Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 3031 word(s) 3031 2021-08-17 12:08:42 |
2 The format is correct Meta information modification 3031 2021-09-28 04:32:36 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Ul Haq, A. Conductive Polymers in Infarction Repair. Encyclopedia. Available online: https://encyclopedia.pub/entry/14481 (accessed on 29 March 2024).
Ul Haq A. Conductive Polymers in Infarction Repair. Encyclopedia. Available at: https://encyclopedia.pub/entry/14481. Accessed March 29, 2024.
Ul Haq, Arsalan. "Conductive Polymers in Infarction Repair" Encyclopedia, https://encyclopedia.pub/entry/14481 (accessed March 29, 2024).
Ul Haq, A. (2021, September 23). Conductive Polymers in Infarction Repair. In Encyclopedia. https://encyclopedia.pub/entry/14481
Ul Haq, Arsalan. "Conductive Polymers in Infarction Repair." Encyclopedia. Web. 23 September, 2021.
Conductive Polymers in Infarction Repair
Edit

The function of the heart pump may be impaired by events such as myocardial infarction, the consequence of coronary artery thrombosis due to blood clots or plaques. A whole heart transplant remains the gold standard so far and the current pharmacological approaches tend to stop further myocardium deterioration, but this is not a long-term solution. Electrically conductive, scaffold-based cardiac tissue engineering provides a promising solution to repair the injured myocardium. The non-conductive component of the scaffold provides a biocompatible microenvironment to the cultured cells while the conductive component improves intercellular coupling as well as electrical signal propagation through the scar tissue when implanted at the infarcted site. The in vivo electrical coupling of the cells leads to a better regeneration of the infarcted myocardium, reducing arrhythmias, QRS/QT intervals, and scar size and promoting cardiac cell maturation. 

conductive polymers striated muscle cell electrical coupling cardiac tissue engineering cardiac muscle repair electrical signals biomimetic material constructs

1. Introduction

Myocardial infarction (MI) is one of the clusters of numerous other cardiovascular diseases and occurs due to the blockage of the coronary artery delivering blood (ischemia) to the ventricle with consequent oxygen shortage to the contractile cells (cardiomyocytes) [1]. The final result is that, after MI, billions of cardiomyocytes (CMs) with limited proliferation capacity are lost and substituted by heterogeneous collagen-rich fibrotic scar tissue [2][3]. Among others, the fibrotic scar does not display contractile capabilities and does not appropriately conduct electric currents, thus generating myocardial arrhythmias and asynchronous beating, contributing to determine heart failure in the worst-case scenario [2]. The current pharmacological approaches are palliative [4] and finalized to prevent intra-coronary blood clotting (thrombolytics, antiplatelet agents, such as aspirin, etc.) and post-ischemic ventricular dilation (ACE-inhibitors, β-blockers, etc.), but do not induce regeneration of the injured cardiac tissue. So far, the heart transplant is the gold standard in post-MI end-stage heart failure, but the lack of organ donors and the possibility of immune rejection makes this approach elusive [5].

In recent years, the parallel progresses in cell biology, materials science, and advanced nano-manufacturing procedures have allowed us to envision the possibility to set up novel strategies (collectively dubbed “tissue engineering”) to combine cells and biomaterials to fabricate myocardium-like structures in vitro to be engrafted into the heart to repair the damaged parts. This approach is characterized by an extreme level of complexity and, as a matter of fact, after more than two decades of extensive efforts, many issues remain to be answered before tissue engineering products can be used at the bedside.

Natural or synthetic biomaterials, such as cardiac patches [6], injectable hydrogels [7], nanofiber composites [8], nanoparticles [9], and 3D hydrogel constructs [10], have been scrutinized to produce structures that mimic the mechanical properties of the extracellular matrix of the myocardium and potentially restoring the cardiac functions [11][12]. However, the issues related to arrhythmias and asynchronous beating of the injured myocardium are not resolved due to the non-conductive nature of most of the polymers used so far. The solution to this issue has been envisioned in conveying electric signals through scaffolds with embedded polymers displaying electroconductive characteristics comparable to the biological tissues (Figure 1) and this has already been demonstrated to enhance cell differentiation to mature CMs [13].

Electrically inert biomaterials can be blended with conductive polymers, such as polyaniline (PANI), polypyrrole (PPy), and Poly(3,4-ethylenedioxythiphene)/PEDOT, which belong to the family of intrinsically conductive organic materials [14]. The resulting electroconductive scaffold would harness the biocompatibility and the mechanical properties of the inert biomaterial and the electrical nature of the conductive component to drive the differentiation of the cultured stem/progenitor cells to cardiomyocyte-like cells with synchronous beating patterns, and enhance the expression of cardiac-specific genes.

GATA4 and Nkx2.5 are cardiac-specific transcription factors that play a fundamental role in myocardial differentiation and cardiac hypertrophy in the early stages of cardiogenesis in a developing embryo, while cardiac troponins are expressed in the cardiac muscles of the mammalian and avian species [15][16][17]. Connexin 43 is the principal gap junction protein of the heart which mediates action potential propagation between cells to synchronise cardiac contraction. In addition to this canonical role, it may act as a transcription regulator [18]. This cell-laden conductive scaffold improves the electrical signal propagation through the scar tissue, when implanted in vivo, and results in repair/regeneration of the injured myocardium with elevated ventricular wall thickness, improved blood pumping ability, reduced scar size, shorter QRS/QT intervals, and lower risk of arrhythmias, as shown in Figure 2 [19].
Ijms 22 08550 g001 550
Figure 1. Electrical conductivity of common materials and biological tissues. The typical electrical conductivity values of intrinsically conductive polymers (top) are compared to those of common materials (centre) and biological tissues (bottom). Values are expressed in S/cm. Adapted with permission from [13].
Ijms 22 08550 g002 550
Figure 2. Conductive scaffold-based cardiac tissue engineering. After MI, damaged and dead cardiomyocytes (CMs) compromise the electrical conduction of the myocardium. The in vitro cell-laden construct can be transplanted at the infarction site to restore the electrical functions and to repair the damaged myocardium.

2. Conductive Polymer-Based Scaffolds in Cardiac Tissue Engineering

2.1. Polyaniline

Polyaniline (PANI) is a widely used material in tissue engineering and biomedical applications (Table 1) due to its intrinsic ability to conduct electric currents and good biocompatibility [20][21][22][23]. The polymer’s conductivity can be tuned by chemical or electrochemical doping (p-doping (oxidation) or n-doping (reduction)) [24]. However, the powdered form of the PANI does not dissolve in its doped form in any common organic solvents and the electro-mechanical properties of the blends and composites depend on the uniform dispersion of the PANI particle in the polymer matrix. PANI particles can be prepared previously and then added to the matrix. Alternatively, aniline is polymerised to polyaniline in the polymer matrix (in situ chemical polymerisation method) [25]. Different PANI blends and composites have been created using various synthesis methods and combining PANI with diverse biomaterials to manufacture suitable conductive constructs for cardiac repair.
Table 1. Polyaniline based conductive constructs for cardiac tissue engineering.
Conductive
Substrate
Mechanical
Properties
Electrical
Properties
Cell Line
or Tissue
Biological Response
Poly-l-Lysine-PANI nanotubes
membranes [26]
    Rat CMs Better CMs proliferation
PLCL, PANI
electrospun
membranes [27]
E = 50 MPa,
εr = 207.85%,
UTS = 0.69 MPa
Four-probe
technique,
σ = 13.8 mS/cm
Human
fibroblasts, NIH-3T3, C2C12
Improved cell adhesion and
metabolic activity
PGLD, PANI
nanotubes
membranes [28]
    Cho cells,
neonatal rat CMs
Good biocompatibility
PU-AP/PCL
porous scaffold [29]
Ec = 4.1 MPa,
C.S = 1.3 MPa
Four-probe
technique,
σ = 10−5 S/cm
Neonatal rat CMs Enhanced Actn4, Cx43, and cTnT2
expressions.
PANI/PCL
patch [30]
  Two-probe
technique,
σ = 80 µS/cm
hMSCs Differentiation of hMSCs to
CM-like cells
PDLA/PANI
electrospun
membranes [31]
  σ = 44 mS/cm primary rat muscle cells Improved cell adhesion
and proliferation
Gelatin/PANI
electrospun
membranes [32]
E = 1384 MPa,
UTS = 10.49 MPa, εr = 9%
Four-probe
technique,
σ = 17 mS/cm
H9c2 Smooth muscle-like morphology
rich in microfilaments
Gelatin/PANI
hydrogels [33]
G’ = 5 Pa,
G” = 26 Pa
Pocket conductivity meter,
σ = 0.45 mS/cm
C2C12,
BM-MSCs
Improved cell-cell signalling
and proliferation
PU-AP/PCL
films [34]
E’ = 10 MPa
at 37 °C
Four-probe
technique,
σ = 10−5 S/cm
L929,
HUVECs
Improved cytocompatibility,
good antioxidant properties
PLGA, PANI
electrospun
meshes [35]
E = 91.7 MPa Four-point probe,
σ = 3.1 mS/cm
Neonatal rat CMs Enhanced Cx43 and cTnI
expressions
PGS/PANI
composites [36]
E = 6 MPa,
UTS = 9.2 MPa,
εr = 40%
Four-probe
technique,
σ = 18 mS/cm
C2C12 Good cell retention, growth,
and proliferation
PCL, amino capped AT films [37] E = 31.2 MPa,
UTS = 48.3 MPa,
εr = 646%
- C2C12 Spindle like morphology,
myotube formation
PCL, PANI
electrospun
membranes [38]
E = 55.2 MPa,
UTS = 10.5 MPa,
εr = 38.0%
Four-point probe,
σ = 63.6 mS/cm
C2C12 Myotube formation
PANI, E-PANI
films [39]
  Z > 10 MΩ/sqr for PANI
Z = 6 MΩ/sqr for
E-PANI
H9c2 Improved proliferation and cell
attachment on E-PANI
PLA/PANI electrospun membranes [40]   Four-probe
technique,
σ = 21 µS/m
H9c2,
rat CMs
Myotube formation from H9c2 cells, enhanced Cx43 and α-actinin
expression, improved
Ca2+ transients for CMs
PCL/SF/PANI
hydrogels [41]
εr = 107%   C2C12 Excellent cell alignment,
myotube formation
Chitosan-AT/PEG-DA hydrogels [42] G’ = 7 kPa Pocket conductivity meter,
σ = 2.42 mS/cm
C2C12,
H9c2
Improved cell viability
PGS-AT
elastomers [43]
E = 2.2 MPa,
UTS = 2.0 MPa,
εr = 141%
- H9c2,
rat CMs
Synchronous CM beating with
improved Ca2+ transients, H9c2 showed good orientation, enhanced Cx43 and α-actinin expression
PANI, Collagen,
HA electrospun
mats [44]
E = 0.02 MPa,
UTS= 4 MPa,
εr = 78%
Four-probe
technique,
σ = 2 mS/cm
Neonatal rat CMs, hiPSCs Synchronous beating of CMs derived from hiPSCs. Enhanced Cx43 and cTnI expression
AP, PLA
films [45]
  Four-point probe,
σ = 10−6 to 10−5 S/cm
H9c2 Pseudopodia like morphology,
improved Ca2+ transients
Chitosan, PANI
patch [46]
E = 6.73 MPa,
UTS = 5.26 MPa,
εr = 79%
Four-probe
technique,
σ = 0.162 S/cm
Rat MI
heart
Improved CV in the infarcted region with healing effects
PA, PANI
patch [47]
Elongation = 84% Digital Avometer,
σ = 2.79 S/m
Pork heart Cardiac ECM mimicking
HPLA/AT
films [48]
εr = 42.7%,
E = 758 MPa
  C2C12 Myotube formation
Dextran-AT/chitosan [49] G’ = 620 Pa at
t = 50 min
Four-probe
technique,
σ = 0.03 mS/cm
L929,
C2C12
high proliferation rate, good in vivo degradation, generation of new
myofibers
Abbreviations: PANI: Polyaniline, PLCL: poly(l-lactide-co-ε-caprolactone), E: elastic modulus, εr: strain at rupture, UTS: ultimate tensile strength, Ec: compressive modulus, C.S: compressive strength, G’: shear storage modulus, G”: shear loss modulus, E’: tensile storage modulus σ: electrical conductivity, CMs: cardiomyocytes, PU: polyurethane, AP: aniline pentamer, PCL: polycaprolactone, hMSCs: human mesenchymal stem cells, hiPSCs: human induced pluripotent stem cells, PDLA: poly (d-lactic acid), BM-MSCs: bone marrow-derived mesenchymal stem cells, PGS: polyglycerol sebacate, AT: aniline tetramer, E-PANI: polyaniline-emeraldine base, PLA: polylactic acid, SF: silk fibroin, PEG-DA: dibenzaldehyde-terminated poly(ethylene glycol), HA: hyaluronic acid, PA: polyamide, HPLA: hyperbranched polylactide, CV: conduction velocity

2.2. Polypyrrole

Polypyrrole (PPy) is another intrinsically conductive polymer that has been extensively studied in the field of biomedical and tissue engineering, as shown in Table 2 [50][51][52][53][54][55]. It has been reported that polypyrrole offers more favourable biological properties than polyaniline. However, under similar experimental conditions, cells cultured on these two polymeric substrates have demonstrated comparable biological responses [56]. PPy can be chemically modified and combined with other polymers to provide a biocompatible electrical microenvironment to the cells. This way, cell viability, proliferation rate, and intercellular coupling can be increased appreciably with a high potential to direct their differentiation toward cardiomyocyte-like phenotype. Poly(L-lactide-co-glycolic acid)/PPy conductive 3D electrospun scaffolds exhibited good biocompatibility when cultured with mouse cardiac progenitor cells and human-induced pluripotent stem cells (hiPSCs). Induced pluripotent stem cells proliferated along the conductive fibre lengths without severe apoptosis up to ten days post-culture [57]. Likewise, PPy-chitosan hydrogels offered no cytotoxicity to rat smooth muscle cells and promoted their proliferation, morphology, and metabolism. These conductive hydrogels significantly enhanced the intracellular Ca2+ transient propagation in cultured neonatal rat CMs and improved myocardial electrical signal conduction with elevated cardiac functions in the rat MI model [58]. Human-induced pluripotent stem cells derived CMs demonstrated a sarcomeric length of 1.86 µm three weeks post culture on nanopatterned silk fibroin/PPy conductive scaffolds [59] while the average sarcomeric length of human CMs in a relaxed state is about 2.2 µm [60]. This demonstrates the ability of the nanopatterned conductive scaffold to achieve near-physiological sarcomeric lengths owing to nanoscale channels which directed the unidirectional growth and proliferation of cells to achieve cardiac muscle-like striated morphology. Seven days post culture, hiPSCs differentiated to cardiomyocyte-like cells with Nkx2.5 and GATA4 expression elevated by 2 and 3.3-fold, respectively, when cultured on Poly(lactide-co-glycolic acid)/PPy conductive membranes [61].
Table 2. Polypyrrole based conductive constructs in cardiac tissue engineering.
Conductive
Substrate
Mechanical
Properties
Electrical
Properties
Cell Line or
Tissue
Biological Response
PCL, PPy films
[62]
Nanoindentation test, E = 0.93 GPa Keithley Parameter Analyzer,
ρ = 1.0 kΩ-cm
HL-1 murine CMs Enhanced Cx43 expression,
improved Ca2+ transients
Chitosan, PPy porous membranes [63] E = 486.7 kPa Three-probe
detector,
σ = 63 mS/m
NRVMs,
rat MI model
Improved cytoskeletal organisation with high beating amplitude, tissue morphogenesis at the MI site
SF, PPy
composites [59]
E = 200 MPa,
UTS = 7 MPa
Four-probe
technique,
σ = 1 S/cm
hPSC-CMs Enhanced expression of Cx43, Myh7, cTnT2, SCN5A genes,
elongated Z-band width and
sarcomeric length
Chitosan, PPy
hydrogels [58]
E = 3 kPa Four-point probe,
σ = 0.23 mS/cm
Neonatal rat CMs,
rat SMCs
Good proliferation with elevated calcium transients and
shorter QRS intervals
PLGA, PPy
membranes [57]
    Mice CPCs,
hiPSCs
Good biocompatibility and
proliferation rate
PLGA, PPy
membranes [61]
    hiPSCs Differentiation of hiPSCs to CMs, enhanced expression of actinin, Nkx2.5, GATA4, and Oct4
PCL, gelatin,
PPy electrospun
membranes [64]
E = 50.3 MPa,
εr = 3.7%
Four-probe
technique,
σ = 0.37 mS/cm
Rabbit primary CMs High proliferation rate,
enhanced expression of Cx43,
cTnT, and α-actinin
PPy, HPAE
hydrogels [65]
G’ = 35 kPa, Four-probe
technique,
σ = 0.65 mS/cm
L929,
BMSCs
Enhanced Cx43, α-SMA
expressions, excellent
cell viability and
biocompatibility
Abbreviations: PPy: polypyrrole, E: elastic modulus, εr: strain at rupture, G’: shear storage modulus, σ: electrical conductivity, NRVMs: neonatal rat ventricular myocytes, hPSCs-CMs: human pluripotent stem cells derived cardiomyocytes, SF: silk fibroin, CPCs: cardiac progenitor cells, hiPSCs: human induced pluripotent stem cells, PLGA: poly(lactic-co-glycolic acid), HPAE: hyperbranched poly(amino ester), BMSCs: bone marrow-derived mesenchymal stem cells, α-SMA: alpha-smooth muscle actin.

2.3. Poly(3,4-Ethylenedioxythiophene)/PEDOT

PEDOT-based electroconductive scaffolds represent a good electrical interface with biological cells/tissues owing to their intrinsic capability to conduct electrical currents, good stability, mechanical properties and biocompatibility, thus providing promising platforms for tissue engineering applications [66][67][68]. To enhance the poor water solubility, it is combined with Poly(styrenesulfonic acid)/PSS in a polymer blend. A conductive PEDOT:PSS/PEG hydrogel was prepared using two-step sequential polymerisation by in situ synthesis of PEDOT within the pre-crosslinked PEG hydrogel. To promote cell adhesion, RGD peptides were covalently attached to the conductive hydrogel surface. In vitro studies with H9C2 myocytes demonstrated good biocompatibility of this conductive hydrogel. It also supported cell adhesion and proliferation up to five days of culture [69]. Better results to promote cardiomyogenic differentiation have been obtained culturing cardiac cells on more complex conductive structures mimicking the fibrous and porous nature of cardiac ECM. A material mimicking elastomeric mechanical properties of the cardiac ECM was obtained by interpenetrating PEDOT into nitrile butadiene rubber (NBR) and Poly(ethylene glycol) dimethacrylate (PEGDM) crosslinked electrospun mats. PEDOT-embedded mats displayed high flexibility and conductivity, and induced enhanced expression of Cx43 and α-actinin in cardiomyocytes after five days of incubation, suggesting the presence of contractility and cell maturation [70]. Table 3 summarizes the applications of PEDOT-based conductive constructs in cardiac tissue engineering and their effect on cellular response.
Table 3. PEDOT based conductive constructs in cardiac tissue engineering.
Conductive
Substrate
Mechanical Properties Electrical
Properties
Cell Line Biological Response
PEG/PEDOT:PSS
hydrogels [69]
Ec = 21 kPa Four-probe
technique,
σ = 16.9 mS/cm
H9c2 Good cell viability and proliferation
GelMA/PEDOT:PSS hydrogels [71] E = 10.3 kPa ElS, Z = 261 kΩ
at 1 Hz
C2C12 Good cell viability and proliferation but high polymer concentration was
detrimental to cells
Collagen/alginate/
PEDOT:PSS
hydrogels [72]
G = 220 Pa,
τmax = 41 Pa
Four-probe
technique,
σ = 3.5 mS/cm
CMs,
hiPSCs-CMs
Good cell viability,
proliferation, and adhesion,
synchronous beating patterns
Alginate/PEDOT
hydrogels [73]
Ec = 175 kPa,
G’ = 100 kPa,
G” = 10 kPa
Electrochemical workstation,
σ = 61 mS/cm
BADSCs Differentiation of BADSCs to
CMs with enhanced expression
of cTnT, α-actinin, Cx43
NBR/PEGDM/PEDOT electrospun
membranes [70]
E = 3.8 MPa,
εr = 75.1%
Four-probe
technique,
σ = 5.8 S/cm
Cardiac
fibroblasts
Well organised sarcomeres, enhanced
expression of α-actinin, Cx43
PCL/PEDOT:PSS
microfibrous
scaffold [74]
E = 13 MPa   H9c2,
primary CMs
Enhanced expression of Cx43 and
α-actinin, synchronous
beating patterns of CMs
Abbreviations: E: elastic modulus, Ec: compressive modulus, εr: strain at rupture, G: shear modulus, G’: shear storage modulus, G”: shear loss modulus, τmax: maximum shear stress, σ: electrical conductivity, PEG: polyethylene glycol, EIS: electrochemical impedance spectroscopy, BADSCs: brown adipose-derived stem cells, NBR: nitrile butadiene rubber, PEGDM: poly(ethylene glycol) dimethacrylate, GelMA: gelatin methacrylate

3. The Underlying Mechanisms of the Positive Role of Conductive Substrates in Cardiac Tissue Engineering

Though the electroconductive scaffolds provide better platforms for the tissue engineering of nerve, cardiac, and skeletal muscle tissues, the mechanism by which these constructs influence cellular behaviour is yet to be known [75]. To uncover this, an equivalent circuit model was proposed for two groups of cardiomyocytes seeded on a conductive substrate. One group was assumed to be active (AG), which could fire spontaneous action potentials, while the other was assumed to be passive (PG), which could not fire spontaneous action potentials, but it was well electrically coupled. AG was considered the only source of ionic current, and the passive group was only affected by the electrical currents from the neighbouring AG. Nano gaps/clefts are formed between the cell membrane of the cultured cardiomyocytes and the substrate surface. Ionic solution filled these clefts during culturing, and it was modelled as seal resistance [76]. Seal resistance is usually generated by the solution between cell-substrate interfacial gaps and has a direct link with the adhesion strength of cells with the substrate [77]. When cells in the active group (AG) fire spontaneous action potentials, the electronic redistribution on the surface of the conductive substrate takes place just beneath the cells. The interfacial clefts between PG cells and substrate fill with more anions due to this electronic redistribution, which depolarises the plasma membrane of cells and could trigger an action potential. Larger seal resistance gives rise to higher excitation potential (Upeak) which indicates a strong electrical coupling between AG and PG. The passive group will fire an action potential once the threshold (Upeak) crosses the resting membrane potential of −90 mV [78][79]. In this way, PG could be stimulated and electrically synchronised to AG under the influence of action potentials generated by the active group, as shown in Figure 3.

Ijms 22 08550 g003 550
Figure 3. A theoretical model of ionic to electronic current transition for cells cultured on a conductive substrate. The ionic current interacting with electrons could direct their flow in other directions. This electronic current, opening the voltage-gated ion channels, would result in the influx of Na+, K+ and Ca2+ ions. This influx will depolarise the plasma membrane and the passive group (PG) will eventually fire action potentials.

4. Conductive Substrates for In Vivo Cardiac Repair

Electroconductive scaffolds represent a leap forward in the effort to manufacture supports to properly address cell fate towards a cardiomyocytic phenotype. The scaffold fabrication is a complex endeavour finalised to replicate in vitro the ECM microenvironment to foster proper cell-cell and cell-matrix interactions driving the cardiomyocyte maturation process. An optimal scaffold must be biocompatible (but not necessarily biodegradable, if the constituent materials are immunopermissive), with (i) adequate porosity (to deliver biologically active factors and remove cell waste) and (ii) mechano-structural design (to deliver physical signals, such as stiffness, micro/nano-topography, etc.), and (iii) with electroconductivity appropriate to mimic the ECM structure and function. Finally, these characteristics must allow the integration of the engineered tissue with the native injured myocardium to restore heart function.

5. Future Research

A biodegradable, intrinsically conductive polymer whose conductivity could be tuned to mimic the electrical features of the myocardium is of great necessity in cardiac tissue engineering. No material has ever been produced which would undergo perpetual contraction and relaxation cycles for an entire span of natural life, like myocardium. To mimic this versatility, future studies should take into consideration the complex biomechanics, anisotropy, micro/nano-architecture, and electrical properties of the myocardium.

References

  1. Nabel, E.G.; Braunwald, E. A Tale of Coronary Artery Disease and Myocardial Infarction. N. Engl. J. Med. 2012, 366, 54–63.
  2. Frangogiannis, N.G. Pathophysiology of myocardial infarction. Compr. Physiol. 2015, 5, 1841–1875.
  3. Bergmann, O.; Zdunek, S.; Felker, A.; Salehpour, M.; Alkass, K.; Bernard, S.; Sjostrom, S.L.; Szewczykowska, M.; Jackowska, T.; Dos Remedios, C.; et al. Dynamics of Cell Generation and Turnover in the Human Heart. Cell 2015, 161, 1566–1575.
  4. Sadahiro, T. Cardiac regeneration with pluripotent stem cell-derived cardiomyocytes and direct cardiac reprogramming. Regen. Ther. 2019, 11, 95–100.
  5. Zammaretti, P.; Jaconi, M. Cardiac tissue engineering: Regeneration of the wounded heart. Curr. Opin. Biotechnol. 2004, 15, 430–434.
  6. Zhang, D.; Shadrin, I.Y.; Lam, J.; Xian, H.Q.; Snodgrass, H.R.; Bursac, N. Tissue-engineered cardiac patch for advanced functional maturation of human ESC-derived cardiomyocytes. Biomaterials 2013, 34, 5813–5820.
  7. Hasan, A.; Khattab, A.; Islam, M.A.; Hweij, K.A.; Zeitouny, J.; Waters, R.; Sayegh, M.; Hossain, M.M.; Paul, A. Injectable Hydrogels for Cardiac Tissue Repair after Myocardial Infarction. Adv. Sci. 2015, 2, 1500122.
  8. Kim, P.H.; Cho, J.Y. Myocardial tissue engineering using electrospun nanofiber composites. BMB Rep. 2016, 49, 26–36.
  9. Hasan, A.; Morshed, M.; Memic, A.; Hassan, S.; Webster, T.J.; Marei, H.E.S. Nanoparticles in tissue engineering: Applications, challenges and prospects. Int. J. Nanomed. 2018, 13, 5637–5655.
  10. Ciocci, M.; Mochi, F.; Carotenuto, F.; Di Giovanni, E.; Prosposito, P.; Francini, R.; De Matteis, F.; Reshetov, I.; Casalboni, M.; Melino, S.; et al. Scaffold-in-scaffold potential to induce growth and differentiation of cardiac progenitor cells. Stem Cells Dev. 2017, 26, 1438–1447.
  11. Carotenuto, F.; Teodori, L.; Maccari, A.M.; Delbono, L.; Orlando, G.; Di Nardo, P. Turning regenerative technologies into treatment to repair myocardial injuries. J. Cell. Mol. Med. 2020, 24, 2704–2716.
  12. Carotenuto, F.; Manzari, V.; Di Nardo, P. Cardiac Regeneration: The Heart of the Issue. Curr. Transpl. Rep. 2021, 8, 67–75.
  13. Solazzo, M.; O’Brien, F.J.; Nicolosi, V.; Monaghan, M.G. The rationale and emergence of electroconductive biomaterial scaffolds in cardiac tissue engineering. APL Bioeng. 2019, 3, 041501.
  14. Quijada, C. Special issue: Conductive polymers: Materials and applications. Materials 2020, 13, 2344.
  15. Wei, B.; Jin, J.P. TNNT1, TNNT2, and TNNT3: Isoform genes, regulation, and structure-function relationships. Gene 2016, 582, 1–13.
  16. Watt, A.J.; Battle, M.A.; Li, J.; Duncan, S.A. GATA4 is essential for formation of the proepicardium and regulates cardiogenesis. Proc. Natl. Acad. Sci. USA 2004, 101, 12573–12578.
  17. Saadane, N.; Alpert, L.; Chalifour, L.E. Expression of immediate early genes, GATA-4, and Nkx-2.5 in adrenergic-induced cardiac hypertrophy and during regression in adult mice. Br. J. Pharmacol. 1999, 127, 1165–1176.
  18. Kotini, M.; Barriga, E.H.; Leslie, J.; Gentzel, M.; Rauschenberger, V.; Schambon, A.; Mayor, R. Gap junction protein Connexin-43 is a direct transcriptional regulator of N-cadherin in vivo. Nat. Commun. 2018, 9, 3846.
  19. He, S.; Wu, J.; Li, S.-H.; Wang, L.; Sun, Y.; Xie, J.; Ramnath, D.; Weisel, R.D.; Yau, T.M.; Sung, H.-W.; et al. The conductive function of biopolymer corrects myocardial scar conduction blockage and resynchronizes contraction to prevent heart failure. Biomaterials 2020, 258, 120285.
  20. Lalegül-ülker, Ö.; Murat, Y. Magnetic and electrically conductive silica-coated iron oxide/polyaniline nanocomposites for biomedical applications. Mater. Sci. Eng. C 2021, 119, 111600.
  21. Wibowo, A.; Vyas, C.; Cooper, G.; Qulub, F.; Suratman, R. 3D Printing of Polycaprolactone–Polyaniline Electroactive Sca ff olds for Bone Tissue Engineering. Materials 2020, 13, 512.
  22. Zhang, M.; Guo, B. Electroactive 3D Scaffolds Based on Silk Fibroin and Water-Borne Polyaniline for Skeletal Muscle Tissue Engineering. Macromol. Biosci. 2017, 17, 1–10.
  23. Karimi-soflou, R.; Nejati, S.; Karkhaneh, A. Electroactive and antioxidant injectable in-situ forming hydrogels with tunable properties by polyethylenimine and polyaniline for nerve tissue engineering. Colloids Surf. B Biointerfaces 2021, 199, 111565.
  24. Choi, C.H.; Park, S.H.; Woo, S.I. Binary and ternary doping of nitrogen, boron, and phosphorus into carbon for enhancing electrochemical oxygen reduction activity. ACS Nano 2012, 6, 7084–7091.
  25. Bhadra, J.; Alkareem, A.; Al-Thani, N. A review of advances in the preparation and application of polyaniline based thermoset blends and composites. J. Polym. Res. 2020, 27, 122.
  26. Fernandes, E.G.R.; Zucolotto, V.; De Queiroz, A.A.A. Electrospinning of hyperbranched poly-L-lysine/polyaniline nanofibers for application in cardiac tissue engineering. J. Macromol. Sci. Part A Pure Appl. Chem. 2010, 47, 1203–1207.
  27. Jeong, S.I.; Jun, I.D.; Choi, M.J.; Nho, Y.C.; Lee, Y.M.; Shin, H. Development of electroactive and elastic nanofibers that contain polyaniline and poly(L-lactide-co-ε-caprolactone) for the control of cell adhesion. Macromol. Biosci. 2008, 8, 627–637.
  28. Moura, R.M.; de Queiroz, A.A.A. Dendronized polyaniline nanotubes for cardiac tissue engineering. Artif. Organs 2011, 35, 471–477.
  29. Baheiraei, N.; Yeganeh, H.; Ai, J.; Gharibi, R.; Ebrahimi-Barough, S.; Azami, M.; Vahdat, S.; Baharvand, H. Preparation of a porous conductive scaffold from aniline pentamer-modified polyurethane/PCL blend for cardiac tissue engineering. J. Biomed. Mater. Res. Part A 2015, 103, 3179–3187.
  30. Borriello, A.; Guarino, V.; Schiavo, L.; Alvarez-Perez, M.A.; Ambrosio, L. Optimizing PANi doped electroactive substrates as patches for the regeneration of cardiac muscle. J. Mater. Sci. Mater. Med. 2011, 22, 1053–1062.
  31. McKeon, K.D.; Lewis, A.; Freeman, J.W. Electrospun poly(D,L-lactide) and polyaniline scaffold characterization. J. Appl. Polym. Sci. 2010, 115, 1566–1572.
  32. Li, M.; Guo, Y.; Wei, Y.; MacDiarmid, A.G.; Lelkes, P.I. Electrospinning polyaniline-contained gelatin nanofibers for tissue engineering applications. Biomaterials 2006, 27, 2705–2715.
  33. Li, L.; Ge, J.; Guo, B.; Ma, P.X. In situ forming biodegradable electroactive hydrogels. Polym. Chem. 2014, 5, 2880–2890.
  34. Baheiraei, N.; Yeganeh, H.; Ai, J.; Gharibi, R.; Azami, M.; Faghihi, F. Synthesis, characterization and antioxidant activity of a novel electroactive and biodegradable polyurethane for cardiac tissue engineering application. Mater. Sci. Eng. C 2014, 44, 24–37.
  35. Hsiao, C.W.; Bai, M.Y.; Chang, Y.; Chung, M.F.; Lee, T.Y.; Wu, C.T.; Maiti, B.; Liao, Z.X.; Li, R.K.; Sung, H.W. Electrical coupling of isolated cardiomyocyte clusters grown on aligned conductive nanofibrous meshes for their synchronized beating. Biomaterials 2013, 34, 1063–1072.
  36. Qazi, T.H.; Rai, R.; Dippold, D.; Roether, J.E.; Schubert, D.W.; Rosellini, E.; Barbani, N.; Boccaccini, A.R. Development and characterization of novel electrically conductive PANI-PGS composites for cardiac tissue engineering applications. Acta Biomater. 2014, 10, 2434–2445.
  37. Deng, Z.; Guo, Y.; Zhao, X.; Li, L.; Dong, R.; Guo, B.; Ma, P.X. Stretchable degradable and electroactive shape memory copolymers with tunable recovery temperature enhance myogenic differentiation. Acta Biomater. 2016, 46, 234–244.
  38. Chen, M.C.; Sun, Y.C.; Chen, Y.H. Electrically conductive nanofibers with highly oriented structures and their potential application in skeletal muscle tissue engineering. Acta Biomater. 2013, 9, 5562–5572.
  39. Bidez, P.R.; Li, S.; Macdiarmid, A.G.; Venancio, E.C.; Wei, Y.; Lelkes, P.I. Polyaniline, an electroactive polymer, supports adhesion and proliferation of cardiac myoblasts. J. Biomater. Sci. Polym. Ed. 2006, 17, 199–212.
  40. Wang, L.; Wu, Y.; Hu, T.; Guo, B.; Ma, P.X. Electrospun conductive nanofibrous scaffolds for engineering cardiac tissue and 3D bioactuators. Acta Biomater. 2017, 59, 68–81.
  41. Wang, L.; Wu, Y.; Guo, B.; Ma, P.X. Nanofiber Yarn/Hydrogel Core-Shell Scaffolds Mimicking Native Skeletal Muscle Tissue for Guiding 3D Myoblast Alignment, Elongation, and Differentiation. ACS Nano 2015, 9, 9167–9179.
  42. Dong, R.; Zhao, X.; Guo, B.; Ma, P.X. Self-Healing Conductive Injectable Hydrogels with Antibacterial Activity as Cell Delivery Carrier for Cardiac Cell Therapy. ACS Appl. Mater. Interfaces 2016, 8, 17138–17150.
  43. Hu, T.; Wu, Y.; Zhao, X.; Wang, L.; Bi, L.; Ma, P.X.; Guo, B. Micropatterned, electroactive, and biodegradable poly(glycerol sebacate)-aniline trimer elastomer for cardiac tissue engineering. Chem. Eng. J. 2019, 366, 208–222.
  44. Roshanbinfar, K.; Vogt, L.; Ruther, F.; Roether, J.A.; Boccaccini, A.R.; Engel, F.B. Nanofibrous Composite with Tailorable Electrical and Mechanical Properties for Cardiac Tissue Engineering. Adv. Funct. Mater. 2020, 30, 8612.
  45. Wang, Y.; Zhang, W.; Huang, L.; Ito, Y.; Wang, Z.; Shi, X.; Wei, Y.; Jing, X.; Zhang, P. Intracellular calcium ions and morphological changes of cardiac myoblasts response to an intelligent biodegradable conducting copolymer. Mater. Sci. Eng. C 2018, 90, 168–179.
  46. Mawad, D.; Mansfield, C.; Lauto, A.; Perbellini, F.; Nelson, G.W.; Tonkin, J.; Bello, S.O.; Carrad, D.J.; Micolich, A.P.; Mahat, M.M.; et al. A Conducting polymer with enhanced electronic stability applied in cardiac models. Sci. Adv. 2016, 2, e1601007.
  47. Yin, Y.; Mo, J.; Feng, J. Conductive fabric patch with controllable porous structure and elastic properties for tissue engineering applications. J. Mater. Sci. 2020, 55, 17120–17133.
  48. Xie, M.; Wang, L.; Guo, B.; Wang, Z.; Chen, Y.E.; Ma, P.X. Ductile electroactive biodegradable hyperbranched polylactide copolymers enhancing myoblast differentiation. Biomaterials 2015, 71, 158–167.
  49. Guo, B.; Qu, J.; Zhao, X.; Zhang, M. Degradable conductive self-healing hydrogels based on dextran-graft-tetraaniline and N-carboxyethyl chitosan as injectable carriers for myoblast cell therapy and muscle regeneration. Acta Biomater. 2019, 84, 180–193.
  50. Aznar-Cervantes, S.; Roca, M.I.; Martinez, J.G.; Meseguer-Olmo, L.; Cenis, J.L.; Moraleda, J.M.; Otero, T.F. Fabrication of conductive electrospun silk fibroin scaffolds by coating with polypyrrole for biomedical applications. Bioelectrochemistry 2012, 85, 36–43.
  51. Zanjanizadeh Ezazi, N.; Shahbazi, M.A.; Shatalin, Y.V.; Nadal, E.; Mäkilä, E.; Salonen, J.; Kemell, M.; Correia, A.; Hirvonen, J.; Santos, H.A. Conductive vancomycin-loaded mesoporous silica polypyrrole-based scaffolds for bone regeneration. Int. J. Pharm. 2018, 536, 241–250.
  52. Broda, C.R.; Lee, J.Y.; Sirivisoot, S.; Schmidt, C.E.; Harrison, B.S. A chemically polymerized electrically conducting composite of polypyrrole nanoparticles and polyurethane for tissue engineering. J. Biomed. Mater. Res. Part A 2011, 98, 509–516.
  53. Pan, X.; Sun, B.; Mo, X. Electrospun polypyrrole-coated polycaprolactone nanoyarn nerve guidance conduits for nerve tissue engineering. Front. Mater. Sci. 2018, 12, 438–446.
  54. Zhou, J.F.; Wang, Y.G.; Cheng, L.; Wu, Z.; Sun, X.D.; Peng, J. Preparation of polypyrrole-embedded electrospun poly(lactic acid) nanofibrous scaffolds for nerve tissue engineering. Neural Regen. Res. 2016, 11, 1644–1652.
  55. Björninen, M.; Gilmore, K.; Pelto, J.; Seppänen-Kaijansinkko, R.; Kellomäki, M.; Miettinen, S.; Wallace, G.; Grijpma, D.; Haimi, S. Electrically Stimulated Adipose Stem Cells on Polypyrrole-Coated Scaffolds for Smooth Muscle Tissue Engineering. Ann. Biomed. Eng. 2017, 45, 1015–1026.
  56. Humpolíček, P.; Kašpárková, V.; Pacherník, J.; Stejskal, J.; Bober, P.; Capáková, Z.; Radaszkiewicz, K.A.; Junkar, I.; Lehocký, M. The biocompatibility of polyaniline and polypyrrole: A comparative study of their cytotoxicity, embryotoxicity and impurity profile. Mater. Sci. Eng. C 2018, 91, 303–310.
  57. Gelmi, A.; Zhang, J.; Cieslar-Pobuda, A.; Ljunngren, M.K.; Los, M.J.; Rafat, M.; Jager, E.W.H. Electroactive 3D materials for cardiac tissue engineering. In Electroactive Polymer Actuators and Devices (EAPAD) 2015; SPIE-International Society for Optics and Photonics: Bellingham, WA, USA, 2015; Volume 9430.
  58. Mihic, A.; Cui, Z.; Wu, J.; Vlacic, G.; Miyagi, Y.; Li, S.H.; Lu, S.; Sung, H.W.; Weisel, R.D.; Li, R.K. A conductive polymer hydrogel supports cell electrical signaling and improves cardiac function after implantation into myocardial infarct. Circulation 2015, 132, 772–784.
  59. Tsui, J.H.; Ostrovsky-Snider, N.A.; Yama, D.M.P.; Donohue, J.D.; Choi, J.S.; Chavanachat, R.; Larson, J.D.; Murphy, A.R.; Kim, D.H. Conductive silk-polypyrrole composite scaffolds with bioinspired nanotopographic cues for cardiac tissue engineering. J. Mater. Chem. B 2018, 6, 7185–7196.
  60. Bird, S.D.; Doevendans, P.A.; Van Rooijen, M.A.; Brutel De La Riviere, A.; Hassink, R.J.; Passier, R.; Mummery, C.L. The human adult cardiomyocyte phenotype. Cardiovasc. Res. 2003, 58, 423–434.
  61. Gelmi, A.; Cieslar-Pobuda, A.; de Muinck, E.; Los, M.; Rafat, M.; Jager, E.W.H. Direct Mechanical Stimulation of Stem Cells: A Beating Electromechanically Active Scaffold for Cardiac Tissue Engineering. Adv. Healthc. Mater. 2016, 5, 1471–1480.
  62. Spearman, B.S.; Hodge, A.J.; Porter, J.L.; Hardy, J.G.; Davis, Z.D.; Xu, T.; Zhang, X.; Schmidt, C.E.; Hamilton, M.C.; Lipke, E.A. Conductive interpenetrating networks of polypyrrole and polycaprolactone encourage electrophysiological development of cardiac cells. Acta Biomater. 2015, 28, 109–120.
  63. Song, X.; Mei, J.; Ye, G.; Wang, L.; Ananth, A.; Yu, L.; Qiu, X. In situ pPy-modification of chitosan porous membrane from mussel shell as a cardiac patch to repair myocardial infarction. Appl. Mater. Today 2019, 15, 87–99.
  64. Kai, D.; Prabhakaran, M.P.; Jin, G.; Ramakrishna, S. Polypyrrole-contained electrospun conductive nanofibrous membranes for cardiac tissue engineering. J. Biomed. Mater. Res. Part A 2011, 99, 376–385.
  65. Liang, S.; Zhang, Y.; Wang, H.; Xu, Z.; Chen, J.; Bao, R.; Tan, B.; Cui, Y.; Fan, G.; Wang, W.; et al. Paintable and Rapidly Bondable Conductive Hydrogels as Therapeutic Cardiac Patches. Adv. Mater. 2018, 30, e1704235.
  66. Lu, B.; Yuk, H.; Lin, S.; Jian, N.; Qu, K.; Xu, J.; Zhao, X. Pure PEDOT:PSS hydrogels. Nat. Commun. 2019, 10, 1043.
  67. Guex, A.G.; Puetzer, J.L.; Armgarth, A.; Littmann, E.; Stavrinidou, E.; Giannelis, E.P.; Malliaras, G.G.; Stevens, M.M. Highly porous scaffolds of PEDOT:PSS for bone tissue engineering. Acta Biomater. 2017, 62, 91–101.
  68. Heo, D.N.; Lee, S.J.; Timsina, R.; Qiu, X.; Castro, N.J.; Zhang, L.G. Development of 3D printable conductive hydrogel with crystallized PEDOT:PSS for neural tissue engineering. Mater. Sci. Eng. C 2019, 99, 582–590.
  69. Kim, Y.S.; Cho, K.; Lee, H.J.; Chang, S.; Lee, H.; Kim, J.H.; Koh, W.G. Highly conductive and hydrated PEG-based hydrogels for the potential application of a tissue engineering scaffold. React. Funct. Polym. 2016, 109, 15–22.
  70. Fallahi, A.; Mandla, S.; Kerr-Phillip, T.; Seo, J.; Rodrigues, R.O.; Jodat, Y.A.; Samanipour, R.; Hussain, M.A.; Lee, C.K.; Bae, H.; et al. Flexible and Stretchable PEDOT-Embedded Hybrid Substrates for Bioengineering and Sensory Applications. ChemNanoMat 2019, 5, 729–737.
  71. Spencer, A.R.; Primbetova, A.; Koppes, A.N.; Koppes, R.A.; Fenniri, H.; Annabi, N. Electroconductive Gelatin Methacryloyl-PEDOT:PSS Composite Hydrogels: Design, Synthesis, and Properties. ACS Biomater. Sci. Eng. 2018, 4, 1558–1567.
  72. Roshanbinfar, K.; Vogt, L.; Greber, B.; Diecke, S.; Boccaccini, A.R.; Scheibel, T.; Engel, F.B. Electroconductive Biohybrid Hydrogel for Enhanced Maturation and Beating Properties of Engineered Cardiac Tissues. Adv. Funct. Mater. 2018, 28, 1803951.
  73. Yang, B.; Yao, F.; Ye, L.; Hao, T.; Zhang, Y.; Zhang, L.; Dong, D.; Fang, W.; Wang, Y.; Zhang, X.; et al. A conductive PEDOT/alginate porous scaffold as a platform to modulate the biological behaviors of brown adipose-derived stem cells. Biomater. Sci. 2020, 8, 3173–3185.
  74. Lei, Q.; He, J.; Li, D. Electrohydrodynamic 3D printing of layer-specifically oriented, multiscale conductive scaffolds for cardiac tissue engineering. Nanoscale 2019, 11, 15195–15205.
  75. Sikorski, P. Electroconductive scaffolds for tissue engineering applications. Biomater. Sci. 2020, 8, 5583–5588.
  76. Wu, Y.; Guo, L. Enhancement of intercellular electrical synchronization by conductive materials in cardiac tissue engineering. IEEE Trans. Biomed. Eng. 2018, 65, 264–272.
  77. Fendyur, A.; Mazurski, N.; Shappir, J.; Spira, M.E. Formation of essential ultrastructural interface between cultured hippocampal cells and gold mushroom-shaped MEA- towards “IN-CELL” recordings from vertebrate neurons. Front. Neuroeng. 2011, 4, 14.
  78. Nerbonne, J.M.; Kass, R.S. Molecular physiology of cardiac repolarization. Physiol. Rev. 2005, 85, 1205–1253.
  79. Santana, L.F.; Cheng, E.P.; Lederer, W.J. How does the shape of the cardiac action potential control calcium signaling and contraction in the heart? J. Mol. Cell. Cardiol. 2010, 49, 901–903.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 499
Revisions: 2 times (View History)
Update Date: 29 Mar 2022
1000/1000