from the American Heart Association. Circulation 2022, 145, e153–e639. [Google Scholar] [CrossRef] [PubMed]
  290. Luxán, G.; Dimmeler, S. The vasculature: A therapeutic target in heart failure? Cardiovasc. Res. 2022, 118, 53–64. [Google Scholar] [CrossRef] [PubMed]
  291. Brutsaert, D.L. Cardiac endothelial-myocardial signaling: Its role in cardiac growth, contractile performance, and rhythmicity. Physiol. Rev. 2003, 83, 59–115. [Google Scholar] [CrossRef] [PubMed]
  292. Gemberling, M.; Karra, R.; Dickson, A.L.; Poss, K.D. Nrg1 is an injury-induced cardiomyocyte mitogen for the endogenous heart regeneration program in zebrafish. Elife 2015, 4, e05871. [Google Scholar] [CrossRef]
  293. D’Uva, G.; Aharonov, A.; Lauriola, M.; Kain, D.; Yahalom-Ronen, Y.; Carvalho, S.; Weisinger, K.; Bassat, E.; Rajchman, D.; Yifa, O. ERBB2 triggers mammalian heart regeneration by promoting cardiomyocyte dedifferentiation and proliferation. Nat. Cell Biol. 2015, 17, 627–638. [Google Scholar] [CrossRef]
  294. Scimia, M.C.; Hurtado, C.; Ray, S.; Metzler, S.; Wei, K.; Wang, J.; Woods, C.E.; Purcell, N.H.; Catalucci, D.; Akasaka, T. APJ acts as a dual receptor in cardiac hypertrophy. Nature 2012, 488, 394–398. [Google Scholar] [CrossRef]
  295. Wang, W.; McKinnie, S.M.; Patel, V.B.; Haddad, G.; Wang, Z.; Zhabyeyev, P.; Das, S.K.; Basu, R.; McLean, B.; Kandalam, V. Loss of Apelin exacerbates myocardial infarction adverse remodeling and ischemia-reperfusion injury: Therapeutic potential of synthetic Apelin analogues. J. Am. Heart Assoc. 2013, 2, e000249. [Google Scholar] [CrossRef]
  296. Narmoneva, D.A.; Vukmirovic, R.; Davis, M.E.; Kamm, R.D.; Lee, R.T. Endothelial cells promote cardiac myocyte survival and spatial reorganization: Implications for cardiac regeneration. Circulation 2004, 110, 962–968. [Google Scholar] [CrossRef]
  297. Laschke, M.W.; Menger, M.D. Prevascularization in tissue engineering: Current concepts and future directions. Biotechnol. Adv. 2016, 34, 112–121. [Google Scholar] [CrossRef]
  298. Rodrigues, I.C.P.; Kaasi, A.; Maciel Filho, R.; Jardini, A.L.; Gabriel, L.P. Cardiac tissue engineering: Current state-of-the-art materials, cells and tissue formation. Einstein 2018, 16, eRB4538. [Google Scholar] [CrossRef]
  299. Nguyen, A.H.; Marsh, P.; Schmiess-Heine, L.; Burke, P.J.; Lee, A.; Lee, J.; Cao, H. Cardiac tissue engineering: State-of-the-art methods and outlook. J. Biol. Eng. 2019, 13, 1–21. [Google Scholar] [CrossRef] [PubMed]
  300. Pfisterer, L.; Korff, T. Spheroid-based in vitro angiogenesis model. In Angiogenesis Protocols; Springer: Berlin/Heidelberg, Germany, 2016; pp. 167–177. [Google Scholar]
  301. Basagiannis, D.; Zografou, S.; Murphy, C.; Fotsis, T.; Morbidelli, L.; Ziche, M.; Bleck, C.; Mercer, J.; Christoforidis, S. VEGF induces signalling and angiogenesis by directing VEGFR2 internalisation through macropinocytosis. J. Cell Sci. 2016, 129, 4091–4104. [Google Scholar] [CrossRef] [PubMed]
  302. Raica, M.; Cimpean, A.M. Platelet-derived growth factor (PDGF)/PDGF receptors (PDGFR) axis as target for antitumor and antiangiogenic therapy. Pharmaceuticals 2010, 3, 572–599. [Google Scholar] [CrossRef] [PubMed]
  303. Sun, H.-J.; Cai, W.-W.; Gong, L.-L.; Wang, X.; Zhu, X.-X.; Wan, M.-Y.; Wang, P.-Y.; Qiu, L.-Y. FGF-2-mediated FGFR1 signaling in human microvascular endothelial cells is activated by vaccarin to promote angiogenesis. Biomed. Pharmacother. 2017, 95, 144–152. [Google Scholar] [CrossRef]
  304. Patel, N.G.; Zhang, G. Responsive systems for cell sheet detachment. Organogenesis 2013, 9, 93–100. [Google Scholar] [CrossRef]
  305. Alghuwainem, A.; Alshareeda, A.T.; Alsowayan, B. Scaffold-free 3-D cell sheet technique bridges the gap between 2-D cell culture and animal models. Int. J. Mol. Sci. 2019, 20, 4926. [Google Scholar] [CrossRef]
  306. Hinton, T.J.; Jallerat, Q.; Palchesko, R.N.; Park, J.H.; Grodzicki, M.S.; Shue, H.-J.; Ramadan, M.H.; Hudson, A.R.; Feinberg, A.W. Three-dimensional printing of complex biological structures by freeform reversible embedding of suspended hydrogels. Sci. Adv. 2015, 1, e1500758. [Google Scholar] [CrossRef]
  307. Maiullari, F.; Costantini, M.; Milan, M.; Pace, V.; Chirivì, M.; Maiullari, S.; Rainer, A.; Baci, D.; Marei, H.E.-S.; Seliktar, D. A multi-cellular 3D bioprinting approach for vascularized heart tissue engineering based on HUVECs and iPSC-derived cardiomyocytes. Sci. Rep. 2018, 8, 1–15. [Google Scholar] [CrossRef]
Figure 1. Architecture of artery, vein, and capillary [6]. Copyright 2020, Wiley-VCH.
Figure 1. Architecture of artery, vein, and capillary [6]. Copyright 2020, Wiley-VCH.
Polymers 15 02015 g001
Figure 2. Preclinical study of cell-free 3D-printed nanofiber vascular grafts. (a) Study design of cell-free patient-specific nanofiber tissue-engineered vascular graft. (A,B) Acquiring the topographical information of the blood vessel from a preoperative angiogram in a sheep model. (C) Three-dimensional-printed mandrel designed with computer-aided design (CAD). (D) Electrospinning fibers around mandrel to fabricate nanofiber vascular grafts. (E) Nanofiber vascular graft. (F) Implantation of fabricated grafts. (G,H) SEM images of nanofibers. (I) Image of the implanted vascular graft. (b) Collagen and elastin deposition at 6 months postimplantation. Picrosirius red (A), Masson’s trichrome (B), Hart’s (C), and von Kossa (D) stainings of native inferior vena cava (native IVC) at 6 months. Picrosirius red (E), Masson’s trichrome (F), Hart’s (G), and von Kossa (H) stainings of tissue-engineered vascular graft (TEVG) at 6 months. Ref. [96] Copyright 2017, Elsevier.
Figure 2. Preclinical study of cell-free 3D-printed nanofiber vascular grafts. (a) Study design of cell-free patient-specific nanofiber tissue-engineered vascular graft. (A,B) Acquiring the topographical information of the blood vessel from a preoperative angiogram in a sheep model. (C) Three-dimensional-printed mandrel designed with computer-aided design (CAD). (D) Electrospinning fibers around mandrel to fabricate nanofiber vascular grafts. (E) Nanofiber vascular graft. (F) Implantation of fabricated grafts. (G,H) SEM images of nanofibers. (I) Image of the implanted vascular graft. (b) Collagen and elastin deposition at 6 months postimplantation. Picrosirius red (A), Masson’s trichrome (B), Hart’s (C), and von Kossa (D) stainings of native inferior vena cava (native IVC) at 6 months. Picrosirius red (E), Masson’s trichrome (F), Hart’s (G), and von Kossa (H) stainings of tissue-engineered vascular graft (TEVG) at 6 months. Ref. [96] Copyright 2017, Elsevier.
Polymers 15 02015 g002
Figure 3. In vitro and in vivo characterization of the PCL and heparin-modified PCL vascular grafts. (A,B) SEM images of PCL and heparin-modified PCL vascular grafts. (C) Optical images of scaffold before and after implantation. (D) Endothelialization (vWF) and smooth muscle regeneration (SMA) in the grafts after 1 month. (A1A3) native vascular, (B1B3) PCL implanted after 1 month, (D1D3) heparin-modified PCL implanted after 1 month. Graft lumen was indicated by *. Ref. [111] Copyright 2018, Oxford University Press.
Figure 3. In vitro and in vivo characterization of the PCL and heparin-modified PCL vascular grafts. (A,B) SEM images of PCL and heparin-modified PCL vascular grafts. (C) Optical images of scaffold before and after implantation. (D) Endothelialization (vWF) and smooth muscle regeneration (SMA) in the grafts after 1 month. (A1A3) native vascular, (B1B3) PCL implanted after 1 month, (D1D3) heparin-modified PCL implanted after 1 month. Graft lumen was indicated by *. Ref. [111] Copyright 2018, Oxford University Press.
Polymers 15 02015 g003
Figure 4. In vivo implantation of PEGDA-based vascular grafts. (A) Microscopic views after implantation and ultrasound images of native blood vessels (native), PHA/PEGDA grafts (PHA/PEGDA), and PHA/PEGDA grafts with uniformly dispersed graphdiyne (PHA/PEGDA-GDY). (B) HE and Masson staining results of implanted grafts. Ref. [123] Copyright 2022.
Figure 4. In vivo implantation of PEGDA-based vascular grafts. (A) Microscopic views after implantation and ultrasound images of native blood vessels (native), PHA/PEGDA grafts (PHA/PEGDA), and PHA/PEGDA grafts with uniformly dispersed graphdiyne (PHA/PEGDA-GDY). (B) HE and Masson staining results of implanted grafts. Ref. [123] Copyright 2022.
Polymers 15 02015 g004
Figure 5. GelMA-based vascular microtubes. (A) Coaxial extrusion printing of microtubes. (B) Biocompatibility and endothelialization of hydrogel microtubes. Ref. [137] Copyright 2020, Wiley-VCH.
Figure 5. GelMA-based vascular microtubes. (A) Coaxial extrusion printing of microtubes. (B) Biocompatibility and endothelialization of hydrogel microtubes. Ref. [137] Copyright 2020, Wiley-VCH.
Polymers 15 02015 g005
Figure 6. Cell-laden elastic constructs using modified human-recombinant elastin. (A) A schematic illustration of cell-laden 3D printing. (B) Immunostaining of the printed structure at 7 days after printing. (C) Scheme of vascularized tissue printing. (DF) Optical and fluorescence images of printed vascularized tissue. (G) Cell viability after printing. Ref. [142] Copyright 2020, Wiley-VCH.
Figure 6. Cell-laden elastic constructs using modified human-recombinant elastin. (A) A schematic illustration of cell-laden 3D printing. (B) Immunostaining of the printed structure at 7 days after printing. (C) Scheme of vascularized tissue printing. (DF) Optical and fluorescence images of printed vascularized tissue. (G) Cell viability after printing. Ref. [142] Copyright 2020, Wiley-VCH.
Polymers 15 02015 g006
Figure 7. Alginate-based vascular grafts. (A) Scheme of coaxial printing of EPCs and atorvastatin-laden blood vessels. (B) Optical images of coaxial needles and microscopic images of alginate tube, EPC-laden tube, and EPCs/atorvastatin-laden tube. (C) Endothelialization of the engineered vessels. (D) Immunostaining images of implanted tubes. Ref. [163] Copyright 2017, Wiley-VCH.
Figure 7. Alginate-based vascular grafts. (A) Scheme of coaxial printing of EPCs and atorvastatin-laden blood vessels. (B) Optical images of coaxial needles and microscopic images of alginate tube, EPC-laden tube, and EPCs/atorvastatin-laden tube. (C) Endothelialization of the engineered vessels. (D) Immunostaining images of implanted tubes. Ref. [163] Copyright 2017, Wiley-VCH.
Polymers 15 02015 g007
Figure 8. Strategies for fabrication of decellularized vascular grafts [64]. (A) Scheme for decellularized vascular grafts in large scale. (B) Scheme for decellularized vasculature grafts. Copyright 2019, Wiley-VCH.
Figure 8. Strategies for fabrication of decellularized vascular grafts [64]. (A) Scheme for decellularized vascular grafts in large scale. (B) Scheme for decellularized vasculature grafts. Copyright 2019, Wiley-VCH.
Polymers 15 02015 g008
Figure 9. Self-organization of ECs and SMCs into artificial mature blood vessels. (A,B) Confocal imaging of engineered vessels immunostainings. (C) Fluorescence intensity distribution of αSMA and CD31. (D) Immunostaining images of laminin, CD31, αSMA, and DAPI of self-assembled vascular grafts. Ref. [195] Copyright 2019.
Figure 9. Self-organization of ECs and SMCs into artificial mature blood vessels. (A,B) Confocal imaging of engineered vessels immunostainings. (C) Fluorescence intensity distribution of αSMA and CD31. (D) Immunostaining images of laminin, CD31, αSMA, and DAPI of self-assembled vascular grafts. Ref. [195] Copyright 2019.
Polymers 15 02015 g009
Figure 10. SEM images of hybrid PLGA–gelatin–elastin fibers (PGE fibers), immunofluorescence images of EC and SMC cultured on PGE fibers, and image of PGE fiber vascular grafts. Ref. [208] Copyright 2011, American Chemical Society.
Figure 10. SEM images of hybrid PLGA–gelatin–elastin fibers (PGE fibers), immunofluorescence images of EC and SMC cultured on PGE fibers, and image of PGE fiber vascular grafts. Ref. [208] Copyright 2011, American Chemical Society.
Polymers 15 02015 g010
Figure 11. Fabrication of customized blood vessel grafts using 3D printing technology. (a) Digital light processing (DLP) stereolithography. (b) Design and fabrication process of 3D-printed graft to treat a coarctation of the aorta. Copyright 2015, Wiley-VCH.
Figure 11. Fabrication of customized blood vessel grafts using 3D printing technology. (a) Digital light processing (DLP) stereolithography. (b) Design and fabrication process of 3D-printed graft to treat a coarctation of the aorta. Copyright 2015, Wiley-VCH.
Polymers 15 02015 g011
Figure 12. Three-dimensional biomimetic microfluidic networks in hydrogels fabricated with laser degradation. (AC) Transport between 3D intertwining microchannels. (D) Dense, tortuous vasculature engineered in PEGDA. (E) Microvasculature endothelialized with ECs. Ref. [247] Copyright 2016, Wiley-VCH.
Figure 12. Three-dimensional biomimetic microfluidic networks in hydrogels fabricated with laser degradation. (AC) Transport between 3D intertwining microchannels. (D) Dense, tortuous vasculature engineered in PEGDA. (E) Microvasculature endothelialized with ECs. Ref. [247] Copyright 2016, Wiley-VCH.
Polymers 15 02015 g012
Figure 13. Cardiomyocytes supported with hierarchical vascular constructs. (A) Immunostaining of printed cardiomyocytes. (B) Spontaneous beat and electromechanical communication observed through calcium transients traveling distances over 1 mm. (C) Scheme of vascularized cardiomyocyte constructs and confocal image of vascular networks and cardiomyocytes. (i) schematic of vascularized cardiomyocytes constructs. (ii) immunofluorescence image of vascular network (CD31, green) and cardiomyocytes (cTnl, red). Scale bar = 100 μm. Ref. [20] Copyright 2021, Wiley-VCH.
Figure 13. Cardiomyocytes supported with hierarchical vascular constructs. (A) Immunostaining of printed cardiomyocytes. (B) Spontaneous beat and electromechanical communication observed through calcium transients traveling distances over 1 mm. (C) Scheme of vascularized cardiomyocyte constructs and confocal image of vascular networks and cardiomyocytes. (i) schematic of vascularized cardiomyocytes constructs. (ii) immunofluorescence image of vascular network (CD31, green) and cardiomyocytes (cTnl, red). Scale bar = 100 μm. Ref. [20] Copyright 2021, Wiley-VCH.
Polymers 15 02015 g013
Table 1. Vascularized tissue engineering.
Table 1. Vascularized tissue engineering.
TissueStrategyGeneral FunctionsExamples of StudiesReferences
Peripheral nerveAutologous graftsEnhanced regeneration in nerve length, size, and sciatic function index (SFI)Ulnar nerve with its blood supply[264–266]
Preimplantment of nerve grafts to obtain microcirculation networkEnhanced axon diameter, myelin thickness, and the number of myelinated axonsVascularized amnion tube obtained through embedding between the femoral artery and vein[267,268]
Blood-vessel-including tabulationRegeneration across the nerve gapNative blood vessel wrapped with the nerve conduit[269]
BoneIon dopingPromoted angiogenesis through HIF-1α/VEGF signaling pathwayMg-doped tantalum scaffold[277–282]
Adding growth factorsImproved neovascularization and synergistic osteogenesisGelMA-based hydrogel with angiogenic peptide QK[283]
Altering topographyPromoted proliferation and differentiation of osteoblasts and ECs; increased bone–matrix interface strength and stimulate mineralizationPore-structured scaffolds based on poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx)[284]
Coculturing of cellsEnhanced capillary and bone formationBMSCs/ECs coculture[285–288]
HeartCoculturing of cellsPromoted vasculature formationEndothelial cells cultured with cardiomyocytes[297–300]
Adding growth factorsEnhanced vascularizationVEGF[301–303]
Microvascular tubes fabricated with cell sheetFormation of well-perfused microchannelsCoculturing ECs with cardiac cell sheets in a collagen hydrogel[304,305]
Prevascularization through 3D printingIntegration to the host’s vasculatureA hierarchical vasculature supporting in vitro functionality of cardiomyocytes[20,306,307]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, J.; Zhang, D.; Wu, L.-P.; Zhao, M. Current Strategies for Engineered Vascular Grafts and Vascularized Tissue Engineering. Polymers 2023, 15, 2015. https://doi.org/10.3390/polym15092015

AMA Style

Chen J, Zhang D, Wu L-P, Zhao M. Current Strategies for Engineered Vascular Grafts and Vascularized Tissue Engineering. Polymers. 2023; 15(9):2015. https://doi.org/10.3390/polym15092015

Chicago/Turabian Style

Chen, Jun, Di Zhang, Lin-Ping Wu, and Ming Zhao. 2023. "Current Strategies for Engineered Vascular Grafts and Vascularized Tissue Engineering" Polymers 15, no. 9: 2015. https://doi.org/10.3390/polym15092015

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop