Open access

Biomaterials for Cardiac Tissue Engineering

Written By

M. Arnal-Pastor, J. C. Chachques, M. Monleón Pradas and A. Vallés- Lluch

Submitted: 15 May 2012 Published: 22 May 2013

DOI: 10.5772/56076

From the Edited Volume

Regenerative Medicine and Tissue Engineering

Edited by Jose A. Andrades

Chapter metrics overview

4,484 Chapter Downloads

View Full Metrics

1. Introduction

1.1. Cardiovascular diseases

Cardiovascular diseases (CVD) are a leading death cause in developed countries (1 of every 3 deaths in the United States in 2008) [1]. Changes in diet and habits are causing CVD to become major mortality pathologies in developing countries too [2] (they are already responsible for a 30% of the world deaths). This group of diseases constitutes a great burden for the national health systems, consuming great percentages of the health systems budgets. In the particular case of the coronary heart diseases (CHD), 3,8 million men and 3,4 million women die a year worldwide because of them [3]. In the United States 1 of every 6 deaths in 2008 was caused by CHD [1].

The heart is a complex organ that pumps 7000 liters of blood to all the tissues in the body per day [4]. This pumping function precisely determines its anatomy. Heart tissue basically is formed by cardiac myocytes (contractile elements) [5], smooth muscle cells, fibroblasts, blood vessels, nerves and the extracellular matrix components (cardiac interstitium and collagen) [6] organized in a very particular way. Myocytes form muscular fibers with changing orientation across the ventricular wall up to 180º [7]. At the same time, muscular fibers are organized into myocardial laminas 4-6 myocytes thick separated from neighboring laminas by extracellular collagen [8]. The particular arrangement of the ventricular myocytes influence the mechanical and electrical function of the heart and small changes in it can lead to severe changes in these functions [9].

The extracellular matrix (ECM) connects the cells into a 3D architecture allowing the coupling of the forces produced by the myocytes. The anatomical model proposed by Torrent-Guasp [8], which considers the heart one muscle band plied in a double helical loop, explains how the ventricles contract and get an efficient pumping in every heart beat, achieving an ejection fraction of the 60% when sarcomeres individually contract 15% only [10].

Myocytes are intimately connected, forming a functional syncytium [8]. Each myocardial cell is coupled in average to 9,1 ± 2,2 [11] myocytes, by 99 [12] gap junctions where the transfer of ionic currents takes place. Gap junctions are a specialized form of cell connection; they are formed by a cluster of ionic channels essential to the rapid propagation of the action potential. The action potential is the electrical impulse responsible for the contraction of the cells [13]. A proper electrical coupling of the cells is critical to avoid arrhythmias and reentries and essential for the contraction to spread as a wave front.

Acute myocardial infarction (AMI) occurs when a coronary artery is clogged, in 80% of the cases, by coronary atherosclerosis with superimposed luminal thrombus [14]. This occlusion leaves the downstream zone of the heart without blood supply, what means lack of oxygen, nutrients and metabolites wash for the affected zone. As a consequence, the aerobic metabolism changes to anaerobic glycolysis [14], leading to a decrease in the pH and reduction in the contractile function. Within 20 to 40 minutes without blood supply cells start to die and as times passes more myocardial tissue is compromised. There is also a zone of the heart affected by the infarction, where myocytes remain viable but lower their activity to reduce the metabolism and oxygen consumption to survive under hypoxic conditions; they can recover their contractibility after revascularization [15].

Clinical practices aim to limit the severity and extension of the AMI by rapidly restoring the blood flow (reperfusion), alleviating the oxygen demand [16] and reducing reperfusion injury. This can be done with different treatments or combinations of them. Pharmacological approaches involve the use of anticoagulant therapies and thrombolytic drugs to eliminate the clot. Vasodilatators like nitrates are also used to favor the dilation of the vessels, aspirin to avoid platelet aggregation, betabloqueants to reduce the heart pace, as well as morphine to reduce the pain are employed. Another group of therapies are the percutaneous coronary interventions; they physically reopen the vessel via catheterization. There are different techniques: the regular angioplasty uses a catheter with a balloon that is inflated in the place of the thrombus to reopen the lumen [17], or allows the permanent implantation of a stent in the vessel to keep it open. There is a wide variety of these devices depending on their composition, whether they release drugs or are biodegradable or not, etc [18, 19].

These therapies restore the blood flow to the infarcted zone; but reperfusion therapy is not exempt of risks: it is a complex process that can induce apoptosis by the microenvironmental changes that the recovery of the blood supply induces (formation of free radicals, calcium release, neutrophils, etc.) [20]. So it has to be done carefully and there is always a compromise between limiting the infarction extension due to the time without oxygen and the induced apoptosis due to the reperfusion. Reperfusion done soon after the onset of the ischemia is very advantageous, saving more tissue by restoring the blood flow than the tissue that will be lost because of the toxic substances released in the reperfusion. All the aforementioned treatments basically limit the damage of the acute episode but do not regenerate the damaged tissue and do not avoid the subsequent ventricular remodeling following an AMI.

In the infarcted area there is a great number of dead myocytes, and the host response to the injury consists in activating the inflammatory response and producing cytokines [21]. Thereupon neutrophils, monocytes and macrophages migrate into this area to remove the necrotic tissue [22]. Then, matrix metalloproteases (MMPs) are activated, which have a deleterious effect on the collagen matrix of the heart and in the surrounding coronary vasculature by degrading them [23]. The weakening of the collagen leads to wall thinning and ventricular dilation, as well as mural realignment of myocytes bundles [24]. After the inflammatory phase and the resorption of the necrotic tissue, there is an increase in the deposition of cross-linked collagen in the infarcted area that leads to scar tissue formation. During the remodelling process a change in the collagen composition occurs, the type I collagen fraction is reduced from 80% to 40% and the collagen III is increased [25].

Against what it was thought, this scar is a living tissue with a fibroblast-like cell population nourished by a neovasculature; these cells regulate the collagen turnover of the scar tissue [22]. The scar tissue has a reduced or absent contractility as compared with the original healthy myocardium [26], what leads to a reduction in the overall cardiac function [27].

The remodeling process initially is a compensatory mechanism to overcome the loss of contractile tissue. But with time this adaptative process of overload becomes maladaptative [15]. To compensate the additional effort, the remaining beating tissue hypertrophies trying to overcome the reduction in the cardiac function. This overload leads to myocyte slippage and fibrotic interstitial growth and to a degenerating process that may end in heart failure. The heart remodeling produces in the ventricles a set of anatomical and functional changes, including increased wall stress, slimming of the wall, chamber dilation, increase of the sphericity, and a significant loss of cardiac function.

The ventricular shape change from elliptical to spherical reduces its ejection fraction, because of a change in the apical loop fiber orientation [28]. Another problem caused by the shape change is that the papillary muscles are separated, what leads to regurgitation, contributing to the overload of the heart [24]. Besides, remodeled hearts are more prone to suffer arrhythmias as the membrane potential is altered and because of the interstitial fibrotic growth that may affect conductivity [15].

The end stage of the degeneration is the heart failure, when the heart is unable to pump enough blood to match the metabolic needs of the tissues. Current treatments aim to avoid reaching this point. Pharmacological treatments aspire to reduce the work load and to protect the cardiac tissues from the accumulated harmful substances [29]. Surgical therapy involves different techniques with different objectives: to restore a proper blood flow in areas that lack it (by-pass surgery), to restore the normal elliptical geometry (Dor and Batista procedures), to restore the wall stress to normal (Dynamic Cardiomyoplasty), to limit the pathologic dilation, etc [10].

1.2. Cell therapy and cardiac tissue engineering

For many years, the heart has been considered a fully differentiated organ, with no myocyte regeneration after birth [30]. Recently it has been proved that myocytes have a limited regenerative capacity, around 1% of the cells per year at the age of 20 and it is reduced to 0,3% at the age of 75 [31]. This regenerative capacity is achieved thanks to a small population of cardiac stem cells [32]. Nevertheless, their regenerative capacity is limited and in any case it is not enough to regenerate the heart if it suffers severe damage, like the one provoked by a myocardial infarction. New therapies under development like cell therapy or tissue engineering, aim to boost this limited regenerative potential of the native tissue by employing cells, drugs, factors or patches.

The aim of cardiac cell therapy is to heal the damaged infarcted tissue by the implantation of cells into or onto the pathologic myocardium by different techniques (figure 1 a). In tissue engineering strategies, different types of cells have been combined with materials and with bioactive molecules if necessary to again try to recover the injured tissue. The employed materials will support cells, provide them 3D organization, protect them, stimulate and guide its growth, maintain them in the site of interest, etc.; in sum, they will act as an artificial extracellular matrix during the regeneration process. But the use of materials either injectable, or ex vivo conformed (gels –patches- or scaffolds) (figure 1 b) has an additional and important effect: the implantation of a material in the scarred ventricular wall, increases its thickness and by Laplace’s law, this increase leads to a reduction in the wall stress. This side-effect could be by itself very positive, even although regeneration did not arrive to happen, to limit ventricular remodeling and improve the quality of life of cardiac patients [29].

Figure 1.

(a) Classical cell therapy in the heart (freely inspired in Strauer BE, Kornowski R, Circulation 2003; 107: 929-934). (b) Tissue engineering approaches with cell sheets, scaffolds or injectable materials (freely inspired in Masuda S et al, Adv. Drug Del. Revs 2008; 60(2): 277-85). (c) Ventricular restrain device.

Advertisement

2. Cardiomyoplasty

2.1. Need for cell cardiomyoplasty

Cardiomyoplasty has evolved from “dynamic” to “cellular cardiomyoplasty”. The term dynamic cardiomyoplasty is referred to a surgical procedure developed in 1987 [33] to wrap the heart with the latissimus dorsi muscle, aiming to support the heart beating and limit the remodeling. Nevertheless, the obtained results were not as good as expected. With the advances in cell therapy, cellular cardiomyoplasty appeared as a promising therapeutical approach. This name encloses the therapies that use the injection of cells, from different origins, directly into the heart to try to obtain an improvement in the reduced heart function after an ischemic insult (figure 1 a).

The injected cells are envisaged to induce angiogenesis, inhibit apoptosis, help to recover hibernating myocardium, activate endogenous repair mechanisms, and create new contractile tissue that will replace the damaged one. Also they are expected to reverse the remodeling process that provoked ventricular dilation [34]. Many cells have been employed and the initial promising results obtained in animal models made this technique moved very fast to clinical trials, even if the mechanisms involved in the observed improvements were unknown. Unfortunately, the results obtained from the clinical trials were not as good as expected, and some were contradictory between them. One possible contributing cause to this discrepancy is that studies are carried out in young healthy animals, while patients susceptible to receive these treatments normally are aged people and in many cases with other co-morbidities [35].

Different ways to deliver cells into the damaged heart have been explored: intracoronary infusion (with the hope that cells will migrate through the vessels and be hosted in the infarcted area) or directly into the infarcted area either by intramyocardial or endocardial injection [36], as shown in figure 1 a. The advantage of injecting them directly into the infarcted area is that this will ensure that the cells are delivered in the site of interest.

2.2. Related problematic

Many different cell types have been employed in the numerous studies that have been done. Autologous cell sources are interesting because they do not require immunosuppression treatment of the patient and there is no risk of illness transmission. On the contrary, allogenic cells could be ready to use whenever a patient needs them, but would require immunosuppressive therapy after their implantation, and there is always a remaining risk of illness transmission. Another disadvantage is that prior to implantation cells need to be extracted and expanded. This whole process in some cases may take several weeks, limiting its application in the acute state. Besides, autologous cells coming from patients that suffer other conditions like diabetes or are simply aged, may have limited proliferation and attachment [37].

An important aspect of this technique is the low engraftment into the heart tissue of the supplied cells. The retention of the cells in the heart seems to be determined by the cell type and delivery route [38]. It has been estimated that in humans 50-75 min after intracoronary injection of bone marrow cells only 1,3-2,6% of the injected cells remain in the myocardium [39]; after 2 hours less than 10% of the injected cells survive [32]. Many causes can be advanced: the heart beats, so cells can easily be pumped out of the heart; the solution in which cells are injected has a low viscosity, so cells can be washed away; the mechanical loss of the cells through the injection hole left by the needle, etc [40]. A different contributing cause to the low cell engraftment is that the injured heart is not a cell-friendly environment, type I collagen fibers have been substituted by type III, which has worse properties in terms of adhesion and promoting angiogenesis, what can induce anoikis [4]. Another problem is cell survival itself. The conditions in the infarcted myocardium are very hostile for the cells: hypoxic conditions (studies show that the survival of injected cells decreases towards the center of the scar), cytokines, inflammatory factors, etc., are present in the damaged myocardium, and can negatively affect the survival of the injected cells. Immunological rejection can be another cause reducing cell survival [41].

An interesting approach is to train cells prior to their implantation for them to resist the hostile conditions they will find in the implantation site. For instance, the resistance to hypoxic conditions is key and needs to be improved even for skeletal myoblasts (which are the cells that have better resistance to lack of oxygen). Privation of glutamine reduces the oxygen consumption rate, what has been proved to improve survival of myoblasts when implanted [42].

The fact that most of the cells did not graft into the host myocardium in the studies performed to date, that there is a very limited transdifferentiation of implanted cells into beating cardiomyocytes (the differentiation reported in animals may have been fusion events between native cardiomyocytes and injected cells [41]), and that a wide range of non-myogenic cells also induce an improvement of the ventricular function [36], suggests that the mechanism leading to this enhancement cannot be only myogenesis regenerating the myocardium. The pathways through which cell implantation induces improvements in cardiac function remain to be elucidated, but different events that can take place simultaneously have been proposed. The most remarkable are the induction of angiogenesis (formation of new vessels) and the improvement in the myocardial perfusion, the reduction of the wall stress because of the increase in cell mass [43] and the paracrine effect of the injected cells [32].

2.3. Cell types investigated

As previously said, many cell types from different origins have been employed: embryonic stem cells, mesenchymal stem cells, bone marrow cells, induced pluripotent stem cells, cardiac stem cells, skeletal myoblasts, umbilical cord blood cells and amniotic fluid stem cells, among others. In what follows the use of these cell types is discussed, with the advantages and disadvantages that each one presents for its application in heart regeneration.

Embryonic Stem Cells (ESC)

ESC can be obtained from the inner mass of an embryo in the blastocyst stage. These cells have the capacity of growing undifferentiated indefinitely, and when they differentiate they can form any cell from the three germ layers. But the use of ESC raises ethical issues, requires immunosuppression, and has the risk to form theratomes. Their use in clinical trials has been limited because of these ethical considerations and risks [36, 44].

A protocol for ESC differentiation into cardiomyocytes and improving their survival when implanted has been established; when these differentiated cells were implanted in rodent models the heart function was improved [45]. In another study in mice, ESC-derived cardiomyocytes implantation reduced the reactive collagen deposition in the ventricular septum, which is one of the remodeling process hallmarks. Nevertheless, the implanted cells were isolated from the host myocardium by scar tissue, although the implanted cardiomyocytes were able to couple functionally to each other [46].

Induced Pluripotent Stem cells (IPS)

Induced pluripotent stem cells are fibroblasts treated with viral factors to recover their pluripotency. Therefore, IPS do not raise the ethical concerns of the ESC. IPS are very interesting because they can be autologous pluripotent cells. However, their application in clinical trials has been limited precisely for the use of viral vectors that may promote malignancy and act as oncogenes [43], as well as for the intrinsic risk of theratomes inherent to their pluripotency [44].

Adult stem cells

These cells have the advantage of being autologous and can be obtained from different sources like bone marrow or adipose tissue. In addition, they can be expanded in vitro and do not raise ethical or immunologic problems [47, 48].

Bone marrow cells (BMC) are easily accessible, can be obtained rapidly and have been reported to have certain plasticity. This property allows them to differentiate in vivo into cardiomyocytes [26] (although this fact remains controversial [42]). They can also differentiate into cardiomyocytes in vitro by supplementing the medium [49]. Studies in animal models demonstrate that the injection of these cells increases neovasculature improving heart function [42]. But the use of BMC is not exempt of risks: intracoronary administration of them can cause microinfarctions due to their big size and irregular shape, making necessary the use of an alternative way of delivery [50]. In clinical trials, results indicated only temporary benefits or no improvement after cell administration [38, 51]. A strategy to enhance the therapeutic efficacy of BMC is to precondition them: BMC treated with growth factors improve the therapeutic effect when implanted and show greater survival rate [52].

Adipose derived stem cells (ASC) can be obtained in great quantity without culturing them. These cells have been implanted in small animal models of AMI and left ventricular function was improved [48]. The underlying mechanisms are unclear, although the hypothesis of a paracrine effect is considered [53]. Clinical trials are ongoing for the implantation of ASCs: PRECISE and APOLLO [54]. These cells are also are under study at the moment in the RECATABI project [55] as part of a strategy that combines them within a three-dimensional polymer scaffold with a peptide gel filling, to lengthen their positive effect and serve as a mechanical support for the dilated ventricle.

Cardiac Stem Cells (CSC)

CSC are undifferentiated cells found in the heart that can become endothelial cells, smooth muscle cells, and functional cardiomyocytes [36]. In undamaged hearts, these cells seem to contribute to the normal self-renewal of the tissue. CSC can be isolated from biopsies and can be expanded in vitro [56], although there is a lack of availability from human origin as they are obtained from biopsies. Human CSC injected in mice hearts after infarction led to functional improvement and to support myocardial regeneration [57]. Currently, autologous cardiosphere-derived cells are being evaluated in the CADUCEUS clinical trial [58].

Skeletal Myoblasts (SM)

SM are cells present in the basal membrane, where they remain in a quiescent state while there is no damage. These cells have better resistance to hypoxic conditions than many other cell types, and can be from autologous origin, but 2 to 3 weeks are necessary to establish and expand myoblasts from skeletal muscle biopsies [36]. These cells are capable to contract; that is the reason why they were expected to attach to the beating cardiomyocytes and contribute to the effective beating by integrating in the working syncytium muscle. Nevertheless, there is no electro-mechanical coupling between the implanted cells and the native cardiomyocytes. This absence of coupling turns the implanted cells into a pro-arrhythmic substrate [44]. The cause for this uncoupling is the lack of the gap junctional protein connexin 43. Therefore, the implantation of a pacemaker or a defibrillator to avoid malignant arrhythmias and sudden death would be necessary when implanting these cells, to obtain a synchronous beating of the heart and the grafted cells [26, 59]. Despite the lack of electro-mechanical coupling of the myoblasts with the host cardiac cells, improvements in the ventricular performance have been observed in animal models, even with a reduced number of grafted cells, suggesting a cytokine-mediated effect [46].

The encouraging preliminary results and its autologous origin made this cell type the first to reach clinical trials. Initial clinical trials carried out with these cells showed symptomatic improvements in the patients, but some of them experienced arrhythmias, making necessary the use of implantable defibrillators [36]. For instance, in the phase II randomized placebo controlled trial MAGIC [60], skeletal myoblasts and a cardioverter defibrillator were implanted during a coronary artery by-pass graft surgery.

Umbilical Cord Blood Cells (UCBC)

UCBC can be easily obtained from the umbilical cord and do not present ethical concerns [42]. These cells have certain plasticity and reduced risk of rejection because they show low immunogenicity [25]. Their injection in animal models has been found to improve their left ventricular function [61].

Amniotic Fluid Stem Cells (AFSC)

Amniotic fluid is extracted for prenatal diagnosis and AFSC are isolated from it. They have many characteristics of ESC and seem to be in an intermediate stage between embryonic and adult stem cells in terms of versatility. Interestingly, these cells do not present ethical concerns and do not present risk of tumorogenicity [62].

Human AFSC have been successfully differentiated into endothelial or cardiac lineages in vitro. When these cells were implanted in an immunosupressed rat model, they contributed to attenuate its left ventricular remodeling, to preserve the thickness of the ventricle and to improve cardiac function [63].

Advertisement

3. Cell sheets

The use of cell sheets is based on the fact that when cells are cultured in normal flasks and enzymatically digested to detach them, the adhesive proteins and membrane receptors are disrupted leaving the cell damaged [64]. The alternative is to grow cell sheets and then detach them from the culture surface in a way that keeps the electromechanical connections between the cells and benefits from the fact that cells are kept together by their own deposited ECM, as figure 1 b displays. In that way, cells maintain the adhesion and membrane proteins, as well as the natural pro-survival and maturation environmental cues that the ECM provides [65]. Altogether, this is expected to help them to survive when implanted onto the infarcted myocardium.

Cells can be cultured, for instance, on temperature-responsive poly(N-isopropylacrylamide) (PnIPAAm)-coated plates. PnIPAAm is a hydrophobic polymer at 37ºC, and cells can attach to its surface. When the temperature is lowered, PnIPAAm suffers a transition to a hydrophilic state and this change causes the attachments of the cell monolayer to the surface to disrupt, and the entire cells sheet detaches from the surface [65]. Other materials, such as a thermo-responsive methylcellulose hydrogel, have been used to successfully obtain cell sheets fragments of human amniotic fluid stem cells (hAFSCs) [66]. Results obtained with these cell sheet fragments were superior to those with dissociated cells in terms of heart function, cell retention, proliferation and vascular density. Moreover, cardiomyocyte sheets were found to functionally integrate with the host tissue in a rat myocardial infarct model [67]. New techniques based on patterning with a gelatin stamp the thermo-responsive substrates allow obtaining complex tissue structures with cells having a determined orientation [68].

The muscle mass loss following an infarction is significant, up to 50 g [69], so the amount of cells needed to overcome this loss is obviously not covered with a single sheet of cells. On the other hand, when several layers of cell sheets are superimposed, they are easier to handle. Some groups have tried to obtain thicker grafts by overlapping several monolayer cardiomyocyte sheets, which adhere one to another forming gap junctions and intercellular adhesions within minutes [70]. But this approach poses a problem: as cell sheets lack of vascularization, the maximum thickness that can be achieved by overlapping them is limited to the depth at which diffusion of oxygen and nutrients can take place (a maximum of three cardiomyocyte sheets can be piled up). To try to overcome this problem, three-layer thick cardiomyocyte sheets were implanted in rats at 1-, 2- and 3-day intervals [71]; in the time between transplantations it was assumed that there is enough time for the cell sheet to be vascularized. With this approach constructs of 1 mm were obtained successfully. But anyway, this option is very invasive, so its application in patients might be limited.

A different approach based on the same idea of providing cell-cell connections and ECM to the implanted cells to improve their retention and survival is to implant them as spherical cell-bodies. Human amniotic fluid stem cells (hAFSC) cultured in a methylcellulose hydrogel to form cell aggregates were implanted in immunosuppressed rats as cell-bodies, and cell retention and engraftment were enhanced as compared with disaggregated cells. This enhancement led to functional improvement and limited the progression of heart failure [72].

Advertisement

4. Injectable gels

4.1. Rationale

As previously stated, cell cardiomyoplasty presents problems in terms of cell attachment and survival. Cells usually reside in a determined microenvironment which regulates their fate and function. The surrounding ECM with its chemical and biophysical cues is a key element, so the lack of cell-ECM interaction limits their survival [73]. To try to overcome the problems of cells supply, alternative approaches are considered in current studies. The use of natural or synthetic materials in an injectable format, alone or together with cells (figure 1 b), has been investigated to limit remodeling and improve both cell attachment and survival upon implantation in the heart. Ideally, they should be tailored to be amenable to delivery with minimally invasive catheter based procedures [69]. The injectable materials have to cure or self-assemble rapidly (without the need or the release of toxic components) once delivered in the site of interest. As injected, they adopt the shape of the cavity, and may increase the stiffness and thickness of the ventricular wall [74]. Simulations showed that the injection of non-contractile materials with proper mechanical properties can contribute to limit the stress the ventricular wall withstands, thus helping to limit the remodeling [75].

These materials can help to keep the cells in the site of interest, provide them a 3D environment and also protect them from the hostile environment represented by the cytokines and hypoxic conditions, reactive oxygen species, etc., consequence of the infarcted condition of the site [41]. The injected gels can provide a cell friendly environment that will prevent anoikis [69]; they can also include adhesion motifs and then actively contribute to cell attachment. Moreover, they can be used as a controlled release system providing in a sustained way drugs or growth factors to improve cells survival, integration and proliferation [32]. And in the case of bioactive materials, their degradation products may provide additional chemicals that stimulate cells.

Among others, the ideal injectable material should be biodegradable, have a low immunogenity, be no cytotoxic, non-adhesive and have antithrombogenic properties, adequate mechanical properties, provide stiffness to the scar but at the same time being compliant with the heart beating and transmit properly the mechanical stimuli to the cells, induce angiogenesis or at least not disturb the angiogenic activity after incorporation, be capable of delivering cells and or bioactive molecules [76]. Next, some of the materials investigated for their potential use as injectable ones are described.

4.2. In situ gelling biomaterials employed

4.2.1. Natural materials

Fibrin

Fibrin is a natural biopolymer that forms the natural provisory matrix for wound healing. It is FDA approved for many applications and there are different preparations commercially available, but it can also be obtained from autologous origin [77]; it is biocompatible, not toxic, or inflammatory [78]. Besides, some of the degradation products of fibrin have interesting properties, like improving healing promotion or a protective effect against myocardial reperfusion injury [79]. Fibrin contains arginine-glycine-asparagine (RGD), which are known cell adhesion motifs [77]; it is cytoprotective for anoxia and provides a favorable microenvironment for cardiomyogenic differentiation of marrow-derived cardiac stem cells [77]. It can also be used as a controlled release system [80]. In sum, fibrin as a gel is a potential candidate to enhance cell adhesion and survival. To obtain the fibrin, fibrinogen monomers in saline solution are mixed with thrombin and they polymerize forming a 3D net by mechanisms similar to normal clotting in vivo [81]. The properties of the network can be tailored by modifying the polymerization process.

A concern about translating the fibrin glue for cardiac tissue engineering into the clinic is the risk of inducing intravascular thrombosis [79]. The concentrations of fibrin amenable to delivery through current percutaneous catheters have been studied, demonstrating the feasibility of using fibrin in a non-invasive injectable application [81]. The injection of fibrin alone was proved to preserve left ventricular geometry and cardiac function in a rat acute MI model [82]. But it has also been combined with many types of cells. As an example, it was employed to deliver bone-marrow derived mesenchymal stem cells, which enhanced cell retention and prevented their redistribution in other organs, improving the beneficial effects of the treatment [81]. Injection of fibrin combined with myoblasts [82], bone marrow stem cells [83] or with autologous endothelial cells [84], improved the results obtained with cells alone.

Chitosan

Chitosan (CHT) is a natural cationic polysaccharide, obtained from the deacetylation of chitin of the mollusks, crustaceans and insects. It is soluble in acidic aqueous solution but after neutralization forms a gel-like precipitate [85]. CHT exhibits numerous positive biological and physicochemical properties: biocompatibility, non immunogenicity, and can be conjugated with various molecules thanks to the amino groups on the polysaccharide backbone [86]. A thermally responsive chitosan-based polymer was capable of scavenging the reactive oxygen species produced by the ischemic conditions and recruit key chemokines for stem cell homing such as SDF-1. As a cell delivery system with adipose-derived mesenchymal stem cells, this material was capable of improving the microenvironment for the cells when injected in the infarcted myocardium of rats, improving their survival and engraftment [87]. Chitosan mixed with collagen has been conjugated with QHREDGS (peptide thought to mediate attachment and survival responses of cardiomyocytes) in the format of a thermoresponsive hydrogel to improve maturation and metabolic activity of cardiomyocytes [86]. Alginate-chitosan nanoparticles have been loaded with placental growth factor (PlGF) to increase the left-ventricular function and vascular density in rats [88].

Matrigel

Matrigel is a commercial ECM proteins mixture that undergoes a temperature mediated sol-gel transition, and is obtained from the ECM of mouse sarcoma cells [27]; its clinical application is limited precisely by the source from which it is obtained. It has been implanted alone and in combination with mouse ESC [89] or neonatal cardiomyocytes [90] into a mice model of infarcted myocardium. The gel prevented worsening of the cardiac function, but animals receiving both Matrigel and cells maintained more wall thickness and preserved better cardiac function in terms of fractional shortening and regional contractility [91].

Hair keratin

Keratin materials can be obtained from hair, importantly from autologous source. More than 30 growth factors are involved in hair morphogenesis, and the residual of them remains in the keratin, what can be beneficial for cardiac repair. Lyophilized keratin powders have the ability to self-assemble upon addition of water, and form gels. Keratin has been implanted onto infarcted rat hearts, and native cardiomyocytes as well as endothelial cells were able to infiltrate the keratin gel, promoting angiogenesis without inducing inflammation; after 2 months animals exhibited preservation of cardiac function and limited ventricular remodeling [92]. These improvements were attributed to the biomaterial’s contribution to the mechanical support to the ventricular wall and the presence of cell binding motifs in it.

Alginate

Alginate is a linear block co-polymer of (1-4)-linked β-D-mannuronate and α-L-guluronate residues obtained from seaweed. It is a negatively charged polysaccharide that gels by the presence of calcium ions and is non-thrombogenic [4]. The properties of this material can be tuned either by changing the concentration of the solutions or by controlling the molecular weight. Greater concentrations will increase mechanical strength but also will increase the solution viscosity and the degradation time of the gel [27].

Alginate has been used as an injectable material in recent and old infarcts in rats, and it was observed that its injection augmented the scar thickness and limited systolic and diastolic dysfunction [93]. It has also been proposed as a controlled delivery system: based on the different binding affinity of alginate to insulin-like growth factor-1 IGF-1 and hepatocyte growth factor HGF, a dual delivery system of these factors was developed [94]. The hydrogel beads protected the proteins from degradation maintaining their bioactivity and increasing the therapeutic effect of the system.

Alginate sustains very low protein adsorption and it does not support mammalian cells attachment [95], but it can be combined with adhesion motifs to improve its attachment properties. Its conjugation with RGD increased the arteriole density in a rodent model of chronic ischemic cardiomyopathy [96]. However, the combination of alginate with RGD and tyrosine–isoleucine– glycine–serine–arginine (YIGSR) reduced the therapeutic effects of the hydrogel in terms of scar thickness, left ventricular dilation and function [97]. Another modification of alginate has been the addition of the electrical conducting polymer polypyrrole [98], which increases arteriogenesis and promotes myofibroblasts infiltration.

Hyaluronic acid

Hyaluronic acid (HA) is a non-sulfated glycosaminoglycan prevalent in the extracellular matrix of many tissues. HA plays an important role in homeostasis, transport of nutrients and also mediates the inflammation and repair processes. It is biocompatible, non-immunogenic, biodegradable and has different biological activities depending on its molecular weight. Precisely the low molecular weight degradation products of HA stimulate angiogenesis and endothelial cell proliferation and migration [99]. It can be functionalized to improve its biological development, for example with PEG-SH4 [100]. Moreover, it is a FDA-approved material for its use in humans in certain applications like dermal and intra-articular injection.

There are already commercially available in situ crosslinkable HA-derived hydrogels. Different types of HA hydrogels have been compared with commercial fibrin, poly(vinyl alcohol)-chitosan and elastin hydrogels, in terms of in vitro degradation rates and cytotoxicity and in vivo degradation, immune response and angiogenic potential [76]. Traut’s grafted HA hydrogel and periodate oxidated HA hydrogel, especially the first one, demonstrated to be the most suitable for new artery formation in ischemic myocardium because they were both digested within 2 weeks with low immune response and strong angiogenesis compared with the other examined hydrogels.

HA alone does not support cell adhesion. Cardiosphere-derived cells were delivered using a thiolated hyaluronan-based hydrogel crosslinked with thiol-reactive poly(ethylene glycol) diacrylate and covalently linked or not with thiolated denatured collagen. It was observed that the retention rate achieved with the hydrogel without collagen was similar to that of cells delivered in phosphate buffer saline (PBS), either by a low physical retention or poor cell survival and adhesion of HA [101]. In the in vivo study in a mouse model of myocardial infarction, some functional benefits were observed though.

Collagen

Collagen supports growth and survival of cardiomyocytes in vitro, and is one of the main components of the ECM in the adult heart [102]. Commercial collagen alone has been implanted in animal models showing improvements in ventricular cardiac function and geometry [103]. In another study in a myocardial infarction model in rats (with ischemia-reperfusion model this time) increased capillarity density and myofibroblasts infiltration after 5 weeks were reported [104].

The therapeutic potential of injectable collagen has been evaluated in combination with different cell types. Bone marrow stem cells were injected via catheter in a swine model in combination with collagen, demonstrating the feasibility of a non-invasive delivery of this system [105]. Collagen was also used as a carrier for mesenchymal stem cells (MSC) transplantation to improve the retention of the cells in the infarcted myocardium [106]. 4 weeks after implantation, rats receiving cells in saline suspension, had the implanted cells in remote organs, whereas in animals receiving the cells with collagen, were detected to a lesser extent in remote organs. However, cardiac function was improved in animals receiving cells in saline and collagen alone but not in the combined collagen MSC group. The mechanisms underlying this negative interaction (controverted in other works) are unknown, but is suggested that collagen may limit oxygen and nutrients diffusion, and compromise cell-cell interactions. In another study, collagen combined with chondroitin 6-sulfate was employed to deliver CD-133+ progenitor cells derived from peripherial blood after expansion in vitro [107]. It was expected that the material would improve cell adhesion and survival into ischemic hind limb athymic rats. The collagen increased two-fold the number of cells retained when implanted alone; the implanted material was vascularized and the injected cells added into vascular structures.

Gelatin

It is a non-immunogenic partially degraded product of collagen [108]. It has been injected as a hydrogel in rat infarcted hearts bare or loaded with basic fibroblast growth factor; adding the factor improved arteriogenesis, ventricular remodeling and function [109]. Basic fibroblast growth factor has also been delivered with gelatin microspheres [110], inducing angiogenesis and improving cardiac function. The loaded nanoparticles induced an increase in the blood flow in the infarct border (thanks to stimulated angiogenesis), and as a result left ventricular function was improved.

ECM-derived materials

A different approach is based on decellularized tissues, their digestion and injection. This type of materials has the advantage of containing a physiological proportion of the native components of the ECM [102] and cues for cell-matrix interactions. ECM coming from different tissues has been studied, and apparently the ECM of each tissue has its unique combination of proteins and proteoglycans. This makes of myocardial decellularized matrix, among all other tissues matrices, the best candidate for myocardial repair when it is available [111]. Decellularized porcine myocardial tissue able to self-assemble into a nanofibrous structure similar to collagen in vitro at 37°C and deliverable in vivo upon catheter injection was tested in rats. It induced endothelial cells and smooth muscle cells migration increasing the arteriole formation at 11 days post-injection [111].

Small intestinal submucosa (SIS) is a dense sheet of acellular extracellular matrix. This material is used in the clinic for accelerated wound healing. SIS supports proliferation, attachment and migration of various cell types and stimulates angiogenesis thanks to the growth factors and binding motifs embedded in the matrix. Two different types of commercial available SIS-derived gels have been studied as an injectable material for cardiac repair in a murine model [112]. The two materials differed in the concentration of basic fibroblast factor, obtaining best result the material richer in this factor. In another work, an emulsion of digested ECM from SIS was injected into infarcted rat hearts, improving cardiac function, increasing neovascularization and promoting cell recruitment [113].

4.2.2. Synthetic materials

Synthetic materials are made in the laboratory from primary building blocks, so their properties can be tuned to match desired characteristics. Besides, they are free from animal origin components and the risks related therewith.

Thermosensitive hydrogels

This group of materials has temperature-dependant sol to gel transition. The great advantage of this group of materials is the possibility to tune their properties for them to undergo the gelation transition around body temperature [114]. In this way they can be comfortably manipulated and injected and only when they are inside the body they will undergo the transition.

Some of the materials of this group are based on N-isopropylacrylamide (NiPAAm). It is non biodegradable, but copolymerized with degradable polymers becomes biodegradable. For instance, NiPAAm was copolymerized with acrylic acid (AAc) and hydroxyethyl methacrylate-poly(trimethylene carbonate) (HEMAPTMC) [115]. The ratio of each material was adjusted to obtain a hydrogel at 37ºC. It can also be degraded in vitro with a mass loss over 85% after 5 months. This material was injected in vivo in rats and proved to preserve the area of the left ventricular cavity and contractility. Tissue ingrowth, a thicker left ventricle (LV) wall and greater capillarity density were also found when compared with PBS controls. After 8 weeks, a layer of smooth muscle cells with contractile phenotype was formed next to the remaining material.

Another family of thermoresponsive hydrogels based on polycaprolactone, N-isopropylacrylamide, 2-hydroxyethyl methacrylate and dimethyl-g-butyrolactone acrylate has been developed [116]. Cardiosphere derived cells (CDC) combined with the hydrogel were suitable for myocardial injection and the solutions formed solid gels within 5 s at 37ºC. Hydrogels with different mechanical properties were obtained and it was shown that they influence the fate of the CDC differentiation. Another thermoresponsive material containing biodegradable dextran chain grafted with hydrophobic poly(ε-caprolactone)-2-hydroxylethyl methacrylate (PCL-HEMA) chain and thermoresponsive poly(N isopropylacrylamide) (PNIPAAm) (Dex-PCL-HEMA/PNIPAAm) has been synthesized. It can shift from sol to gel within 30 s and is reversible within the same time frame [117]. It was injected in rabbits, 4-days post-infarction. Histological analyses one month later indicated that the material prevented the scar expansion and thinning of the wall. Left ventricular ejection fraction was increased and it attenuated left ventricular systolic and diastolic dilation.

Poly (Ethylene Glycol) (PEG)

A strategy based on non-biodegradable in situ crosslinkable PEG hydrogel has been developed, to provide a permanent support to limit the remodeling [118]. Its therapeutic effects were tested in rat myocardial infarction model at short and long term. Beneficial effects were observed at 4 weeks, but at long term (13 weeks) it was unable to prevent the dilation. Besides, the material injection induced some inflammatory response.

An injectable α-cyclodextrin/poly(ethylene glycol)–b-polycaprolactone-(dodecanedioic acid)-polycaprolactone–poly(ethylene glycol) (MPEG–PCL–MPEG) hydrogel was used to deliver and encapsulate bone marrow stem cells into infarcted myocardium [119]. The CD/MPEG-PCL-MPEG hydrogel alone does not induce angiogenesis, but can serve as a support in the infarcted zone and contribute to inhibit the left ventricular remodeling. One month after the injection of the gel combined with cells, cell retention and survival and the density of vessels were increased when compared with cells injection alone; moreover, the gel was absorbed, ventricular dilation was limited and the ventricular ejection fraction improved.

PEG-based temperature-sensitive hydrogels have also been combined with growth factors or other molecules. VEGF was mixed or conjugated with the aliphatic polyester hydrogel poly(δ-valerolactone)-block-poly(ethylene glycol)-block-poly(δ-valerolactone) (PVL-b-PEG-b-PVL); the sustained VEGF release during the degradation time of the hydrogel translated into an improvement of the myocardial and functional recovery, in dependence of the preparation method [120]. In another work, a metalloproteinase-responsive PEG-based hydrogel was synthesized to be a thymosin β4 (a pro-angiogenic and pro-survival factor) delivering scaffold. It was implanted combined with endothelial and smooth muscle cells derived from human embryonic stem cells (hESC) in rats [121]. The gel provides structural organization and when was loaded with cells and thymosin b4 enhanced more contractile performance than when the hydrogel was only loaded with the factor, because of their paracrine effect. Another PEG-based hydrogel, α-cyclodextrin/MPEG–PCL–MPEG, was tested as a delivery system for erythropoietin (EPO) [122], a hormone that plays a protective role in the infarcted myocardium. Rats treated with this system showed limited cell apoptosis and increased neovasculature formation; also infarct size was reduced and cardiac function improved.

PEG in the format of nanoparticles has also been studied. They can be injected intravenously, circulate in the body for long periods and bind only to desired tissues. Nanoparticles targeting the infarcted myocardium were developed based on the overexpression of angiotensin II type 1 (AT1) receptor in the infarcted heart [123]. The system was formed by a vehicle and a targeter, a ligand specific to AT1 that will make the nanoparticles bind specifically. The vehicle was 142 nm diameter PEGylated liposomes, which could carry therapeutic molecules and release them in a controlled way. This system was proved to target the infarcted heart in mice model, but not the healthy.

Self Assembling Peptides (SAPs)

SAPs are short peptides capable of forming hydrogels at physiological pH and osmolarity [124]. When the SAPs solution is placed in contact with ions or pH is changed, the charges are partially neutralized and a hydrophobic packing takes place forming beta-sheet structures, constituting fibers that build a 3D network if the concentration is high enough. Fibers shape is different depending on the nature of the employed peptides. In the particular case of the RAD16 ionic peptides family (R: arginine, A: alanine, D: aspartate) fibers thicknesses are of 5-10 nm.

Peptides can be combined with cells to encapsulate them within the peptide network [125]. RAD16-I (AcN-RADARADARADARADA-CNH2) has proved to be a useful synthetic gel capable of maintaining the cells in the site of interest, and has been used as a delivery system of different types of cells to the heart. On the contrary, when it was implanted alone limited improvements were observed in the infarct area and the remodeling process. RAD16-II (AcN-RARADADARARADADA-CNH2) peptide has been shown to create microenvironments in the infarcted myocardium that are infiltrated with endothelial and smooth muscle cells, suggesting a potential for vascularization [124]. It was also observed that combining RAD16-II with neonatal cardiomyocytes the density of endogenous α-sarcomeric actin positive cells increased.

As stated, SAPs gels can be modified to incorporate growth factors or drugs. The self assembling peptide RAD16-II has been used as a drug vehicle to deliver both platelet derived growth factor and fibroblast growth factor (PDGF-BB and FGF-2) [126]. The first is arteriogenic and the second is angiogenic; their combination targets endothelial cells (EC) and vascular smooth muscle cells (VSMC). Infarct size and cardiomyocyte apoptosis were considerably reduced in rats. The capillary and arterial density was recovered, and cardiac function was almost recovered. This system also induced long-lasting vessel formation. RAD16-II combined with IGF-1, a cytokine that protects and promotes cardiomyocytes growth, has also been used as a delivery system for cardiomyocytes [127]. The addition of IGF-1 acted reducing cell apoptosis and improving systolic function.

Advertisement

5. Preformed gels and scaffolds

5.1. Rationale

An alternative approach in the field of cardiac tissue engineering involves the use of biomaterials to produce patches ex vivo and implant them epicardially onto the infarcted tissue, conveniently adapted to its size and shape. These patches can be pre-loaded with cells (incorporated within their pores in the case of microporous scaffolds, or encapsulated in the case of a gel conformed before implantation, as shown in figure 1 b) and growth factors or drugs, and act as a cell supply, a mechanical reinforcement to the infarct scar to avoid ventricular dilation and a drug release system simultaneously.

5.2. Requirements of the scaffolds

In this strategy a key aspect is to find a material that matches the required properties. The material needs also to be cell-friendly, non-cytotoxic and promote cell attachment and proliferation, and it must also be non-immunogenic [128]. The scaffolds should provide a 3D environment to the cells with a porous structure able to guide cardiomyocytes alignment and promote maturation, also induce the development of a contractile phenotype and the electro-mechanical coupling of the implanted cells among them, and also with the host tissue [129, 32] and need to be easily vascularized [37].

The mechanical properties exhibited by the scaffolds should be adequate to their application in heart tissue engineering. It implies that they should ideally be compliant with contractions and exhibit non-linear elasticity, as well as be capable to adapt to the shape of the heart in all phases of the heart beat. Anisotropy to mimic the directionally-dependent electrical and mechanical properties of the native myocardium is important too [130]. Besides, the stiffness of the material employed affects to a great extent the phenotype and contractile properties of the neonatal cardiomyocytes [131, 132], and has to be carefully tuned to match physiological conditions. During heart development, the ECM on which cardiomyocytes maturation takes place, stiffen 9 times. An interesting approach to mimic it is the development of materials with time dependant mechanical properties [133]. For instance, hyaluronic acid hydrogels that stiffen with time form more contractile units when compared with cultures in hydrogels without such time-dependant stiffness.

Attending to the type of strategy, three groups can be distinguished, in terms of the nature of the matrices: biologically-derived materials, synthetic (either biodegradable or biostable) materials and decellularized tissues. With the use of biodegradable scaffolds, it is expected that the matrix will degrade as the surrounding tissue is regenerated; the degradation products should not be toxic and metabolized by the body. By using permanent scaffolds, the idea is that they will be infiltrated by the host tissue and contribute to the regeneration, but also act as a permanent mechanical restraint to limit ventricular dilation. The approach of scaffolds derived from decellularized tissue is based on the use of tissues whose cells are removed and the remaining ECM maintains the architecture and mechanical properties similar to those of the native tissue. Obtaining a scaffold matching the desired properties is a hard task, as many different properties are required; thus, materials exhibiting different properties have been mixed in more advanced strategies to obtain a composite that combines them.

5.3. Related problematic

As all the approaches described so far, this one also has some advantages, disadvantages and unsolved problematic. An important disadvantage is that the application of a patch in the heart needs a much more invasive technique than a catheter-delivered system, as it requires a surgical procedure to be implanted. As advantage, the fact that the materials are synthesized and conveniently prepared out of the body can be outlined. It implies that there is no limitation in the preparation procedure and in the use of solvents (if they are properly removed at the end of the fabrication process and do not induce cytotoxicity). Therefore, the range of chemistries and techniques available to obtain scaffolds with different architectures is broadening. Besides, cells can be pre-cultured in vitro within them prior to implantation if desired. In addition, the mechanical properties of polymer scaffolds may be tuned to match more closely those of the heart muscle than with gelly biomaterials.

Unlike native myocardium, where the greatest distance between capillaries is around 20 microns [69], scaffolds are not vascularized a priori. Then, cells seeded in the scaffolds have their oxygen and nutrients supply limited to their molecular diffusion through the thickness of the scaffold. Given the fact that cardiomyocytes have great consumption rates of nutrients and oxygen, diffusion is insufficient supply for thick constructs. Consequently, to obtain a thick engineered tissue with viable cells through all its thickness, pre-vascularization or improved diffusion throughout the scaffold until it is vascularized is key for the implant to succeed. Otherwise, cell density will be concentrated in the external parts and cell viability will be compromised in the center of the scaffold if the distance to the surface is greater than a critical value estimated around 100 microns [134]. For example, the influence of oxygen concentration in cell density and viability in collagen scaffolds has been studied, the former decreasing linearly with the distance to the surface and the latter exponentially [135]. These results indicate that in order to guarantee an appropriate oxygen concentration throughout the scaffold, additional measures need to be taken.

Many attempts have been done in this direction, like the addition of oxygen carriers to the culture medium to simulate the effect of the hemoglobin in the blood. Their addition contributed to improve mass transport and to increase cell density [136]. Another strategy includes the use of scaffolds releasing growth factors to enhance the vascularization process, like basic fibroblast growth factor [137], vascular endothelial growth factor (VEGF) [138] and Thymosin beta-4 [139]. Another approach is the addition of the growth factor platelet derived growth factor BB to the culture medium to protect cardiomyocytes from apoptosis [140]. In a different methodology, channeled scaffolds were produced to simulate the capillary structure of the native tissues and guide endothelial cells growth. The porosity might be adjusted to increase capillary infiltration but it is limited to the maximum size of the pores on which endothelial cells can form vascular structures [141]. An alternative involves the use of decellularized tissues that already provide a native vascular network [142, 143]. The culture of endothelial cells prior to implantation of cardiac myocytes has also been explored [144], and reduced cardiomyocytes apoptosis and necrosis was found. Another possibility is to pre-implant the scaffold to pre-vascularize it prior to its implantation in the final site: alginate scaffolds loaded with angiogenic and pro-survival factors (Matrigel, SDF-1, VEGF and IGF-1) were pre-implanted into the omentum of rats [145]. It proved to be a very interesting in vivo “bioreactor”, providing to the patch a functional vascular network that maintained the viability of the transplanted cells.

Pre-culturing the scaffolds in vitro in bioreactors has also been a considered an option. There are many types of bioreactors (stirring, spinning flasks rotating, perfusion, etc.), but not all of them improve enough the diffusion to lead to uniform cell density and compact tissue formation. As an example, in a study where rotating bioreactors were used to culture polyglycolic acid (PGA) scaffolds [146], functional and interconnected cells only were found in the peripheral parts, where there was a better diffusion of the oxygen. Perfusion bioreactors have been developed to try to reduce diffusional limitations by establishing interstitial flow through the scaffolds in order to allow the formation of thick tissues with uniform cell density throughout them. The effect of culturing scaffolds in perfusion bioreactors was compared with culturing them in spinner flasks [134] or orbital mixed dishes [147]. In both studies results were improved with the perfusion bioreactors; when cultured in the others, high cell density was only found in the outer layers. However, a limitation of perfusion bioreactors is the medium flow rate, because of the hydrodynamic shear the interstitial flow inflicts to the cells, which could maintain them in a rounded morphology or even wash them out if it is too high. This finding led to the combination of the perfusion culture with the use of channeled scaffolds that provided separated compartments for medium flow [148]. Even more, this strategy has been successfully combined and used simultaneously with a selective pre-seeding of the scaffold in the channels with endothelial cells using a perfusion seeding technique, which provides uniform seeding throughout the entire scaffold without the use of cell carriers [149].

Another step was made when the pulsatile perfusion bioreactor [150] was developed. It was expected that the pulsatile interstitial medium flow would provide mechanical conditioning and improved mass transport, intending that all together would lead to a tissue with better contractile properties. Indeed, scaffolds cultured under these conditions had enhanced contractile properties. A different type of bioreactor, with bidirectional slow flow perfusion obtained with an oscillatory system was tested with culture medium loaded or not with Insuline-like growth factor-I [151]. The advantage of the combined strategies was revealed.

However, despite the great efforts put and the improvements achieved, obtaining vascularized constructs is still an unsolved problem.

5.4. Preparation techniques

Many different techniques have been proposed to obtain 3D porous structures with different topographies and porosities, basically based in phase separation procedures or the use porogen templates to create the pores. Now with the introduction of controlled computer assisted systems, new possibilities are open. Next, a brief description of the main techniques employed to prepare scaffolds for heart tissue engineering is outlined.

The electrospinning technique is based in the application of a high voltage to a polymer melted or in a solution that leads to the formation of ultrathin nonwoven fibers [152], which are projected on a collector giving rise to fiber mats with controlled thicknesses. The fibers diameters can be obtained in the range of the ECM proteins. This technique also allows the preparation of aligned fibers, which can be applied to obtain aligned cardiac cells [153].

The particle leaching technique is based on the use of a porogen that is mixed with a polymer solution or a melted polymer. This porogen is removed after the solvent has been eliminated (solvent casting, freeze extraction) or the polymer has solidified after cooling, leaving empty spaces (pores) with the size and shape of the porogen template (and also small pores for the elimination of the solvent, if used). Porosity and pores interconnection can be tuned by changing the porogen-polymer ratio. Gas foaming avoids the use of solvents and high temperatures, because the pores are obtained by exposition to a high pressure gas followed by a pressure decrease with nucleation and growth of pores. The freeze-drying technique consists in freezing a polymeric solution and then lyophilize it to remove the solvent in the frozen state and obtain a solid porous structure [154]. Different morphologies can be obtained by changing the freezing conditions [155].

Microfluidic patterning consists in forcing a polymer solution through a channeled mould previously obtained with the desired geometry. Once the polymer is consistent, the mould is removed and the scaffold or patterned surface is ready. Selective laser sintering is a technique based in the use of a CO2 laser to sinterize selectively the powder of a material to form the cross section of each layer of a 3D object.

Microcontact printing is a technique that allows cell adhesion guidance [156]. It consists in the use of a stamp, with the pattern to be followed by the cells. The stamp is inked with the solution that is expected to promote the adhesion (laminin, ECM proteins, etc.) and then pressed against the substrate to transfer the solution. By loading the solution with growth factors, cell differentiation can also be induced in patterns [157].

5.5. Biomaterials employed as scaffolds

Many different types of materials have been considered for cardiac tissue engineering. According to their origin we can distinguish: biologically-derived materials, decellularized tissues and synthetic materials. Natural materials include collagen, gelatin, fibrin, silk and alginate; and synthetic materials include polyurethane (PU), polylactide acid (PLA), polyglycolic acid (PGA), polycaprolactone (PCL), or polyglicerolsebacate (PGS), among others.

5.5.1. Natural materials

Collagen

There are a number of commercial collagen patches, widely used by clinicians for other purposes, which are now under study as epicardial patches, because it has been reported to be a good substrate for cell attachment and infiltration [158]. They have been combined with different cell types and molecules. Unfortunately, collagen sponges have a great swelling rate and poor mechanical performance in aqueous medium.

Collagen can be used in two formats, as a porous scaffold or as a hydrogel. To obtain the scaffold a collagen solution is lyophilized and then rehydrated and seeded with cells. In the case of hydrogels, a collagen solution is mixed with cells ex vivo and then gelled. As a gel entrapping embryonic chick cardiomyocytes [159], it was found to beat and arrange as a highly organized tissue-like when pulses with different frequency were applied.

The potential of collagen scaffolds as an attractant for neovascularization was demonstrated in a study with rats [160]. Collagen sponges implanted in both healthy and cryoinjured hearts were almost absorbed after 2 months, but the remaining structures were populated by new arterioles and capillaries. In another study, collagen has been combined with chondroitin 6-sulfate to obtain porous scaffolds. These scaffolds delivered MSC in the infarcted region in a rat model, promoting neovascularization [161].

The therapeutical potential of collagen as epicardial patch has been compared with injectable approaches. Collagen matrices loaded with mesenchymal stem cells (MSC) [162], and collagen scaffolds loaded with human umbilical cord blood cells (hUCBCs) [25], gave better results than the injection of cells alone in mice. In the MAGNUM phase I clinical trial [163], intrainfarct cell therapy of autologous BMC was combined with collagen scaffolds loaded with BMC. This treatment was found to be safe and contribute to limit left ventricular remodeling by increasing the thickness of the ventricle wall and then reducing the stress of the wall.

Collagen has been modified to incorporate bioactive molecules to improve its biological behavior. Its scaffolds have been modified with RGD [164] and cardiac markers of cardiospheres derived from cardiac progenitor cultured on them were upregulated. Collagen functionalized with interleukin-10 plasmid [165] (an anti-inflammatory plasmid) increased 5 times cell retention and modulated inflammation.

Gelatin

Gelatin is obtained from chemical denatured collagen; it is therefore weaker and degrades faster than it [27]. It has been reported to provoke unspecific inflammatory response upon degrading; at first this can be considered an undesired effect, but for certain applications it might be beneficial for the positive impact that can have on angiogenesis [166]. A commercial gelatin sponge bare or cultured either with fetal or adult rat heart cells was implanted to replace the resected right ventricular outflow tract (ROVT) of rats [167]. After 4 weeks a great inflammation was observed and after 12 weeks the patches had endothelial cells on the endocardial surface. Nonetheless, the authors concluded that a material inducing less inflammatory response is needed.

Fibrin

Fibrin can be used as an injectable gel, but can also be preformed ex vivo, which broadens the possibilities of fabrication. For example, SDF-1 (a factor that is up-regulated for a period of time after a myocardial infarction, and contributes to mobilize cells from bone marrow and peripherial blood to the damaged tissue) was covalently bound to a PEGylated fibrin patch [168] and implanted in an AMI mouse model; the SDF-1 loaded patch reduced more significantly the scar area expansion and improved the left ventricular function than the un-loaded patch.

Alginate

Alginate scaffolds obtained by the freeze drying technique have been extensively explored in myocardial regeneration. Loaded with fetal cardiac cells and implanted in infarcted rats, they limited left ventricular dilation [169]. However, cultured with neonatal or fetal cardiomyocytes in static conditions, cell aggregates were formed due to the non-adhesive nature of the alginate [170].

To improve cell adhesion and survival modifications of alginate scaffolds have been investigated. For example, it has been modified to incorporate the adhesion peptide RGD [171], which improved cell adhesion, reduced apoptosis, accelerated tissue regeneration and led to the organization of cardiomyocytes in myofibers in vitro, and also with a combination of RGD and the heparin-binding peptide G4SPPRRARVTY (HBP) [172], with better results.

Polysaccharides

Polysaccharide-based scaffolds have also been investigated with myocardial regeneration purposes. The effectiveness of freeze-dried pullulan and dextran patches was compared to mesenchymal stem cells endocardial delivery alone in a rat myocardial infarction model [173], the scaffolds improving the cell engraftment and survival at 1 and 2 months.

Silk

Because of silk fibroin good mechanical properties, biological performance, and its easy processing to obtain different morphologies, it has generated interest in the tissue engineering field. Silk is produced by some insects like spiders or silkworms, and is considered a non-degradable material by the FDA [174]. Silk fibroin has been combined by chitosan and hyaluronic acid to produce microparticles that were pressed and crosslinked with genipin to obtain cardiac patches [175]. MSC cultured on the composite patches exhibited greater proliferation and cardiomyogenic differentiation than in silk patches.

Recently, non-mulberry silk fibroin from Antheraea mylitta has been investigated as a material for cardiac tissue engineering [176]. It has better mechanical properties than mulberry silk, contains RGD sequences, is non-cytotoxic and induces low level of inflammatory response. When neonatal rat cardiomyocytes were seeded in an Antherea mylitta silk lyophilized scaffold, the results were better than those obtained with a mulberry silk.

Decellularized-tissue derived scaffolds

Decellularized extracellular matrices have been used as scaffolds in many studies and also in preclinical and human clinical applications [177]. The decellularization process consists in a set of washes to remove the cells but maintain as much as possible the architecture, proteins and adhesion molecules. The more aggressive the washes and treatments are, the lower the risk of allogenic immune reaction is, but undesired washout of adhesion proteins and architecture damage can be associated [65].

Decellularized sheets have been tested in combination with fibrin, TGF-beta, and MSC and tested in a nude rat model of infarction with positive results [178]. A patch of urinary bladder-derived extracellular matrix (UBM) was implanted in pigs, as a left ventricular wall replacement after infarction, and compared with a polytetrafluoroethlyene (ePTFE) [177] one. At three months, the results were better with the UBM: it was reabsorbed and a cellularized and vascularized tissue rich in collagen was formed.

Sliced decellularized porous scaffolds of acellular bovine pericardia have been combined with cell sheets from bone marrow stem cells, cultured and implanted in rats replacing the resected infarcted myocardium [179]. The patch pores were filled by cells, new vessels and new muscle fibers, indicating that the graft was integrating. Cardiac function was improved and the dilated left ventricle was restored after implantation. In a revolutionary study entire rat hearts were decellularized, and then re-cellularized with neonatal cardiac cells [180]. The architecture was conserved and the preserved vasculature was perfusable. Seeded cardiomyocytes coupled electromechanically and after 8 days under external electrodes stimulation the re-cellularized heart beat and was capable to pump blood.

5.5.2. Synthetic materials

Synthetic materials are prepared in the laboratory, allowing precise control over their mechanical properties, degradation, morphology and porosity that can be tuned as desired [181]. However, they may not have as good biological performance as biologically derived materials [4].

Polylactic Acid and Polyglycolic Acid (PLA and PGA)

Polylactic acid is a biocompatible, biodegradable and FDA-approved polymer; it degrades into lactic acid (non-cytotoxic), and has been widely used in patients, for example as sutures. However, its degradation products can induce a slight, undesired, acidification of the microenvironment [65]. Polyglycolic acid is a thermoplastic too; it has also been used in the clinic and degrades into non-toxic products. However, neither PLLA nor PGA exhibit the desired elasticity to match that of native heart tissue. In many studies PLA and PGA have been combined as poly(lactic–co-glycolic acid) (PLGA), or other polyesters, to modify their properties as desired. Electrospun PLGA fibrous membranes with different compositions (having different hydrophobicity and degradation rates) [4] were found to align cardiomyocytes in the direction of the nanofibers, the best results being those of the slightly hydrophobic copolymers. Porous beads of PLGA seeded with human amniotic fluid stem cells (hAFSCs) have been tested as a cell delivery vehicle or “cellularized microscaffold” [182]; after implantation by intramyocardial injection in a rat infarct model, they showed good retention of the cells in the site of interest. PLGA has been treated with laminin [183] to improve its biological development and combined with carbon nanofibers (CNF) to increase its conductivity and cytocompatibility [184]. PLLA-PLGA scaffolds loaded with Matrigel have been co-cultured with endothelial cells, cardiomyocytes and embryonic fibroblasts simultaneously [185], for EC to provide vasculature and act synergically with cardiomyocytes to improve cell survival and proliferation.

Poly (epsilon-caprolactone) (PCL)

Poly(epsilon-caprolactone) is a FDA-approved biocompatible polyester, as PLA and PGA. It is more elastic because of its lower glass transition temperature, and behaves as a rubber at body temperature. Its degradation does not produce acidification because it occurs more slowly [158]. It has been proposed for myocardial regeneration for example in 3D constructs obtained by overlapping electrospun PCL nanofibrous mats (up to 5 layers) on which neonatal cardiomyocytes were cultured [186]. The layers established morphologic and electrical connections between them and exhibited synchronized beating, and no ischemia was found in the center of the constructs.

It is usually combined with PLA, PGA or its copolymer. Poly-glycolide-co-caprolactone (PGCL) biodegradable porous scaffolds have been studied as cell vehicles for bone marrow-derived mononuclear cells (BMMNC) in rat myocardial infarction models [187]. BMNC migrated from the scaffold and neovasculature over the implant was detected; left ventricular function improvement and limitation of the progression of the left ventricular dilation was also observed. Scaffolds made of poly(DL-lactide-co-caprolactone) (PLACL), PLGA, and type I collagen [158], cultured with neonatal rat heart cells, have been compared. The composite scaffolds gave better results than controls (collagen and PLGA sponges) in terms of cellularity, contractility and cardiac markers expression (Tn-I and Cx-43). Perfusion culture improved cell density distribution.

Polyurethanes (PU)

Polyurethanes are synthetic biocompatible materials widely used in the biomedical field. Their mechanical properties and biodegradability can be tuned by changing their composition. PU degrades in vivo through hydrolytic chain scission, which is accelerated by the enzymes action and loads, among other factors [188], but with the appropriate composition non-biodegradable polyurethanes can be obtained [189]. This family of polymers can be used to obtain fibrous scaffolds by electrospinning with different mechanical properties depending on the fibers orientation [190] or porous elastic scaffolds [191]. Polyester urethane urea (PEUU) elastic porous scaffolds have been implanted in sub-acute infarctions in rats and were found to promote the formation of smooth muscle bundles, to increase the ventricle thickness and to improve contractile function [192]. Cell attachment on polyurethane-based porous scaffolds can be improved by pre-treating them with laminin [193].

Poly(glycerol sebacate) (PGS)

Poly(glycerol sebacate) is a biocompatible and biodegradable elastomer capable of recovering from deformation. It can be obtained by polycondensation of glycerol and sebacic acid. By changing the synthesis temperature, the properties of the resulting material can be tuned to match the desired mechanical properties. The degradation rates can also be adjusted from fast degradation to nearly inert [194].

By the use of excimer laser microablation, 3D porous PGS scaffolds with anisotropic structural and mechanical properties were obtained [195, 130]. These scaffolds induced neonatal cardiac cells alignment in the absence of external stimuli and matched the mechanical properties of adult rat right ventricle. Moreover, they allowed cell contractility when stimulated. For its interesting mechanical properties, PGS has been coaxially electrospun with gelatin to form a nanofibrous mat with PGS in the core and gelatin in the shell [196] to enhance cell adhesion and proliferation. PGS has been modified to incorporate acrylic groups in different number (to modify its mechanical properties and degradation) and electrospun in combination with gelatin [197].

Acrylate based materials

Acrylate based materials have not been widely exploited for cardiac tissue engineering yet but the interest on them is increasing, for their versatility of processing and variety of properties obtained. For example, scaffolds made of poly(2-hydroxyethyl methacrylate-co-methacrylic acid) (P(HEMA-co-MAA) hydrogel have been obtained by fibers and microspheres templating to obtain spherical pores and parallel channels [198], which allow simultaneously mass transfer and guidance of the cardiomyocyte bundles. Mechanical properties were adjusted intentionally for the elastic modulus to be lower than that of native myocardium in order to make possible the mechanical stimulation of the cells when implanted in vivo. In [199], poly(ethyl acrylate) (PEA) scaffolds are filled with HA gel; the scaffolds provide the three-dimensional environment and mechanical properties and the gel may act as an encapsulating medium for the cells and may be also used as a medium for drug or growth factors release. RAD16-I gel may also be used as a filler in PEA scaffolds, where it acts as a diffusion medium and improves cell seeding efficiency (figure 2).

Figure 2.

(a) Scanning electron microscopy (SEM) image of poly(ethyl acrylate), PEA, elastomeric membranes with interconnected spherical pores. (b) CryoSEM image (cross section) of a PEA scaffold whose pores are filled with the self-assembling peptide (SAP) gel RAD16-I. (c) Adipose stem cells (nuclei stained in blue and actin cytoskeleton stained in green) seeded in a PEA scaffold with a SAP gel filling. Confocal laser scanning microscopy image of a 50 μm thick internal slice.

5.6. Electrical and mechanical stimulation

Electrical stimulation

External electrical fields have been shown to contribute to the differentiation towards cardiomyocytes of different cell types, such as embryonic stem cells (ESC) [200] or BMSC [201] seeded in collagen scaffolds, and to the development of conductive and contractile properties of neonatal cardiac cells, in this case seeded with Matrigel in a collagen porous scaffold [202]. It has been proposed that the intracellular endogenous reactive oxygen species (ROS) produced when an electric field is applied contribute to the hESC differentiation [203].

In an attempt of optimizing the electrical stimulation parameters [204], it has been determined that the electrode material is very important, and best results have been obtained for carbon electrodes. Amplitude and frequency of the stimulation have also a great influence in the cultured cardiac tissue. Micropatterned electrodes can be of interest as they allow spatial control of the electric field [205].

Polymeric scaffolds limit cardiomyocytes electric communication, what restricts the synchronous beating of the engineered tissue. To improve it, gold nanowires were incorporated to a porous alginate scaffold [206]. Another approach to obtain elastic and electrical conductive scaffolds consisted in impregnating thiol-HEMA/HEMA scaffolds with gold nanoparticles [207]. In both cases even without electrical stimulation the improvement in the scaffold conductivity had positive physiological effects.

Mechanical stimulation

Mechanical stress has a great impact on cell proliferation, ECM formation and hypertrophy (increased cell size), and has been intensively studied in the field of cardiac tissue engineering. Embryonic chick and neonatal rat cardiac myocytes mixed with collagen and mechanically stimulated exhibited hypertrophy and improvement of contractile function [208]. Cardiac myocytes from neonatal rats mixed with collagen I and Matrigel and casted in rings subjected to mechanical stretch [209] showed histological characteristics of adult cardiac tissue. Action potential measurements indicated electrophysiological behavior akin to cardiac tissue. Constructs produced by simultaneously electrospinning PU and electrospraying mesenchymal stem cells [210] were cultured in spinner flasks with stretching, which led to cells alignment, cardiac markers increase and ion channels development. Similarly, cells isolated from neonatal rat hearts seeded in chitosan-collagen I channeled porous scaffolds [211] and cultured under high mechanical stimulation induced cell alignment, elongation and the presence of gap junctions connecting the cells. Mechanical stress applied to human cardiac cells cultured in a gelatin scaffold improved cell distribution and proliferation within the scaffold, increased the production of the ECM, and the structure and organization was similar to normal myocardium, likely because the stretching of the scaffold favors nutrients and oxygen exchange improving cell microenvironment [212].

Advertisement

6. Ventricular restraints

After Chachques and Carpentier work [213], it was found that wrapping the heart even with a passive muscle flap had beneficial effects; this finding led to the development of the ventricular restraint therapy [214]. In this approach the aim is not to regenerate the ischemic tissue, but to avoid the progress of the adverse remodeling following a myocardial infarction. It is based on the application of a mechanical restraint (schematized in figure 1 c), which should limit or revert ventricular dilation. A variety of synthetic meshes have been proposed to achieve this goal.

A bilayer membrane with polypropylene in one side to promote tissue ingrowth (or at least limit the ventricular dilation) and with polytetrafluoroethylene in the other side to prevent pericardial adhesions was studied in a chronic infarction model of pig as a restraint [215]. The use of this patch induced improvements once the remodeling process following an infarction had started. The use of a non-biodegradable material is intentional as authors considered that a permanent mechanical reinforcement would be necessary to limit the remodeling.

To determine the extent at which a mechanical restraint is beneficial, a comparative study of two types of restrain was carried out in sheep: a patch over the infarct (non-biodegradable Marlex mesh) or a wrap (non-biodegradable Merseline mesh) [216]. The use of the mesh wrapping the ventricle reduced the remodeling whereas the patch applied over the infarct did not yield considerable improvements when compared with controls (untreated infarcted animals).

Paracor heartnet is a nitinol mesh proposed as a restrain device that is under clinical study in patients with severe dilated cardiomyopathy. In a study, six months after the implantation in 51 patients, results obtained suggested clinical benefits tending to reverse remodeling and that it could consequently be reliably implanted [217]. The PEERLESS-HF trial is the last carried out with this device so far [218]. It proved to be safe and improved patient’s quality of life and ventricular dilation; however, no improvement in the peak of VO2 was produced (which was an end-point of the trial), what led to stop enrollment in the trial. Nevertheless, a new clinical trial is planned. In another study in an animal model, it was shown that the heartnet can alter myocardial blood flow patterns in dilated cardiomyopathy, although it remains unclear if these changes are clinically relevant [219].

Another left ventricular restraint proposed is Acorn Corcap, a polyester mesh that is also being assessed in clinical trials after the positive results obtained in animal models [220]. 5 years after implantation it exhibited safety, a sustained reverse remodeling with a significant reduction in the left ventricular end diastolic volume and a slight increase in the sphericity index [221]. However, in an echocardiographic study using tissue velocity imaging, no improvement in cardiac output was achieved [222].

Limited results obtained with the ventricular restraint therapy can be, among other reasons, because of the absence of tissue regeneration. A more advanced approach combines the ventricular restraint therapy with a regenerative strategy such as patches or scaffolds loaded with cells. For instance, the Acorn Corcap and a collagen matrix loaded with MSC has been implanted in sheeps, and the combination was found to limit the fibrosis produced as foreign body reaction against the Corcap and improve the systolic and diastolic function [223].

Advertisement

7. Concluding remarks

Several therapeutic strategies have been proposed in the last decade to limit the adverse spread of the ischemic tissue and ventricle dilation or even to generate new myocardial tissue. These treatments consist in cellular therapy (so-called cellular cardiomyoplasty) where cells of different origin are implanted by different techniques onto the infarcted ventricle with the hope that cells will contribute to the generation of new contractile tissue to replace the scar, electrically coupled with the host myocardium. But despite the intense efforts and work put in the field, attempts so far have failed. Most of the implanted cells die soon after transplantation due to the fact that the cells cannot withstand the mechanical forces they experience in the host tissue. Mechanisms underlying the slight improvements observed are still undetermined; the paracrine effect is usually considered the way through which cells act, but the precise mechanisms are not completely understood yet. Besides, for this therapeutic approach to evolve to a realistic alternative to conventional treatments, some critical issues are still to be clarified: the way of delivery to maximize cell engraftment and minimize cell loss and death, the ideal cell type to be used, and the optimal time of cell administration (if they are implanted too soon, the inflammatory process kills the implanted cells, but if it is too late, the presence of the fibrotic scar limits their beneficial effects). New strategies already under study envision to improve cell survival by pre-conditioning the cells, pre-treating the host tissue or combining cells with other elements.

A possible way of localizing the appropriate cells in the target diseased tissue is to entrap them in a cell-friendly gelling biomaterial. Besides, gels can incorporate bioactive molecules for their controlled supply, and their preparation procedure (in the case of in situ gelling materials) avoids any invasive surgery. The injection of gelly materials alone onto the infarcted myocardium has shown some beneficial effects by itself and contributes somehow to limit the ventricular remodeling, for their slight role as mechanical support. Combining cells with gelly materials contributes, to some extent, to increase the cells residence time in the site of interest, and enhances cells adhesion and survival by providing them a better microenvironment. However, the consistency of these materials is generally too weak to withstand the synchronous contraction of the heart muscle without spreading from their target location, and their mechanical properties are too low to reach significant improvements in terms of containment of the dilated ventricle and post-infarct ventricular dysfunction.

Alternative tissue engineering strategies combine cells with three-dimensional scaffolds or patches to host them and improve their survival, induce the formation of new blood vessels and extracellular matrix and at the same time support the native tissue mechanically. The advantages of using myocardial patches or scaffolds are not only their usually superior mechanical properties, but also their wide versatility in terms of chemistries and morphology. There are many fabrication techniques for the preparation of scaffolds, leading to very different architectures, and these options are broadening with the computer-assisted techniques. Generally, positive results have been obtained by using scaffolds. In studies in which the therapeutic efficiency of a material was compared when used as an injectable gel or as a pre-fabricated scaffold or patch, the scaffold gave better results. When the scaffolds were loaded with growth factors or adhesion motives, in most of the cases the outcome was better. Mechanical and electrical stimulation are of help for cardiomyocytes to mature within the scaffolds and develop the characteristics and structures typical of cardiac tissue. Unfortunately, the implantation of epicardiac patches is much more invasive than that of injectable gels, and they need to be vascularized to ensure the success of the graft. Many attempts have addressed these questions but a satisfying solution has not been found yet.

Advertisement

Acknowledgments

The authors acknowledge the support of the FP7 NMP3-SL-2009-229239 project ‘‘Regeneration of cardiac tissue assisted by bioactive implants (RECATABI)’’.

References

  1. 1. Roger VL et al. Heart Disease and Stroke Statistics - 2012 Update A Report From the American Heart Association. Circulation 2012; 125: 2-220.
  2. 2. Vasan SV, Benjamin EJ, Sullivan LM, D’agostino RB. The burden of increasing worldwide cardiovascular disease. In: Fuster V, Walsh RA, O'Rourke RA, Poole-Wilson P (ed.) Hurst the Heart. 12th edition McGraw-Hill Professional; 2010 p17-46.
  3. 3. World Hearth Organitation. WHO: Programes and projects: Cardiovascular disease: The Atlas of Heart Disease and Stroke; 2004. http://www.who.int/cardiovascular_diseases/resources/atlas/en/ (accessed 03 June 2012)
  4. 4. Venugopal JR, Prabhakaran MP, Mukherjee S, Ravichandran R, Dan K, Ramakrishna S. Biomaterial Strategies for Alleviation of Myocardial Infarction. Journal of the Royal Society Interface 2012; 9(66): 1-19. doi:10.1098/rsif.2011.0301.
  5. 5. Walker CA, Spinale FG. The Structure and Function of the Cardiac Myocite: a Review of Fundamental Concepts. The Journal of Thoracic and Cardiovascular Surgery 1999; 118: 375-82.
  6. 6. Di Donato M, Toso A, Dor V, Sabatier M, Barletta G, Menicanti L, Fantini F and the RESTORE Group. Surgical Ventricular Restoration Improves Mechanical Intraventricular Dyssynchrony in Ischemic Cardiomyopathy. Circulation 2004; 109: 2536-43.
  7. 7. Smaill BH, LeGrice IJ, Hooks DA, Pullan AJ, Caldwell BJ, Hunter PJ. Cardiac Structure and Electrical Activation: Models and Measurement. Clinical and Experimental Pharmacology and Physiology 2004; 31 (12): 913-9.
  8. 8. Kocica MJ, Corno AF, Carreras-Costa F, Ballester-Rodes M, Moghbel MC, Cueva CNC, Lackovic V, Kanjuh V, Torrent-Guasp F. The Helical Ventricular Myocardial Band: Global, Three-Dimensional, Functional Architecture of the Ventricular Myocardium. European Journal Cardio-Thoracic Surgery 2006; 29: 21-40. DOI: 10.1016/j.ejcts.2006.03.011
  9. 9. LeGrice IJ, Smaill BH, Chai LZ, Edgar SG, Gavin JB, Hunter PJ. Laminar Structure of the Heart: Ventricular Myocyte Arrangement and Connective Tissue Architecture in the Dog. American Journal of Physiology 1995; 269: H571-82.
  10. 10. Chen FY, Cohn LH. The Surgical Treatment of Heart Failure. A New Frontier: Nontransplant Surgical Alternatives in Heart Failure. Cardiology in Review 2002; 10(6): 326-33.
  11. 11. Hoyt RH, Cohen ML, Saffitz JE. Distribution and Three-Dimensional Structure of Intercellular Junctions in Canine Myocardium. Circulation Research 1989; 64: 563-74.
  12. 12. Spach MS, Heidlage JF. The Stochastic Nature of Cardiac Propagation at a Microscopic Level. Electrical Description of Myocardial Architecture and its Application to Conduction. Circulation Research 1995; 76: 366-80.
  13. 13. Severs NJ. The Cardiac Muscle Cell. BioEssays 2000; 22: 188-199.
  14. 14. Burke AP, Virmani R. Pathology of myocardial ischemia, infarction, reperfusion and sudden death. In: Fuster V, Walsh RA, O'Rourke RA, Poole-Wilson P (ed.) Hurst the Heart. 12th edition McGraw-Hill Professional; 2010. p1321-1338.
  15. 15. Baig MK, Mahon N, McKenna WJ, Caforio ALP, Bonow RO, Francis GS, Gheorghiade M. The Pathophysiology of Advanced Heart Failure. Heart & Lung 1999; 28(2): 87-101.
  16. 16. Ferrero JM Jr, Trénor B, Montilla F, Saiz J, Ferrero Á, Rodriguez B. Ischemia. In: Wiley Encyclopedia of Biomedical Engineering. (ed.) John Wiley & Sons, Inc; 2006. p1-17.
  17. 17. Douglas JS Jr, King SB III.Percutaneous coronary intervention. In:. Fuster V, Walsh RA, O'Rourke RA, Poole-Wilson P (ed.) Hurst the Heart. 12th edition McGraw-Hill Professional; 2010. p1427-1457.
  18. 18. Lally C. Kelly DJ, Prendergast PJ. Stents. In: Wiley Encyclopedia of Biomedical Engineering. (ed.) John Wiley & Sons, Inc; 2006. p1-10.
  19. 19. Stefanini GG, Kalesan B, Serruys PW, Heg D, Buszman P, Linke A, Ischinger T, Klauss V, Eberli F, Wijns W, Morice MC, Di Mario C, Corti R, Antoni D, Sohn HY, Eerdmans P, van Es GA, Meier B, Windecker S, Jüni P. Long-term clinical outcomes of biodegradable polymer biolimus-eluting stents versus durable polymer sirolimus-eluting stents in patients with coronary artery disease (LEADERS): 4 year follow-up of a randomised non-inferiority trial. Lancet 2011; 378: 1940-8.
  20. 20. Ruwende C, Visovatti S, Pinsky DJ. Molecular and cellular mechanisms of myocardial ischemia-reperfusion injury. In: Fuster V, Walsh RA, O'Rourke RA, Poole-Wilson P (ed.) Hurst the Heart. 12th edition McGraw-Hill Professional; 2010. p1339-1350.
  21. 21. Nian M, Lee P, Khaper N, Liu P. Inflammatory Cytokines and Postmyocardial Infarction Remodeling. Circulation Research 2004; 94: 1543-1553.
  22. 22. Sun Y, Kiani MF, Postlethwaite AE, Weber KT. Infarct Scar as Living Tissue. Basic Research in Cardiology 2002; 97: 343-347. doi: 10.1007/s00395-002-0365-8.
  23. 23. Christman KL, Lee RJ. Biomaterials For the Treatment of Myocardial Infarction. Journal American College of Cardiology 2006; 48: 907-13.
  24. 24. Mann DL. Mechanisms and Models in Heart Failure: a Combinatorial Approach. Circulation 1999; 100: 999-1008. DOI: 10.1161/01.CIR.100.9.999.
  25. 25. Cortes-Morichetti M, Frati G, Schussler O, Duong Van Huyen JP, Lauret E, Genovese JA, Carpentier AF, Chachques JC. Association Between a Cell-Seeded Collagen Matrix and Cellular Cardiomyoplasty for Myocardial Support and Regeneration. Tissue engineering 2007; 13(11): 2681-2687. doi: 10.1089/ten.2006.0447.
  26. 26. Jawad H, Ali NN, Lyon AR, Chen QZ, Harding SE, Boccaccini AR. Myocardial Tissue Engineering: a Review. Journal of Tissue Engineering and Regenerative Medicine 2007; 1: 327–342.
  27. 27. Nelson DM, Mab Z, Fujimoto KL, Hashizume R, Wagner WR. Intra-Myocardial Biomaterial Injection Therapy in the Treatment of Heart Failure: Materials, Outcomes and Challenges. Acta Biomaterialia 2011; 7: 1-15.
  28. 28. Chachques JC, Salanson-Lajos C, Lajos P, Shafy A, Alshamry A, Carpentier A. Cellular Cardiomyoplasty for Myocardial Regeneration. Asian Cardiovascular & Thoracic Annals 2005; 13: 287-296.
  29. 29. Chen QZ, Harding SE, Ali NN, Lyon AR, Boccaccini AR. Biomaterials in Cardiac Tissue Engineering: Ten Years of Research Survey. Materials Science and Engineering 2008; 59: 1-37.
  30. 30. Anversa P, Leri A, Kajstura J, Nadal-Ginard B. Myocyte Growth and Cardiac Repair. Journal of Molecular and Cellular Cardiology 2002; 34: 91-105.
  31. 31. Bergmann O, Bhardwaj RD, Bernard S, Zdunek S, Barnabe-Heider F, Walsh S, Zupicich J, Alkass K, Buchholz BA, Druid H, Jovinge S, Frisén J. Evidence for Cardiomyocyte Renewal in Humans. Science 2009; 324(5923): 98-102.
  32. 32. Wang F, Guan J. Cellular Cardiomyoplasty and Cardiac Tissue Engineering for Myocardial Therapy. Advanced Drug Delivery Reviews 2010; 62: 784–797.
  33. 33. Chachques JC, Grandjean PA, Tommasi JJ, Perier P, Chauvaud S, Bourgeois I, Carpentier A. Dynamic Cardiomyoplasty: A New Approach to Assist Chronic Myocardial Failure. Life Support System 1987; 5(4): 323-7.
  34. 34. Chachques JC. Development of Bioartificial Myocardium Using Stem Cells and Nanobiotechnology Templates. Cardiology Research and Practice 2011; 2011: 806795. doi:10.4061/2011/806795.
  35. 35. Wu J, Zeng F, Weisel RD, Li RK. Stem Cells for Cardiac Regeneration by Cell Therapy and Myocardial Tissue Engineering. Advances in Biochemical Engineering/Biotechnology 2009; 114: 107-128. doi: 10.1007/10_2008_37.
  36. 36. Pendyala L, Goodchild T, Gadesam RR, Chen J, Robinson K, Chronos N, Hou D. Cellular cardiomyoplasty and cardiac regeneration. Current Cardiology Reviews 2008; 4: 72-80.
  37. 37. Leor J, Amsalem Y, Cohen S. Cells, scaffolds, and molecules for myocardial tissue engineering. Pharmacology and Therapeutics 2005; 105(2): 151-63.
  38. 38. Zhou R, Acton PD, Ferrari VA. Imaging stem cells implanted in infarcted myocardium. Journal American college of cardiology 2006; 48(10): 2094-2106.
  39. 39. Hofmann M, Wollert KC, Meyer GP, Menke A, Arseniev L, Hertenstein B, Ganser A, Knapp WH, Drexler H. Monitoring of Bone Marrow Cell Homing into the Infarcted Human Myocardium. Circulation 2005; 111: 2198-202.
  40. 40. Teng CJ, Luo J, Chiu RC, Shum-Tim D. Massive Mechanical Loss of Microspheres with Direct Intramyocardial Injection in the Beating Heart: Implications for Cellular Cardiomyoplasty. Journal of Thoracic and Cardiovascular Surgery 2006; 132(3): 628-32. doi: 10.1016/j.jtcvs.2006.05.034.
  41. 41. Schussler O, Chachques JC, Mesana TG, Suuronen EJ, Lecarpentier Y, Ruel M. 3-Dimensional structures to enhance cell therapy and engineer contractile tissue. Asian Cardiovascular & thoracic annals 2010; 18(2): 188-198.
  42. 42. Zenovich AG, Davis BH, Taylor DA. Comparison of Intracardiac Cell Transplantation: Autologous Skeletal Myoblasts Versus Bone Marrow Cells. Handbook of Experimental Pharmacology 2007; 180: 117–165.
  43. 43. Forte E, Chimenti I, Barile L, Gaetani R, Angelini F, Ionta V, Messina E, Giacomello A. Cardiac Cell Therapy: The Next (Re)Generation. Stem Cell Reviews and Reports 2011; 7(4): 1018-1030. doi:10.1007/s12015-011-9252-8.
  44. 44. Qian H, Yang Y, Huang J, Dou K, Yang G. Cellular cardiomyoplasty by catheter-based infusion of stem cells in clinical settings. Transplant Immunology 2006; 16: 135-147.
  45. 45. Laflamme MA, Chen KY, Naumova AV, Muskheli V, Fugate JA, Dupras SK, Reinecke H, Xu C, Hassanipour M, Police S, O'Sullivan C, Collins L, Chen Y, Minami E, Gill EA, Ueno S, Yuan C, Gold J, Murry CE. Cardiomyocytes derived from human embryonic stem cells in pro-survival factors enhance function of infarcted rat hearts. Nature Biotechnology 2007; 25(9): 1015-24. doi:10.1038/nbt1327.
  46. 46. Ebelt H, Jungblut M, Zhang Y, Kubin T, Kostin S, Technau A, Oustanina S, Niebrügge S, Lechmann J, Werdan K, Braun T. Cellular cardiomyoplasty: improvement of left ventricular function correlates with the release of cardioactive cytokines. Stem cells 2007; 25(1): 236-244.
  47. 47. Zimmet JM, Hare JM. Emerging role for bone marrow derived mesenchymal stem cells in myocardial regenerative therapy. Basic Research in Cardiology 2005; 100(6): 471-481. doi: 10.1007/s00395-005-0553-4.
  48. 48. Madonna R, Geng YJ, De Caterina R. Adipose tissue-derived stem cells: characterization and potential for cardiovascular repair. Arteriosclerosis, Thrombosis and Vascular Biology 2009; 29(11): 1723-9.
  49. 49. Heng BC, Haider HK, Sim EK, Cao T, Ng SC. Strategies for Directing the Differentiation of Stem Cells into the Cardiomyogenic Lineage in Vitro. Cardiovascular Research 2004; 62(1): 34–42.
  50. 50. Vulliet PR, Greeley M, Halloran SM, MacDonald KA, Kittleson MD. Intra-Coronary Arterial Injection of Mesenchymal Stromal Cells and Microinfarction in Dogs. The Lancet 2004; 363(9411): 783-4.
  51. 51. Meyer GP, Wollert KC, Lotz J, Steffens J, Lippolt P, Fichtner S, Hecker H, Schaefer A, Arseniev L, Hertenstein B, Ganser A, Drexler H. Intracoronary Bone Marrow Cell Transfer After Myocardial Infarction: Eighteen Months Follow-up Data from the Randomized, Controlled BOOST (BOne marrOw transfer to enhance ST-elevation infarct regeneration) Trial. Circulation 2006; 113: 1287–94, doi: 10.1161/CIRCULATIONAHA.105.575118.
  52. 52. Hahn JY, Cho HJ, Kang HJ, Kim TS, Kim MH, Chung JH, Bae JW, Oh BH, Park YB, Kim HS. Pre-treatment of Mesenchymal Stem Cells with a Combination of Growth Factors Enhances Gap Junction Formation, Cytoprotective Effect on Cardiomyocytes, and Therapeutic Efficacy for Myocardial Infarction. Journal of the American College of Cardiology 2008; 51(9): 933-43.
  53. 53. Miyahara Y, Nagaya N, Kataoka M, Yanagawa B, Tanaka K, Hao H, Ishino K, Ishida H, Shimizu T, Kangawa K, Sano S, Okano T, Kitamura S, Mori H. Monolayered mesenchymal stem cells repair scarred myocardium after myocardial infarction. Nature Medicine 2006; 12(4): 459-465.
  54. 54. www.clinicaltrials.gov (accessed 13 November 2012).
  55. 55. REgeneration of CArdiac Tissue Assisted by Bioactive Implants. funded by the European Comission under the 7th FP, www.recatabi.com.
  56. 56. Messina E, De Angelis L, Frati G, Morrone S, Chimenti S, Fiordaliso F, Salio M, Battaglia M, Latronico MV, Coletta M, Vivarelli E, Frati L, Cossu G, Giacomello A. Isolation and Expansion of Adult Cardiac Stem Cells From Human and Murine Heart. Circulation Research 2004; 95: 911-921.
  57. 57. Smith RR, Barile L, Cho HC, Leppo MK, Hare JM, Messina E, Giacomello A, Abraham MR, Marbán E. Regenerative potential of cardiosphere-derived cells expanded from percutaneous endomyocardial biopsy specimens. Circulation 2007; 115(7): 896-908.
  58. 58. Makkar RR, Smith RR, Cheng K, Malliaras K, Thomson LEJ, Berman D, Czer LSC, Marbán L, Mendizabal A, Johnston PV, Russell SD, Schuleri KH, Lardo AC, Gerstenblith G, Marbán E. Intracoronary cardiosphere-derived cells for heart regeneration after myocardial infarction (CADUCEUS): a prospective, randomised phase 1 trial. The Lancet 2012; 379: 895-904. doi: 10.1016/S0140- 6736(12)60195-0.
  59. 59. Shafy A, Lavergne T, Latremouille C, Cortes-Morichetti M, Carpentier A, Chachques JC. Association of electrostimulation with cell transplantation in ischemic heart disease. Journal of Thoracic and Cardiovascular Surgery 2009; 138(4): 994-1001.
  60. 60. Cleland JG, Coletta AP, Abdellah AT, Nasirb M, Hobsonb N, Freemantlec N, Clarka AL. Clinical Trials Update from the American Heart Association 2006: OAT, SALT 1 and 2, MAGIC, ABCD, PABA-CHF, IMPROVECHF, and Percutaneous Mitral Annuloplasty. European Journal of Heart Failure 2007; 9: 92-7.
  61. 61. Hirata Y, Sata M, Motomura N, Takanashi M, Suematsu Y, Ono M, Takamoto S. Human umbilical cord blood cells improve cardiac function after myocardial infarction. Biochemical and Biophysical Research Communications 2005; 327(2): 609-14. doi:10.1016/j.bbrc.2004.12.044.
  62. 62. Walther G, Gekas J, Bertrand OF. Amniotic Stem Cells for Cellular Cardiomyoplasty: Promises and Premises. Catheterization and Cardiovascular Interventions 2009; 73(7): 917–924.
  63. 63. Yeh YC, Wei HJ, Lee WY, Yu CL, Chang Y, Hsu LW, Chung MF, Tsai MS, Hwang SM, Sung HW. Cellular Cardiomyoplasty with Human Amniotic Fluid Stem Cells: In Vitro and In Vivo Studies. Tissue Engineering Part A 2010; 16(6): 1925-36.
  64. 64. Shimizu T, Yamato M, Kikuchi A, Okano T. Cell sheet engineering for myocardial tissue reconstruction. Biomaterials 2003; 24(13): 2309-2316.
  65. 65. Alcon A, Cagavi Bozkulak E, Qyang Y. Regenerating functional heart tissue for myocardial repair. Cellular and Molecular Life Sciences 2012; 69(16): 2635-56. doi:10.1007/s00018-012-0942.
  66. 66. Yeh YC, Lee WY, Yu CL, Hwang SM, Chung MF, Hsu LW, Chang Y, Lin WW, Tsai MS, Wei HJ, Sung HW. Cardiac repair with injectable cell sheet fragments of human amniotic fluid stem cells in an immune-suppressed rat model. Biomaterials 2010; 31(25): 6444-53.
  67. 67. Furuta A, Miyoshi S, Itabashi Y, Shimizu T, Kira S, Hayakawa K, Nishiyama N, Tanimoto K, Hagiwara Y, Satoh T, Fukuda K, Okano T, Ogawa S. Pulsatile cardiac tissue grafts using a novel three-dimensional cell sheet manipulation technique functionally integrates with the host heart, in vivo. Circulation Research 2006; 98(5): 705–712.
  68. 68. Williams C, Xie AW, Yamato M, Okano T, Wong JY. Stacking of aligned cell sheets for layer-by-layer control of complex tissue Structure. Biomaterials 2011; 32(24): 5625-32.
  69. 69. Vunjak-Novakovic G, Lui KO, Tandon N, Chien KR. Bioengineering Heart Muscle: A Paradigm for Regenerative Medicine. Annual Reviewof Biomedical Engineering 2011; 13: 245–67. doi: 10.1146/annurev-bioeng-071910-124701.
  70. 70. Haraguchi Y, Shimizu T, Yamato M, Kikuchi A, Okano T. Electrical coupling of cardiomyocyte sheets occurs rapidly via functional gap junction formation. Biomaterials 2006; 27(27): 4765–4774.
  71. 71. Shimizu T, Sekine H, Yang J, Isoi Y, Yamato M, Kikuchi A, Kobayashi E, Okano T. Polysurgery of cell sheet grafts overcomes diffusion limits to produce thick, vascularized myocardial tissues. The Journal of the Federation of American Societies for Experimental Biology 2006; 20(6): 708–710.
  72. 72. Lee WY, Wei HJ, Lin WW, Yeh YC, Hwang SM, Wang JJ, Tsai MS, Chang Y, Sung HW. Enhancement of cell retention and functional benefits in myocardial infarction using human amniotic-fluid stem-cell bodies enriched with endogenous ECM. Biomaterials 2011; 32(24): 5558-67.
  73. 73. Ye Z, Zhou Y, Cai H, Tan W. Myocardial regeneration: Roles of stem cells and hydrogels. Advanced Drug Delivery Reviews 2011; 63(8): 688-97.
  74. 74. Habib M, Shapira-Schweitzer K, Caspi O, Gepstein A, Arbel G, Aronson D, Seliktar D, Gepstein L. A combined cell therapy and in-situ tissue-engineering approach for myocardial repair. Biomaterials 2011; 32(30): 7514-23.
  75. 75. Wall ST, Walker JC, Healy KE, Ratcliffe MB, Guccione JM. Theoretical impact of the injection of material into the myocardium: a finite element model simulation. Circulation 2006; 114(24): 2627-35.
  76. 76. Shen X, Tanaka K, Takamori A. Coronary Arteries Angiogenesis in Ischemic Myocardium: Biocompatibility and Biodegradability of Various Hydrogels. Artificial Organs 2009; 33(10): 781-7. doi: 10.1111/j.1525-1594.2009.00815.x.
  77. 77. Guo HD, Wang HJ, Tan YZ, Wu JH. Transplantation of marrow derived cardiac stem cells carried in fibrin improves cardiac function after myocardial infarction. Tissue Engineering Part A 2011; 17(1-2): 45-58.
  78. 78. Christman KL, Vardanian AJ, Fang Q, Sievers RE, Fok HH, Lee RJ. Injectable fibrin scaffold improves cell transplant survival, reduces infarct expansion, and induces neovasculature formation in ischemic myocardium. Journal of the American College of Cardiology 2004; 44(3): 654-60.
  79. 79. Barsotti MC, Felice F, Balbarini A, Di Stefano R. Fibrin as a scaffold for cardiac tissue Engineering. Biotechnology and Applied Biochemistry 2011; 58(5): 301-10. doi: 10.1002/bab.49.
  80. 80. Christman KL, Fang Q, Yee MS, Johnson KR, Sievers RE, Lee RJ. Enhanced neovasculature formation in ischemic myocardium following delivery of pleiotrophin plasmid in a biopolymer. Biomaterials 2005; 26(10): 1139-44.
  81. 81. Martens TP, Godier AF, Parks JJ, Wan LQ, Koeckert MS, Eng GM, Hudson BI, Sherman W, Vunjak-Novakovic G. Percutaneous cell delivery into the heart using hydrogels polymerizing in situ. Cell Transplantation 2009; 18(3): 297-304.
  82. 82. Christman KL, Fok HH, Sievers RE, Fang Q, Lee RJ. Fibrin glue alone and skeletal myoblasts in a fibrin scaffold preserve cardiac function after myocardial infarction. Tissue Engineering 2004; 10 (3-4): 403-9.
  83. 83. Ryu JH, Kim IK, Cho SW, Cho MC, Hwang KK, Piao H, Piao S, Lim SH, Hong YS, Choi CY, Yoo KJ, Kim BS. Implantation of bone marrow mononuclear cells using injectable fibrin matrix enhances neovascularization in infarcted myocardium. Biomaterials 2005; 26(3): 319–326.
  84. 84. Chekanov V, Akhtar M, Tchekanov G, Dangas G, Shehzad MZ, Tio F, Adamian M, Colombo A, Roubin G, Leon MB, Moses JW, Kipshidze NN. Transplantation of autologous endothelial cells induces angiogenesis. Pacing and Clinical Electrophysiology 2003; 26(1 Pt 2): 496-9.
  85. 85. Chenite A, Chaput C, Wang D, Combes C, Buschmann MD, Hoemann CD, Leroux JC, Atkinson BL, Binette F, Selmani A. Novel injectable neutral solutions of chitosan form biodegradable gels in situ. Biomaterials 2000; 21(21): 2155-61.
  86. 86. Reis LA, Chiu LL, Liang Y, Hyunh K, Momen A, Radisic M. A peptide-modified chitosan–collagen hydrogel for cardiac cell culture and delivery. Acta Biomaterialia 2012; 8(3): 1022-36.
  87. 87. Liu Z, Wang H, Wang Y, Lin Q, Yao A, Cao F, Li D, Zhou J, Duan C, Du Z, Wang Y, Wang C. The influence of chitosan hydrogel on stem cell engraftment, survival and homing in the ischemic myocardial microenvironment. Biomaterials 2012; 33(11): 3093-106.
  88. 88. Binsalamah ZM, Paul A, Khan AA, Prakash S, Shum-Tim D. Intramyocardial sustained delivery of placental growth factor using nanoparticles as a vehicle for delivery in the rat infarct model. International Journal of Nanomedicine 2011; 6: 2667-78.
  89. 89. Kofidis T, Lebl DR, Martinez EC, Hoyt G, Tanaka M, Robbins RC. Novel injectable bioartificial tissue facilitates targeted, less invasive, large-scale tissue restoration on the beating heart after myocardial injury. Circulation 2005; 112(9 Suppl): I173-7.
  90. 90. Zhang P, Zhang H, Wang H, Wei Y, Hu S. Artificial matrix helps neonatal cardiomyocytes restore injured myocardium in rats. Artificial Organs 2006; 30(2): 86-93.
  91. 91. Kofidis T, de Bruin JL, Hoyt G, Lebl DR, Tanaka M, Yamane T, Chang CP, Robbins RC. Injectable bioartificial myocardial tissue for large-scale intramural cell transfer and functional recovery of injured heart muscle. The Journal of Thoracic and Cardiovascular Surgery 2004; 128(4): 571-8.
  92. 92. Shen D, Wang X, Zhang L, Zhao X, Li J, Cheng K, Zhang J. The amelioration of cardiac dysfunction after myocardial infarction by the injection of keratin biomaterials derived from human hair. Biomaterials 2011; 32(35): 9290-9.
  93. 93. Landa N, Miller L, Feinberg MS, Holbova R, Shachar M, Freeman I, Cohen S, Leor J. Effect of injectable alginate implant on cardiac remodeling and function after recent and old infarcts in rat. Circulation 2008; 117(11): 1388-96.
  94. 94. Ruvinov E, Leor J, Cohen S. The promotion of myocardial repair by the sequential delivery of IGF-1 and HGF from an injectable alginate biomaterial in a model of acute myocardial infarction. Biomaterials 2011; 32(2): 565-78.
  95. 95. Rowley JA, Madlambayan G, Mooney DJ. Alginate hydrogels as synthetic extracellular matrix materials. Biomaterials 1999; 20(1): 45-53.
  96. 96. Yu J, Gu Y, Du KT, Mihardja S, Sievers RE, Lee RJ. The effect of injected RGD modified alginate on angiogenesis and left ventricular function in a chronic rat infarct model. Biomaterials 2009; 30(5): 751-6.
  97. 97. Tsur-Gang O, Ruvinov E, Landa N, Holbova R, Feinberg MS, Leor J, Cohen S. The effects of peptide-based modification of alginate on left ventricular remodeling and function after myocardial infarction. Biomaterials 2009; 30(2): 189-95.
  98. 98. Mihardja SS, Sievers RE, Lee RJ. The effect of polypyrrole on arteriogenesis in an acute rat infarct model. Biomaterials 2008; 29(31): 4205–10.
  99. 99. Gaffney J, Matou-Nasri S, Grau-Olivares M, Slevin M. Therapeutic applications of hyaluronan. Molecular BioSystems 2010; 6(3): 437–443. doi: 10.1039/b910552m.
  100. 100. Yoon SJ, Fang YH, Lim CH, Kim BS, Son HS, Park Y, Sun K. Regeneration of ischemic heart using hyaluronic acid-based injectable hydrogel. Journal of Biomedical Materials Research Part B: Applied Biomaterials 2009; 91(1): 163-71.
  101. 101. Cheng K, Blusztajn A, Shen D, Li TS, Sun B, Galang G, Zarembinski TI, Prestwich GD, Marbán E, Smith RR, Marbán L. Functional performance of human cardiosphere-derived cells delivered in an in situ polymerizable hyaluronan-gelatin hydrogel. Biomaterials 2012; 33(21): 5317-24.
  102. 102. Duan Y, Liu Z, O'Neill J, Wan LQ, Freytes DO, Vunjak-Novakovic G. Hybrid gel composed of native heart matrix and collagen induces cardiac differentiation of human embryonic stem cells without supplemental growth factors. Journal of Cardiovascular Translational Research 2011; 4(5): 605-15.
  103. 103. Dai W, Wold LE, Dow JS, Kloner RA. Thickening of the infarcted wall by collagen injection improves left ventricular function in rats: a novel approach to preserve cardiac function after myocardial infarction. Journal of the American College of Cardiology 2005; 46(4): 714-9.
  104. 104. Huang NF, Yu J, Sievers R, Li S, Lee RJ. Injectable biopolymers enhance angiogenesis after myocardial infarction. Tissue Engineering 2005; 11(11-12): 1860-6.
  105. 105. Thompson CA, Nasseri BA, Makower J, Houser S, McGarry M, Lamson T, Pomerantseva I, Chang JY, Gold HK, Vacanti JP, Oesterle SN. Percutaneous transvenous cellular cardiomyoplasty. A novel nonsurgical approach for myocardial cell transplantation. Journal of the American College of Cardiology 2003; 41(11): 1964-71.
  106. 106. Dai W, Hale SL, Kay GL, Jyrala AJ, Kloner RA. Delivering stem cells to the heart in a collagen matrix reduces relocation of cells to other organs as assessed by nanoparticle technology. Regenerative Medicine 2009; 4(3): 387-95.
  107. 107. Suuronen EJ, Veinot JP, Wong S, Kapila V, Price J, Griffith M, Mesana TG, Ruel M. Tissue-engineered injectable collagen-based matrices for improved cell delivery and vascularization of ischemic tissue using CD133+ progenitors expanded from the peripheral blood. Circulation 2006; 114(1 Suppl): I138-44.
  108. 108. Zhang F, He C, Cao L, Feng W, Wang H, Mo X, Wang J. Fabrication of gelatin–hyaluronic acid hybrid scaffolds with tunable porous structures for soft tissue engineering. International Journal of Biological Macromolecules 2011; 48(3): 474-81.
  109. 109. Shao ZQ, Takaji K, Katayama Y, Kunitomo R, Sakaguchi H, Lai ZF, Kawasuji M. Effects of intramyocardial administration of slow-release basic fibroblast growth factor on angiogenesis and ventricular remodeling in a rat infarct model. Circulation Journal 2006; 70(4): 471-7.
  110. 110. Iwakura A, Fujita M, Kataoka K, Tambara K, Sakakibara Y, Komeda M, Tabata Y. Intramyocardial sustained delivery basic fibroblast growth factor improves angiogenesis and ventricular function in a rat infarct model. Heart Vessels 2003; 18: 93–9.
  111. 111. Singelyn JM, DeQuach JA, Seif-Naraghi SB, Littlefield RB, Schup-Magoffin PJ, Christman KL. Naturally derived myocardial matrix as an injectable scaffold for cardiac tissue engineering. Biomaterials 2009; 30(29): 5409-16.
  112. 112. Okada M, Payne TR, Oshima H, Momoi N, Tobita K, Huard J. Differential efficacy of gels derived from small intestinal submucosa as an injectable biomaterial for myocardial infarct repair. Biomaterials 2010; 31(30): 7678-83.
  113. 113. Zhao ZQ, Puskas JD, Xu D, Wang NP, Mosunjac M, Guyton RA, Vinten-Johansen J, Matheny R. Improvement in cardiac function with small intestine extracellular matrix is associated with recruitment of C-kit cells, myofibroblasts, and macrophages after myocardial infarction. Journal of the American College of Cardiology 2010; 55(12): 1250-61.
  114. 114. Jeong B, Kim SW, Bae YH. Thermosensitive sol-gel reversible hydrogels. Advanced Drug Delivery Reviews 2002; 54(1): 37-51.
  115. 115. Fujimoto KL, Ma Z, Nelson DM, Hashizume R, Guan J, Tobita K, Wagner WR. Synthesis, characterization and therapeutic efficacy of a biodegradable, thermoresponsive hydrogel designed for application in chronic infarcted myocardium. Biomaterials 2009; 30(26): 4357-68.
  116. 116. Li Z, Guo X, Matsushita S, Guan J. Differentiation of cardiosphere-derived cells into a mature cardiac lineage using biodegradable poly(N-isopropylacrylamide) hydrogels. Biomaterials 2011; 32(12): 3220-32.
  117. 117. Wang T, Wu DQ, Jiang XJ, Zhang XZ, Li XY, Zhang JF, Zheng ZB, Zhuo R, Jiang H, Huang C. Novel thermosensitive hydrogel injection inhibits post-infarct ventricle remodelling. European Journal of Heart Failure 2009; 11(1): 14-9.
  118. 118. Dobner S, Bezuidenhout D, Govender P, Zilla P, Davies N. Asynthetic nondegradable polyethylene glycol hydrogel retards adverse post-infarct left ventricular remodeling. Journal of Cardiac Failure 2009; 15(7): 629-36.
  119. 119. Wang T, Jiang XJ, Tang QZ, Li XY, Lin T, Wu DQ, Zhang XZ, Okello E. Bone marrow stem cells implantation with a-cyclodextrin/MPEG– PCL–MPEG hydrogel improves cardiac function after myocardial infarction. Acta Biomaterialia 2009; 5(8): 2939-44.
  120. 120. Wu J, Zeng F, Huang XP, Chung JC, Konecny F, Weisel RD, Li RK. Infarct stabilization and cardiac repair with a VEGF-conjugated, injectable Hydrogel. Biomaterials 2011; 32(2): 579-86.
  121. 121. Kraehenbuehl TP, Ferreira LS, Hayward AM, Nahrendorf M, van der Vlies AJ, Vasile E, Weissleder R, Langer R, Hubbell JA. Human embryonic stem cell-derived microvascular grafts for cardiac tissue preservation after myocardial infarction. Biomaterials 2011; 32(4): 1102-9.
  122. 122. Wang T, Jiang XJ, Lin T, Ren S, Li XY, Zhang XZ, Tang QZ. The inhibition of postinfarct ventricle remodeling without polycythaemia following local sustained intramyocardial delivery of erythropoietin within a supramolecular hydrogel. Biomaterials 2009; 30(25): 4161-7.
  123. 123. Dvir T, Bauer M, Schroeder A, Tsui JH, Anderson DG, Langer R, Liao R, Kohane DS. Nanoparticles targeting the infarcted heart. Nano Letters 2011; 11(10): 4411-4. doi: 10.1021/nl2025882.
  124. 124. Davis ME, Motion JP, Narmoneva DA, Takahashi T, Hakuno D, Kamm RD, Zhang S, Lee RT. Injectable self-assembling peptide nanofibers create intramyocardial microenvironments for endothelial cells.Circulation 2005; 111(4): 442-50.
  125. 125. Tokunaga M, Liu ML, Nagai T, Iwanaga K, Matsuura K, Takahashi T, Kanda M, Kondo N, Wang P, Naito AT, Komuro I. Implantation of cardiac progenitor cells using self-assembling peptide improves cardiac function after myocardial infarction. Journal of Molecular and Cellular Cardiology 2010; 49(6): 972-83.
  126. 126. Kim JH, Jung Y, Kim SH, Sun K, Choi J, Kim HC, Park Y, Kim SH. The enhancement of mature vessel formation and cardiac function in infarcted hearts using dual growth factor delivery with self-assembling peptides. Biomaterials 2011; 32(26): 6080-8.
  127. 127. Davis ME, Hsieh PC, Takahashi T, Song Q, Zhang S, Kamm RD, Grodzinsky AJ, Anversa P, Lee RT. Local myocardial insulin-like growth factor 1 (IGF-1) delivery with biotinylated peptide nanofibers improves cell therapy for myocardial infarction. Proceedings of the National Academy of Sciences of the United States of America 2006; 103(21): 8155-60.
  128. 128. Nerem RM. The challenge of imitating nature. In Lanza R, Langer R, Vacanti J.. Principles of tissue engineering. San Diego (Ca) USA: Academic press; 1997 p.9-15.
  129. 129. Jawad H, Lyon AR, Harding SE, Ali NN, Boccaccini AR. Myocardial tissue engineering. British Medical Bulletin 2008; 87: 31-47.
  130. 130. Engelmayr GC Jr, Cheng M, Bettinger CJ, Borenstein JT, Langer R, Freed LE. Accordion-like honeycombs for tissue engineering of cardiac anisotropy. Nature Materials 2008; 7: 1003–10.
  131. 131. Bhana B, Iyer RK, Chen WL, Zhao R, Sider KL, Likhitpanichkul M, Simmons CA, Radisic M. Influence of substrate stiffness on the phenotype of heart cells. Biotechnology and Bioengineering 2010; 105(6): 1148-60.
  132. 132. Marsano A, Maidhof R, Wan LQ, Wang Y, Gao J, Tandon N, Vunjak-Novakovic G. Scaffold stiffness affects the contractile function of three-dimensional engineered cardiac constructs. Biotechnology Progress 2010; 26(5): 1382-90.
  133. 133. Young JL, Engler AJ. Hydrogels with time-dependent material properties enhance cardiomyocyte differentiation in vitro. Biomaterials 2011; 32(4): 1002-9.
  134. 134. Carrier RL, Rupnick M, Langer R, Schoen FJ, Freed LE, Vunjak-Novakovic G. Perfusion Improves Tissue Architecture of Engineered Cardiac Muscle. Tissue Engineering 2002; 8(2): 175-88.
  135. 135. Radisic M, Malda J, Epping E, Geng W, Langer R, Vunjak-Novakovic G. Oxygen gradients correlate with cell density and cell viability in engineered cardiac tissue. Biotechnology and Bioengineering 2006; 93(2): 332-43.
  136. 136. Radisic M, Park H, Chen F, Salazar-Lazzaro JE, Wang Y, Dennis R, Langer R, Freed LE, Vunjak-Novakovic G. Biomimetic approach to cardiac tissue engineering: oxygen carriers and channeled scaffolds. Tissue Engineering 2006; 12(8): 2077-91.
  137. 137. 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. Journal of Biomedical Materials Research Part A 2003; 65(4): 489-97.
  138. 138. Miyagi Y, Chiu LL, Cimini M, Weisel RD, Radisic M, Li RK. Biodegradable collagen patch with covalently immobilized VEGF for myocardial repair. Biomaterials 2011; 32(5): 1280-90.
  139. 139. Chiu LL, Radisic M. Controlled release of thymosin β4 using collagen–chitosan composite hydrogels promotes epicardial cell migration and angiogenesis. Journal of Controlled Release 2011; 155(3): 376-85. doi: 10.1016/j.jconrel.2011.05.026.
  140. 140. Vantler M, Karikkineth BC, Naito H, Tiburcy M, Didié M, Nose M, Rosenkranz S, Zimmermann WH. PDGF-BB protects cardiomyocytes from apoptosis and improves contractile function of engineered heart tissue. Journal of Molecular and Cellular Cardiology 2010; 48(6): 1316-23.
  141. 141. Davis ME, Hsieh PC, Grodzinsky AJ, Lee RT. Custom design of the cardiac microenvironment with biomaterials. Circulation Research 2005; 97(1): 8-15.
  142. 142. Kaully T, Kaufman-Francis K, Lesman A, Levenberg S. Vascularization--the conduit to viable engineered tissues. Tissue Engineering Part B Reviews 2009; 15(2): 159-69.
  143. 143. Bar A, Haverich A, Hilfiker A. Cardiac tissue engineering: ‘‘Reconstructing the motor of life’’. Scandinavian Journal of surgery 2007; 96 (2): 154-8.
  144. 144. Narmoneva DA, Vukmirovic R, DavisME, Kamm RD, Lee RT. Endothelial cells promote cardiacmyocyte survival and spatial reorganization: implications for cardiac regeneration. Circulation 2004; 110(8): 962-8.
  145. 145. Dvir T, Kedem A, Ruvinov E, Levy O, Freeman I, Landa N, Holbova R, Feinberg MS, Dror S, Etzion Y, Leor J, Cohen S. Prevascularization of cardiac patch on the omentum improves its therapeutic outcome. Proceedings of the National Academy of Sciences from the United States of America 2009; 106 (35): 14990-5.
  146. 146. Bursac N, Papadaki M, White JA, Eisenberg SR, Vunjak-Novakovic G, Freed LE. Cultivation in rotating bioreactors promotes maintenance of cardiac myocyte electrophysiology and molecular properties. Tissue Engineering 2003; 9(6): 1243-53.
  147. 147. Radisic M, Yang L, Boublik J, Cohen RJ, Langer R, Freed LE, Vunjak-Novakovic G. Medium perfusion enables engineering of compact and contractile cardiac tissue. American Journal of Physiology Heart and Circulatory Physiology 2004; 286(2): H507-16.
  148. 148. Radisic M, Marsano A, Maidhof R, Wang Y, Vunjak-Novakovic G. Cardiac tissue engineering using perfusion bioreactor systems. Nature Protocols 2008; 3(4): 719-38.
  149. 149. Maidhof R, Marsano A, Lee EJ, Vunjak-Novakovic G. Perfusion Seeding of Channeled Elastomeric Scaffolds with Myocytes and Endothelial Cells for Cardiac Tissue Engineering. Biotechnology Progress 2010; 26(2): 565-72.
  150. 150. Brown MA, Iyer RK, Radisic M. Pulsatile perfusion bioreactor for cardiac tissue engineering. Biotechnology Progress 2008; 24(4): 907-20.
  151. 151. Cheng M, Moretti M, Engelmayr GC, Freed LE. Insulin-like Growth Factor-I and Slow, Bi-directional Perfusion Enhance the Formation of Tissue-Engineered Cardiac Grafts. Tissue Engineering Part A 2009; 15(3): 645-53.
  152. 152. Li D, Xia Y. Electrospinning of nanofibers: reinventing the wheel? Advanced Materials 2004; 16(14): 1151–1170. doi: 10.1002/adma.200400719.
  153. 153. Orlova Y, Magome N, Liu L, Chen Y, Agladze K. Electrospun nanofibers as a tool for architecture control in engineered cardiac tissue. Biomaterials 2011; 32(24): 5615-24.
  154. 154. Blan NR, Birla RK. Design and fabrication of heart muscle using scaffold-based tissue engineering. Journal of Biomedical Materials Research Part A 2008; 86(1): 195-208.
  155. 155. Madihally SV, Matthew HW. Porous chitosan scaffolds for tissue engineering. Biomaterials 1999; 20(12): 1133-42.
  156. 156. Cimetta E, Pizzato S, Bollini S, Serena E, De Coppi P, Elvassore N. Production of arrays of cardiac and skeletal muscle myofibers by micropatterning techniques on a soft substrate. Biomedical Microdevices 2009; 11(2): 389-400.
  157. 157. Chiang CK, Chowdhury MF, Iyer RK, Stanford WL, Radisic M. Engineering surfaces for site-specific vascular differentiation of mouse embryonic stem cells. Acta Biomaterialia 2010; 6(6): 1904-16.
  158. 158. Park H, Radisic M, Lim JO, Chang BH, Vunjak-Novakovic G. A novel composite scaffold for cardiac tissue engineering. In Vitro Cellular and Developmental Biology Animal 2005; 41(7): 188-96.
  159. 159. Eschenhagen T, Fink C, Remmers U, Scholz H, Wattchow J, Weil J, Zimmermann W, Dohmen HH, Schäfer H, Bishopric N, Wakatsuki T, Elson EL. Three-dimensional reconstitution of embryonic cardiomyocytes in a collagen matrix: a new heart muscle model system. The Journal of the Federation of American Societies for Experimental Biology 1997; 11(8): 683-94.
  160. 160. Callegari A, Bollini S, Iop L, Chiavegato A, Torregrossa G, Pozzobon M, Gerosa G, De Coppi P, Elvassore N, Sartore S. Neovascularization induced by porous collagen scaffold implanted on intact and cryoinjured rat hearts. Biomaterials 2007; 28(36): 5449-61.
  161. 161. Xiang Z, Liao R, Kelly MS, Spector M. Collagen-GAG scaffolds grafted onto myocardial infarcts in a rat model: a delivery vehicle for mesenchymal stem cells. Tissue Engineering 2006; 12(9): 2467-78.
  162. 162. Simpson DL, Dudley SC Jr. Modulation of human mesenchymal stem cell function in a three-dimensional matrix promotes attenuation of adverse remodelling after myocardial infarction. Journal of Tissue Engineering and Regenerative Medicine 2011 Nov 18. doi: 10.1002/term.511.
  163. 163. Chachques JC, Trainini JC, Lago N, Masoli OH, Barisani JL, Cortes-Morichetti M, Schussler O, Carpentier A. Myocardial assistance by grafting a new bioartificial upgraded myocardium (MAGNUM clinical trial): one year follow-up. Cell Transplantation 2007; 16(9): 927-34.
  164. 164. Chimenti I, Rizzitelli G, Gaetani R, Angelini F, Ionta V, Forte E, Frati G, Schussler O, Barbetta A, Messina E, Dentini M, Giacomello A. Human cardiosphere-seeded gelatin and collagen scaffolds as cardiogenic engineered bioconstructs. Biomaterials 2011; 32(35): 9271-81.
  165. 165. Holladay CA, Duffy AM, Chen X, Sefton MV, O'Brien TD, Pandit AS. Recovery of cardiac function mediated by MSC and interleukin-10 plasmid functionalised scaffold. Biomaterials 2012; 33(5): 1303-14.
  166. 166. Akhyari P, Kamiya H, Haverich A, Karck M, Lichtenberg A. Myocardial tissue engineering: the extracellular matrix. Journal of Cardio-thoracic Surgery 2008; 34: 229-241.
  167. 167. Sakai T, Li RK, Weisel RD, Mickle DA, Kim ET, Jia ZQ, Yau TM. The fate of a tissue-engineered cardiac graft in the right ventricular outflow tract of the rat. Journal of Thoracic and Cardiovascular Surgery 2001; 121(5): 932-42.
  168. 168. Zhang G, Nakamura Y, Wang X, Hu Q, Suggs LJ, Zhang J. Controlled release of stromal cell-derived factor-1 alpha in situ increases c-kit+ cell homing to the infarcted heart. Tissue Engineering 2007; 13(8): 2063-71.
  169. 169. Leor J, Aboulafia-Etzion S, Dar A, Shapiro L, Barbash IM, Battler A, Granot Y, Cohen S. Bioengineered cardiac grafts: A new approach to repair the infarcted myocardium? Circulation 2000; 102(19 Suppl 3): III56-61.
  170. 170. Dar A, Shachar M, Leor J, Cohen S. Optimization of cardiac cell seeding and distribution in 3D porous alginate scaffolds. Biotechnology and Bioengineering 2002; 80(3): 305-12.
  171. 171. Shachar M, Tsur-Gang O, Dvir T, Leor J, Cohen S. The effect of immobilized RGD peptide in alginate scaffolds on cardiac tissue engineering. Acta Biomaterialia 2011; 7(1): 152-62.
  172. 172. Sapir Y, Kryukov O, Cohen S. Integration of multiple cell-matrix interactions into alginate scaffolds for promoting cardiac tissue regeneration. Biomaterials 2011; 32(7): 1838-47.
  173. 173. Le Visage C, Gournay O, Benguirat N, Hamidi S, Chaussumier L, Mougenot N, Flanders JA, Isnard R, Michel JB, Hatem S, Letourneur D, Norol F. Mesenchymal stem cell delivery into rat infarcted myocardium using a porous polysaccharide-based scaffold: a quantitative comparison with endocardial injection. Tissue Engineering Part A 2012; 18(1-2): 35-44.
  174. 174. Cao Y, Wang B. Biodegradation of Silk Biomaterials. International Journal of Molecular Sciences 2009; 10(4): 1514-1524.
  175. 175. Yang MC, Wang SS, Chou NK, Chi NH, Huang YY, Chang YL, Shieh MJ, Chung TW. The cardiomyogenic differentiation of rat mesenchymal stem cells on silk fibroin–polysaccharide cardiac patches in vitro. Biomaterials 2009; 30(22): 3757-65.
  176. 176. Patra C, Talukdar S, Novoyatleva T, Velagala SR, Mühlfeld C, Kundu B, Kundu SC, Engel FB. Silk protein fibroin for cardiac tissue engineering. Biomaterials 2012; 33(9): 2673-80.
  177. 177. Robinson KA, Li J, Mathison M, Redkar A, Cui J, Chronos NA, Matheny RG, Badylak SF. Extracellular matrix scaffold for cardiac repair. Circulation 2005; 112(9 Suppl): I135-43.
  178. 178. Godier-Furnémont AF, Martens TP, Koeckert MS, Wan L, Parks J, Arai K, Zhang G, Hudson B, Homma S, Vunjak-Novakovic G. Composite scaffold provides a cell delivery platform for cardiovascular repair. Proceedings of the National Academy of Sciences of the United States of America 2011; 108(19): 7974-9.
  179. 179. Wei HJ, Chen CH, Lee WY, Chiu I, Hwang SM, Lin WW, Huang CC, Yeh YC, Chang Y, Sung HW. Bioengineered cardiac patch constructed from multilayered mesenchymal stem cells for myocardial repair. Biomaterials 2008; 29(26): 3547-56.
  180. 180. Ott HC, Matthiesen TS, Goh SK, Black LD, Kren SM, Netoff TI, Taylor DA. Perfusion-decellularized matrix: using nature’s platform to engineer a bioartificial heart. Nature Medicine 2008; 14(2): 213–221.
  181. 181. Giraud MN, Armbruster C, Carrel T, Tevaearai HT. Current state of the art in myocardial tissue engineering. Tissue Engineering 2007; 13(8): 1825-36.
  182. 182. Huang CC, Wei HJ, Yeh YC, Wang JJ, Lin WW, Lee TY, Hwang SM, Choi SW, Xia Y, Chang Y, Sung HW. Injectable PLGA porous beads cellularized by hAFSCs for cellular cardiomyoplasty. Biomaterials 2012; 33(16): 4069-77.
  183. 183. McDevitt TC, Angello JC, Whitney ML, Reinecke H, Hauschka SD, Murry CE, Stayton PS. In vitro generation of differentiated cardiac myofibers on micropatterned laminin surfaces. Journal of Biomedical Materials Research 2002; 60(3): 472-9.
  184. 184. Stout DA, Basu B, Webster TJ. Poly(lactic–co-glycolic acid): Carbon nanofiber composites for myocardial tissue engineering applications. Acta Biomaterialia 2011; 7(8): 3101-12. doi:10.1016/j.actbio.2011.04.028 4.
  185. 185. Caspi O, Lesman A, Basevitch Y, Gepstein A, Arbel G, Habib IH, Gepstein L, Levenberg S. Tissue engineering of vascularized cardiac muscle from human embryonic stem cells. Circulation Research 2007; 100(2): 263-72.
  186. 186. Ishii O, Shin M, Sueda T, Vacanti JP. In vitro tissue engineering of a cardiac graft using a degradable scaffold with an extracellular matrix–like topography. Journal of Thoracic and Cardiovascular Surgery 2005; 130(5): 1358-63.
  187. 187. Piao H, Kwon JS, Piao S, Sohn JH, Lee YS, Bae JW, Hwang KK, Kim DW, Jeon O, Kim BS, Park YB, Cho MC. Effects of cardiac patches engineered with bone marrow-derived mononuclear cells and PGCL scaffolds in a rat myocardial infarction model. Biomaterials 2007; 28(4): 641-9.
  188. 188. Gorna K, Gogolewski S. Biodegradable polyurethanes for implants. II. In vitro degradation and calcification of materials from poly(epsilon-caprolactone)-poly(ethylene oxide) diols and various chain extenders. Journal of Biomedical Materials Research 2002; 60(4): 592-606.
  189. 189. Zhang JY, Beckman EJ, Piesco NP, Agarwal S. A new peptide-based urethane polymer: synthesis, biodegradation, and potential to support cell growth in vitro. Biomaterials 2000; 21(12): 1247-1258.
  190. 190. Rockwood DN, Akins RE Jr, Parrag IC, Woodhouse KA, Rabolt JF. Culture on electrospun polyurethane scaffolds decreases atrial natriuretic peptide expression by cardiomyocytes in vitro. Biomaterials 2008; 29(36): 4783-91. doi:10.1016/j.biomaterials.2008.08.034.
  191. 191. Guan J, Fujimoto KL, Sacks MS, Wagner WR. Preparation and characterization of highly porous, biodegradable polyurethane scaffolds for soft tissue applications. Biomaterials 2005; 26(18): 3961-3971. doi:10.1016/j.biomaterials.2004.10.018.
  192. 192. Fujimoto KL, Tobita K, Merryman WD, Guan J, Momoi N, Stolz DB, Sacks MS, Keller BB, Wagner WR. An elastic, biodegradable cardiac patch induces contractile smooth muscle and improves cardiac remodeling and function in subacute myocardial infarction. Journal of the American College of Cardiology 2007; 49(23): 2292-300. doi: 10.1016/j.jacc.2007.02.050.
  193. 193. Siepe M, Giraud MN, Liljensten E, Nydegger U, Menasche P, Carrel T, Tevaearai HT. Construction of skeletal myoblast-based polyurethane scaffolds for myocardial repair. Artificial Organs 2007; 31(6): 425-33.
  194. 194. Chen QZ, Bismarck A, Hansen U, Junaid S, Tran MQ, Harding SE, Ali NN, Boccaccini AR. Characterisation of a soft elastomer poly(glycerol sebacate) designed to match the mechanical properties of myocardial tissue. Biomaterials 2008; 29(1): 47-57.
  195. 195. Jean A, Engelmayr GC Jr. Finite element analysis of an accordion-like honeycomb scaffold for cardiac tissue engineering. Journal of Biomechanics 2010; 43(15): 3035-43.
  196. 196. Ravichandran R, Venugopal JR, Sundarrajan S, Mukherjee S, Ramakrishna S. Poly(glycerol sebacate)/gelatin core/shell fibrous structure for regeneration of myocardial infarction. Tissue Engineering Part A 2011; 17(9-10): 1363-73. doi: 10.1089/ten.tea.2010.0441.
  197. 197. Ifkovits JL, Devlin JJ, Eng G, Martens TP, Vunjak-Novakovic G, Burdick JA. Biodegradable fibrous scaffolds with tunable properties formed from photo-cross-linkable poly(glycerol sebacate). ACS Applied Materials and Interfaces 2009; 1(9): 1878-86.
  198. 198. Madden LR, Mortisen DJ, Sussman EM, Dupras SK, Fugate JA, Cuy JL, Hauch KD, Laflamme MA, Murry CE, Ratner BD. Proangiogenic scaffolds as functional templates for cardiac tissue engineering. Proceedings of the National Academy of Sciences of the United States of America 2010; 107(34): 15211-6. doi: 10.1073/pnas.1006442107.
  199. 199. Arnal-Pastor M, Vallés-Lluch A, Keicher M, Pradas MM. Coating typologies and constrained swelling of hyaluronic acid gels within scaffold pores. Journal of Colloid and Interface Science 2011; 361(1): 361-9.
  200. 200. Sauer H, Rahimi G, Hescheler J, Wartenberg M. Effects of electrical fields on cardiomyocyte differentiation of embryonic stem cells. Journal of Cellular Biochemistry 1999; 75(4): 710–723.
  201. 201. Haneef K, Lila N, Benadda S, Legrand F, Carpentier A, Chachques JC. Development of bioartificial myocardium by electrostimulation of 3D collagen scaffolds seeded with stem cells. Heart International 2012; 7(2): e14.
  202. 202. Radisic M, Park H, Shing H, Consi T, Schoen FJ, Langer R, Freed LE, Vunjak-Novakovic G. Functional assembly of engineered myocardium by electrical stimulation of cardiac myocytes cultured on scaffolds. Proceedings of the National Academy of Sciences of the United States of America 2004; 101(52): 18129-34.
  203. 203. Serena E, Figallo E, Tandon N, Cannizzaro C, Gerecht S, Elvassore N, Vunjak-Novakovic G. Electrical stimulation of human embryonic stem cells: Cardiac differentiation and the generation of reactive oxygen species. Experimental Cell Research 2009; 315(20): 3611-9.
  204. 204. Tandon N, Marsanno A, Maidhof R, Wan L. Park H, Vunjak-Novakovic G. Optimization of electrical stimulation parameters for cardiac tissue engineering. Journal of Tissue Engineering and Regenerative Medicine 2011; 5: e115–e125.
  205. 205. Tandon N, Marsano A, Maidhof R, Numata K, Montouri-Sorrentino C, Cannizzaro C, Voldmand J, Vunjak-Novakovic G. Surface-patterned electrode bioreactor for electrical stimulation. Lab on a Chip 2010; 10: 692–700.
  206. 206. Dvir T, Timko BP, Brigham MD, Naik SR, Karajanagi SS, Levy O, Jin H, Parker KK, Langer R, Kohane DS. Nanowired three-dimensional cardiac patches. Nature Nanotechnology 2011; 6(11): 720-5. doi:10.1038/nnano.2011.160.
  207. 207. You J-O, Rafat M, Ye GJC Auguste,DT. Nanoengineering the heart: conductive scaffolds enhance connexin 43 expression. Nano Letters 2011; 11(9): 3643–3648.
  208. 208. Fink C, Ergün S, Kralisch D, Remmers U, Weil J, Eschenhagen T. Chronic stretch of engineered heart tissue induces hypertrophy and functional improvement. Federation of American Societies for Experimental Biology Journal 2000; 14(5): 669-79.
  209. 209. Zimmermann WH, Schneiderbanger K, Schubert P, Didié M, Münzel M, Heubach F,Kostin S, Neuhuber WL, Eschenhagen T. Tissue engineering of a differentiated cardiac muscle construct. Circulation Research 2002; 90: 223-230.
  210. 210. Guan J, Wang F, Li Z, Chen J, Guo X, Liao J, Moldovan NI. The stimulation of the cardiac differentiation of mesenchymal stem cells in tissue constructs that mimic myocardium structure and biomechanics. Biomaterials 2011; 32(24): 5568-80.
  211. 211. Zhang T, Wan LQ, Xiong Z, Marsano A, Maidhof R, Park M, Yan Y, Vunjak-Novakovic G. Channelled scaffolds for engineering myocardium with mechanical stimulation. Journal of Tissue Engineering and Regenerative Medicine 2011. doi: 10.1002/term.481.
  212. 212. Akhyari P, Fedak PW, Weisel RD, Lee TY, Verma S, Mickle DA, Li RK. Mechanical stretch regimen enhances the formation of bioengineered autologous cardiac muscle grafts. Circulation 2002; 106(12 Suppl 1): I137-42.
  213. 213. Chachques JC, Jegaden O, Mesana T, Glock Y, Grandjean PA, Carpentier AF, et al. Cardiac bioassist: results of the French multicenter cardiomyoplasty study. Asian Cardiovascular and Thoracic Annals 2009; 17: 573-80.
  214. 214. Kwon MH, Cevasco M, Schmitto JD, Chen FY. Ventricular restraint therapy for heart failure: A review, summary of state of the art, and future directions. Journal of Thoracic and Cardiovascular Surgery 2012; 144(4): 771-777.
  215. 215. Liao SY, Siu CW, Liu Y, Zhang Y, Chan WS, Wu EX, Wu Y, Nicholls JM, Li RA, Benser ME, Rosenberg SP, Park E, Lau CP, Tse HF. Attenuation of left ventricular adverse remodeling with epicardial patching after myocardial infarction. Journal of Cardiac Failure 2010; 16 (7): 590-8.
  216. 216. Enomoto Y, Gorman JH 3rd, Moainie SL, Jackson BM, Parish LM, Plappert T, Zeeshan A, St John-Sutton MG, Gorman RC. Early ventricular restraint after myocardial infarction: extent of the wrap determines the outcome of remodeling. Annals of Thoracic Surgery 2005; 79(3): 881-7.
  217. 217. Klodell CT, Aranda JM, McGiffin DC, Rayburn BK, Sun B, Abraham WT, Pae WE, Boehmer JP, Klein H, Huth C. Worldwide surgical experience with the Paracor HeartNet cardiac restraint device. The Journal of Thoracic and Cardiovascular Surgery 2008; 135(1): 188–195.
  218. 218. Costanzo MR, Ivanhoe RJ, Kao A, Anand IS, Bank A, Boehmer J, Demarco T, Hergert CM, Holcomb RG, Maybaum S, Sun B, Vassiliades TA Jr, Rayburn BK, Abraham WT. Prospective evaluation of elastic restraint to lessen the effects of heart failure (PEERLESS-HF) trial. Journal of Cardiac Failure 2012; 18(6): 446-58. doi: 10.1016/j.cardfail.2012.04.004.
  219. 219. Dixon JA, Goodman AM, Gaillard WF 2nd, Rivers WT, McKinney RA, Mukherjee R, Baker NL, Ikonomidis JS, Spinale FG. Hemodynamics and myocardial blood flow patterns after placement of a cardiac passive restraint device in a model of dilated cardiomyopathy. Journal of Thoracic and Cardiovascular Surgery 2011; 142: 1038-45.
  220. 220. Pilla JJ, Blom AS, Brockman DJ, Ferrari VA, Yuan Q, Acker MA. Passive ventricular constraint to improve left ventricular function and mechanics in an ovine model of heart failure secondary to acute myocardial infarction. Journal of Thoracic and Cardiovascular Surgery 2003; 126(5): 1467-76.
  221. 221. Mann DL, Kubo SH, Sabbah HN, Starling RC, Jessup M, Oh JK, Acker MA. Beneficial effects of the CorCap cardiac support device: five-year results from the Acorn Trial. Journal of Thoracic and Cardiovascular Surgery 2012; 143(5): 1036-42.
  222. 222. Olsson A, Bredin F, Franco-Cereceda A. Echocardiographic findings using tissue velocity imaging following passive containment surgery with the Acorn CorCap cardiac support device. European Journal of Cardio-Thoracic Surgery 2005; 28: 448-53.
  223. 223. Shafy A, Fink T, Zachar V, Lilaa N, Carpentier A, Chachques JC. Development of cardiac support bioprostheses for ventricular restoration and myocardial regeneration. European Journal of Cardio-Thoracic Surgery 2012; 0: 1–9. doi:10.1093/ejcts/ezs480.

Written By

M. Arnal-Pastor, J. C. Chachques, M. Monleón Pradas and A. Vallés- Lluch

Submitted: 15 May 2012 Published: 22 May 2013