Open access peer-reviewed chapter

Enhancement of Hydrogels’ Properties for Biomedical Applications: Latest Achievements

Written By

Ángel Serrano-Aroca

Submitted: 21 April 2017 Reviewed: 13 October 2017 Published: 01 August 2018

DOI: 10.5772/intechopen.71671

From the Edited Volume

Hydrogels

Edited by Sajjad Haider and Adnan Haider

Chapter metrics overview

1,774 Chapter Downloads

View Full Metrics

Abstract

Currently, there are many hydrogels used in many important biomedical fields such as therapeutic delivery, contact lenses, corneal prosthesis, bone cements, wound dressing, 3D tissue scaffolds for tissue engineering, etc., due to their excellent biocompatibility and water sorption properties. Many of these hydrophilic polymers have been already approved by the US Food and Drug Administration (FDA) for various applications. However, many of their potential uses required for many biomedical applications often are hindered by their low mechanical strength, antimicrobial and/or antifouling activity, biological interactions, water sorption and diffusion, porosity, electrical and/or thermal properties, among others. Thus, new advanced hydrogels have been developed as multicomponent systems in the form of composite or nanocomposite materials, which are expected to exhibit superior properties to increase the potential uses of these materials in the biomedical industry. Even though the great advances achieved so far, much research has to be conducted still in order to find new strategies to fabricate novel hydrogels able to overcome many of these problems.

Keywords

  • hydrogels
  • polymers
  • composites
  • nanocomposites
  • biomedical
  • properties

1. Introduction

Hydrogels are used in many fields of the biomedical industry such as therapeutic delivery [1], intraocular lenses, contact lenses and corneal prosthesis in ophthalmology [2], bone cements for orthopedics [3], wound dressing [4], 3D tissue scaffolds in regenerative medicine [5], etc., due to their excellent properties such as biocompatibility, water sorption and suitable mechanical performance, among others [6]. Many of these hydrogels have been approved by the US Food and Drug Administration (FDA) for diverse applications and are currently produced massively. However, many of their potential uses required for many biomedical applications are sometimes hindered by their low mechanical strength, biological interactions, electrical and/or thermal properties, water sorption and diffusion, antimicrobial and/or antifouling activity, porosity etc. Thus, new advanced hydrogels have been developed and are currently under intensive research to solve all these problems by means of multicomponent polymeric systems or by combination with other materials and/or nanomaterials to form composites or nanocomposites with enhanced required properties.

Advertisement

2. Mechanical properties

The enhancement of mechanical properties is one of the most desirable achievements in the field of hydrogel engineering and many researchers are currently working in this complex scientific area. Most hydrogels possesses very low mechanical properties, especially in the swollen state. Thus, hydrogels can be reinforced through many established kinds of methods and techniques: block copolymers, in which hydrophobic and hydrophilic domains alternate [7], increasing crosslinking density [8], by means of binary systems composed of two or more mixed polymers as interpenetrating polymer networks [9], by plasma grafting of a hydrogel onto a hydrophobic substrate [10, 11, 12], self-reinforced composite materials composed of fibers embedded in a matrix of the same polymer [13] and with the sol–gel reaction to produce nanosilica reinforcement [14]. However, more recent studies have shown new procedures to improve the mechanical properties of hydrogels with the incorporation of nanomaterials such as graphene (2010 Nobel Prize in Physics) and their derivatives: carbon nanotubes (CNT) [15], graphene oxide (GO) [16, 17], reduced graphene oxide (rGO) [18], etc.

2.1. IPNs

Interpenetrating polymer networks (IPN) have gained greater attention during last decades, mainly due to their biomedical applications as reinforced polymer networks. In 2003, Gong et al. [19] reported their discovery of a new hydrogel architecture that produced extraordinarily materials with enhanced mechanical properties, which they termed a double-network (DN) hydrogel. The DN hydrogel was originally believed to be an interpenetrating polymer network (IPN) of a soft neutral polymer network within a more highly crosslinked network prepared by a two-step sequential free-radical polymerization. The first step consisted of the synthesis of a highly crosslinked network, and the second step involved swelling this first network with a water soluble monomer that was then polymerized within it. The second polymerization step was conducted with or without adding a crosslinking agent. Thus, the use of an IPN, which consists of two separate but interwoven polymer networks, is a chemical procedure that is often used in polymer science to control, enhance, and/or combine functional properties. These advanced multicomponent polymeric systems are composed of crosslinked polymer networks without any covalent bonds between them, where at least one of them is synthesized and/or crosslinked within the immediate presence of the other. It is important to differentiate between the six basic multicomponent polymeric structures: mechanical blends, graft copolymers, block copolymers, AB-crosslinked copolymer, semi-IPNs and full-IPNs (see Figure 1) [20].

Figure 1.

Schematic representation of (a) mechanical blends, (b) graft copolymers, (c) block copolymers, (d) AB-crosslinked copolymer, (e) semi-IPNs and (f) full-IPNs. Modified from [20].

If a crosslinker is present in the polymeric system, fully-IPN [21] result, while in the absence of crosslinking, a network having linear polymers embedded within the first crosslinked network is formed (semi- or pseudo-IPN) [22, 23]. IPNs are prepared usually in the form of simultaneous interpenetrating polymer networks (SINs), in which the precursors of both networks are mixed and the two networks are synthesized at the same time, or in the form of sequential IPNs, by swelling of a single-polymer network into a solution containing the mixture of monomer, initiator and activator, usually with a crosslinker. Thus, poly(2-hydroxyethyl methacrylate) (PHEMA) networks were greatly reinforced by poly(ethylene glycol) in SINs and semi-SINs composed of both polymers [24]. Full-IPNs and semi-IPNs of weak gelatin hydrogels were also prepared by the sequential mode of synthesis with polyacrylic acid (PAA) to be evaluated for tissue response in rats [25]. The mechanical properties of pseudo-SIPNs and pseudo-IPNs hydrogels, where the prefix pseudo denotes connectivity of the two network, showed that non-linear tensile properties of pseudo-SIPNs are rate-dependent, but for pseudo-IPNs they are not, which is a consequence of the viscoelastic behavior of a pseudo-SIPN versus elastic performance of the pseudo-IPN [26]. In that study, the mechanical properties of triple-network (TN) hydrogels synthesized from pseudo SIPNs and pseudo-IPNs showed that the presence of a loosely crosslinked third network changes the mechanical behavior of pseudo-SIPNs and pseudo-IPNs by homogenizing the stress within the sample for finite deformations.

IPN hydrogels are also developed with the aim of enhancing the mechanical strength and swelling/deswelling response [27]. For example, interpenetrating polymer network (IPN) hydrogels of chitosan/poly(acrylic acid) (PAA) synthesized by the UV irradiation method showed that even in the swollen state, the present chitosan/PAA IPNs possessed good mechanical properties [28].

“Smart” hydrogels are able to significantly change their volume/shape in response to small alterations of certain parameters of the environment. These responsive hydrogels have numerous applications, being the most of them focused on biological and therapeutic demands [29, 30], and sensing applications [31].

Encapsulation of cells in interpenetrating network (IPN) hydrogels of two biocompatible materials-agarose and poly(ethylene glycol) (PEG) diacrylate with superior mechanical integrity has been developed [32].

Although IPNs based on hydrogels have been extensively reported, the combination of liquid crystalline (LC) property based hydrogels has been rarely explored. In this case, the anisotropic and molecular order of liquid crystals can be combined with the responsive isotropic properties of hydrogels. Thus, advanced stimuli-responsive materials based on interpenetrating liquid crystal-hydrogel polymer networks have been recently fabricated consisting of a cholesteric liquid crystalline network that reflects color and an interwoven poly(acrylic acid) network that provides a humidity and pH response [33].

2.2. Composite hydrogels

Diverse types of chemical modifications of hydrogels do not have a significant change of the overall mechanical strength because of the main structural skeleton of these polymers or copolymers are still weak. However, by fiber reinforcement, the addition of fabrics imparts high strength to the polymer networks, which form the main skeleton of the composites. Thus, fiber-reinforced hydrogels usually consists of a polymer matrix imbedded with high strength fibers, such as glass, aramid and carbon [34]. In such kind of materials, the mechanical properties are presumed to be improved and the biocompatible characteristics of the polymer should remain the same. Thus, hydrogels such as PHEMA, which is one of the most popular biomaterials, has been manufactured by adding various kinds of weaved and knitted fabrics and fibers, in order to improve overall qualities in advanced wound dressing usage [35]. However, in the recent decades, natural fibers as an alternative reinforcement in polymer composites have attracted the attention of many research groups due to their advantages over conventional glass and carbon fibers [36]. These natural fibers include flax, hemp, jute, sisal, kenaf, coir, kapok, banana, henequen and many others, which offer various advantages over man-made glass and carbon fibers such as low cost, low density, comparable specific tensile properties, non-abrasive to the equipment, non-irritation to the skin, reduced energy consumption, less health risk, renewability, recyclability and bio-degradability [37]. Thus, ultra-long chitin natural fibers were incorporated into hydrophobic Poly(methyl methacrylate) (PMMA) to prepare PMMA/Chitin composite hydrogels with improved properties [38]. This achievement is a significantly environmental move toward the sustainable utilization of marine-river crab shell wastes for biomedical applications in good agreement with green chemistry principles.

2.3. Nanocomposite hydrogels

Another alternative and very promising way of reinforcing hydrogels consist of the incorporation of nanomaterials such as silica, graphene and its derivatives, nanofibres or many other nanoparticles. Silica is a biocompatible material which has been reported to possess bioactive properties [39]. Silica can improve the mechanical properties of hydrogels through nanosilica filling or the well-known sol–gel process, which offers a new approach to the synthesis of nanocomposite materials with domain sizes approaching the molecular level [40]. Thus, for example, a biphasic matrix of a hybrid (inorganic–organic) nanocomposite materials of poly(2-hydroxyethyl acrylate) (PHEA) with a silica network obtained by an acid catalyzed sol–gel process of tetraethoxysilane (TEOS) showed a very significant improvement of the mechanical properties of the pure hydrogel [41].

Another reinforcement strategy that can be used consists of the combination of interpenetrated polymer networks and nanosilica filling. Thus, for example, poly(acrylic acid) and alginate IPN material with the incorporation of nanosilica greatly increased the compressive strength of the pure components [42].

Graphene (GN) is a two-dimensional monolayer of sp2-bonded carbon atoms [43], which has attracted increasing attention in the last decade owing to its excellent electrical and thermal conductivities [44, 45] and great mechanical strength [46]. Furthermore, graphene promotes adherence of human osteoblasts and mesenchymal stromal cells [47], which render this nanomaterial and its derivatives very promising in the biomedical research field. It has been recently reported that its oxidized form, graphene oxide (GO), can greatly enhance the compression performance of alginate hydrogels even in a minuscule concentration [48]. The improvement of mechanical strength of poly(acrylamide) (PAM), which generally exhibit pronounced weakness and brittleness, by incorporating GO to the polymer matrix has also been reported [49]. GO is also a 2D nanomaterial obtained from natural graphite that can be easily exfoliated into monolayer sheets. GO has many hydrophilic oxygenated functional groups, including hydroxyl (−OH), epoxy (−C-O-C-), carbonyl (−C=O) and carboxyl (−COOH), which groups enable its dispersion in water solution [50] and render possible its use in many synthetic procedures. The diversity of unique properties of graphene oxide, including great tensile modulus (1.0 TPa), ultimate strength (130 GPa), electrical and thermal properties [51], render GO an ideal carbon nanomaterial for variety of applications toward the development of new advanced materials. Thus, the addition of GO nanosheets increased the Young’s modulus and maximum stress of poly (acrylic acid)/gelatin composite hydrogels significantly as compared with control (0.0 wt. % GO). The highest Young’s modulus was observed for hydrogel with GO (0.2 wt. %)/PAA (20 wt. %), whereas the highest maximum stress was detected for GO (0.2 wt. %)/PAA (40 wt. %) specimen. These results suggests that GO nanosheets could be used to improve mechanical properties of hydrogel materials, which are very promising for tissue engineering applications in regenerative medicine [17]. In the biomedical area, a very innovative strategy for three-dimensional self-assembly of graphene oxide sheets and DNA to form multifunctional hydrogels with high mechanical strength, environmental stability, and dye-loading capacity, has also been recently reported [52]. Furthermore, the promising properties of GO with the availability of oxygen-containing functional groups has led the synthesis of 3D crosslinked GO networks able to improve the mechanical properties of alginate hydrogels even more than single GO nanosheets (Figure 2) [53].

Figure 2.

Confocal microscopy of (a) GO/alginate (GO/Alg) (b & d) crosslinked GO/alginate (cGO/Alg), (c) dynamic mechanical analysis (storage modulus (E′) and loss tangent (tan δ)) (e) morphology after swelling in water for 3 hours. Modified from Ref [53].

Other derivatives of graphene, such as carbon nanotubes (CNTs), discovered by Iijima [54], in the form of single wall carbon nanotubes (SWCNTs) and multi-wall carbon nanotubes (MWCNTs), as well as carbon nanofibres (CNFs) are being used for reinforcing and enhancement of many other hydrogels’ properties [55, 56, 57, 58, 59].

Reinforcement can be also conducted with plant fiber-based nanofibres by a successful fibrillation of wood pulp fibers into nanofiber bundles, which are thin enough to work as well as bacterial cellulose in maintaining the transparency of materials [60]. Other novel nanocomposite hydrogels such as those prepared with polyacrylamide (PAM) as a matrix material reinforced with natural chitosan nanofibres via in situ free-radical polymerization showed that these nanofibres acted as a multifunctional cross-linker and a reinforcing agent in the hydrogel polymer system producing a compression strength and a storage modulus significantly higher than those of the pure hydrogel [61].

Other nanoparticles such as clay have been employed to reinforce hydrogels such as polyvinyl alcohol (PVA) [62]. These polymer-clay nanocomposite hydrogels, fabricated for wound healing, constitute a class of materials in which the polymer matrix is reinforced by uniformly dispersed inorganic particles (usually 10 wt.% or less) having at least one dimension in the nanometer scale and exhibiting superior mechanical and thermal properties when compared to pure polymer or conventional composites [63].

Advertisement

3. Electrical properties

The electrical properties are very important in some biomedical areas because it has been demonstrated that various types of electrical stimulation can regulate cell physiological activities such as division [64], migration [65], differentiation and cell death [66]. The electrical stimulation has also been useful in promoting healing for spinal cord repair and cancer therapy due to its non-invasiveness of these polymers [67, 68, 69]. Therefore, much emphasis is being done in the development of new conductive hydrogels for biomedical applications. Graphene has been considered to be very effective electrode material due to its excellent conductivity [44], but its production is still very expensive and more new composite materials are expected to be developed with its derivative graphene oxide. However, GO has a very low conductivity due to their oxygen-containing functional groups and must be modified to reduce graphene oxide (rGO) in order to develop electrically conductive hydrogels. Thus, for example, following a single-step procedure starting from a homogeneous water dispersion of GO, it is possible to undergo reduction induced by the UV radiation during the photopolymerization of a resin [70]. Recently, transparent conductive films have been produced by grafting poly(acryl amide)/poly(acrylic acid) on the GO surface followed by a reduction to rGO nanosheets by a two-step chemical reduction with increased conductivity [71] and inorganic–organic double-network (DN) conductive hydrogel of rGO and poly(acrylic acid) has been prepared by a two-step synthesis with a reduction-induced in situ self-assembly [72]. Even more recently, a nacre-inspired nanocomposite of rGO and PAA has been prepared via a vacuum-assisted filtration self-assembly process (see Figure 3). The abundant hydrogen bonding between GO and PAA results in both high strength and toughness of the bioinspired nanocomposites, which are higher than that of pure reduced GO. Moreover, this composite also displays high electrical conductivity, which renders it very promising material in many biomedical applications such as flexible electrodes and artificial muscles.

Figure 3.

Fabrication process of rGO–PAA composites: (a) the GO nanosheets/PAA homogeneous solution was filtered by vacuum-assisted filtration into GO–PAA composites. Then after hydroiodic acid (HI) reduction, the rGO–PAA composites were obtained. (b) a digital photograph of rGO–PAA composites (c) and (d) cross-section surface morphology with different magnifications of rGO–PAA composites. Reprinted with permission from Ref. [73].

Carbon nanotubes (CNTs) have also been attracting intensive attention because of their excellent electrical properties with a superb conductivity, remarkable mechanical properties with many potential technological applications [74]. CNTs offer the possibility of developing ultrasensitive electrochemical biosensors due to their unique electrical properties. Thus, nanofibrous membranes filled with multiwalled carbon nanotubes (MWCNT) have been electrospun from the mixture of poly(acrylonitrile-co-acrylic acid) (PANCAA) and MWCNT to develop a glucose biosensor for diabetics [75]. Other hydrogels with very promising biomedical applications consists of dielectrophoretically aligned carbon nanotubes, which control electrical and mechanical properties of gelatin methacrylate (GelMA) hydrogels [76]. The contractile muscle cells cultured on these materials demonstrated higher maturation compared with cells cultured on pristine and randomly distributed CNTs in GelMA hydrogels.

Advertisement

4. Thermal properties

Even though hydrogels do not need to endure temperatures higher than that of the human body, the improvement of thermal properties can increase its long-term operation. Thus, for example, the incorporation of polyurethane into polyacrylamide network in the form of an interpenetrating polymer network enhanced the thermal properties due to higher crosslinking density imparted by the hard segment content [22]. Though silica can improve the mechanical properties of hydrogels, the differential scanning calorimetry results of PHEMA/SiO2 hybrids are complicated, showing two glass transition temperatures (Tg) [77]. However, composite hydrogels with functionalized graphene sheets (FGNS) showed an unprecedented shift in Tg of up to 40 and 30°C in poly(acrylonitrile) with 1 wt. % of this nanomaterial [78].

Another strategy to improve the thermal properties of hydrogels is by nanoparticle filling. Thus, crosslinking metal nanoparticles added into the polymer backbone of PHEMA hydrogels enabled the preparation of thermally stable, soft, magnetic field-driven actuators with muscle-like flexibility [79]. Furthermore, thermal degradation can also be improved by this filling procedure. For example, the mechanical and thermal properties of a renewable and biocompatible hydrogel of gelatin were improved through cross-linking by cellulose nanowhiskers [80].

In the biomedical field, hydrogels are hydrophilic polymers, which are able to absorb large amounts due to contact with cells or tissue in the human body. Therefore, the thermal analysis of water and its influence on the swollen hydrogel properties becomes essential [12, 81, 82].

Advertisement

5. Water sorption and diffusion

Water sorption and diffusion of hydrogels are also very important in biomedical applications because these properties play a very important role in cell survival, especially in tissue engineering [5]. Thus, synthetic hydrogels such as PHEMA or PHEA are very important hydrophilic materials as these polymers were able to absorb and swell retaining large amounts of water within their structure [83, 84, 85, 86]. The excellent water sorption property has made these kind of biomaterials very promising in a wide range of biomedical applications such as controlled drug delivery, tissue engineering, wound healing, etc. [6, 87]. The ability of hydrogels to absorb water arises from hydrophilic functional groups attached to the polymeric backbone, while their resistance to dissolution arises from cross-links between network chains [88]. However, these single-network hydrogels have weak mechanical properties in the swollen state and slow response at swelling. Therefore, although reinforcement of hydrogels is absolutely necessary, as already mentioned, the improvement of mechanical properties can significantly affect water sorption. For example, water sorption can be dramatically reduced by the reinforcement produced by the combination of hydrophilic and hydrophobic functional groups of polymers as multicomponent polymeric systems (Figure 1).

Reinforcement of hydrogels by GO loading can enhance significantly water sorption and diffusion. Thus, the swelling rates of graphene oxide / poly(acrylic acid-co-acrylamide) nanocomposite hydrogels increased with increase in the GO loadings to 0.30 wt. % and then decreased with further increasing GO loadings. It is worth noting that the hydrogel with only 0.10 wt. % GO exhibited significant improvement of swelling capacity in neutral medium, and could also retain relatively higher swelling rates to a certain degree in acidic and basic solutions. Furthermore, it has been reported very recently that a very low filling of GO can produce a very significant increase of water diffusion (almost 6 times faster) in crosslinked alginate (Figure 4) [48]. Therefore, these GO-based super-absorbent hydrogels have very potential applications in many fields such as biomedical engineering and hygienic products [50].

Figure 4.

Cryo-scanning electron micrograph of crosslinked alginate synthesized with a minuscule amount of GO and 18 wt.% of calcium chloride (with respect to the mass of sodium alginate) in the swollen state after 2 minutes of immersion in water at 24 ± 0.5°C (a). Apparent diffusion coefficients of liquid water (mean ± standard deviation) in calcium alginate hydrogels with different crosslinker contents with (black columns) and without (gray columns) 0.1 wt.% of GO (b). Reprinted with permission from Ref [48].

The mechanism of water diffusion [89] can also be altered by the reinforcement of hydrogels through any of the methods shown in Section 2. Thus, very promising biomaterials for drug-releasing such as poly(acrylic acid)-GO composite hydrogels exhibit non-Fickian anomalous diffusion and their deswelling ratio decreases with increasing GO content [51]. Superabsorbent polymers of sodium lignosulfonate-grafted poly(acrylic acid-co-acrylamide), prepared by a new ultrasound synthetic method, shows also a non-Fickian water diffusion transport with a maximum water absorbency of 1350 g·g−1 [90]. PHEA hydrogels exhibit also a non-Fickian diffusion behavior s [83, 86]. However, other polymer chemically very similar, PHEMA, which is a very important water-swellable biomedical polymer, is controlled by Fickian diffusion [91]. Thus, copolymerized hydrogels based on 2-hydroxyethyl methacrylate (HEMA) and epoxy methacrylate (EMA) synthesized by bulk polymerizations showed that the swelling process of these polymers is also Fickian and the equilibrium water content (EWC) decreased with increasing EMA content due to its hydrophobicity [92].

It is remarkable that pH has a big influence in the swelling properties and diffusion mechanism of hydrogels. Thus, the swelling properties of semi-interpenetrating polymer networks of acrylamide-based polyurethanes decreased in acidic pH while a reverse trend was observed in basic pH. Nevertheless, these semi-IPNs were found to be hydrolytically stable in phosphate buffer solution, which render them potential materials for biomedical applications [22]. PAA is a pH-sensitive and biocompatible polymer that is being used in many biomedical fields [30] and has attracted considerable interest because of its therapeutic use, due to its ability to swell reversibly with changes in pH. Thus, GO functionalized with PAA (GO-PAA) by in situ atom transfer radical polymerization (ATRP) showed potential use as an intracellular protein carrier using bovine serum albumin (BSA) as a model protein [93]. This application is very important because proteins participate in all vital body processes and these perform an essential function inside cells as enzymes, transduction signals, and gene regulation. Another pH-sensitive terpolymer hydrogel, poly(acryl amide-co-2-acrylamido-2-methyl-1-propanesulfonic acid-co-acrylamido glycolic acid), with applications in drug release showed a quasi-Fickian diffusion mechanism with partly chain relaxation controlled diffusion. These hydrogels demonstrated a sharp change in its water absorbency and molecular weight between crosslinks of the network with a change in the swelling media pH [94].

The effect of temperature on swelling properties of hydrogels is also very important [92]. Thus, hydrogels can be modified to exhibit fast temperature sensitivity, and improved oscillating swelling-deswelling properties as for example the thermosensitive poly(N-isopropyl acrylamide-co-acrylic acid) hydrogels [95].

Advertisement

6. Antimicrobial and antifouling activity

Microbial infections can lead to implant failure, which may cause major economic losses and suffering among patients despite the use of preoperative antibiotic prophylaxis and the aseptic processing of materials. Therefore, novel antimicrobial materials are urgently in need for medical uses [96]. For that reason, much effort is being done in the development of advanced hydrogels with inherent antimicrobial properties. Thus, syringe-injectable bioadhesive hydrogels prepared from mixing polydextran aldehyde and branched polyethylenimine, able to kill both Gram-negative and Gram-positive bacteria, while sparing human erythrocytes [97] and injectable conductive self-healed hydrogels based on quaternized chitosan-g-polyaniline (QCSP) and benzaldehyde group functionalized poly(ethylene glycol)-co-poly(glycerol sebacate) (PEGS-FA) with antibacterial, anti-oxidant and electroactive dressing for cutaneous wound healing have been developed [98].

Other hydrogels such as chitosan and its derivatives has been widely used as implant coatings for its intrinsic properties such as non-toxic, osteoconductive, pH responsive, anti-microbial, biocompatible and cell adhesive [99, 100]. Nevertheless, the development of new chitosan derivatives or composites with superior antimicrobial activity is still under research. Thus, for example, a novel hydrogel coating produced by electrophoretic co-deposition of chitosan/alkynyl chitosan showed high antibacterial effect against Escherichia coli and Staphylococcus aureus [101] by disk diffusion method [102] (see Figure 5). Antibacterial polymer coating adhered on the surface of medical implants and devices have attracted great interests in the last decades for its ability to reduce implant-associated infections [103, 104].

Figure 5.

Antimicrobial results of chitosan and alkynyl chitosan against E. coli and S. aureus by using disk diffusion method. Images (a–e) are paper disks containing 1 wt. % chitosan, 1 wt. % alkynyl chitosan (ACS1, ACS2, ACS3 and ACS4), respectively against E. coli; images (f–j) correspond to paper disks containing 1 wt. % chitosan, 1 wt. % alkynyl chitosan (ACS1, ACS2, ACS3 and ACS4), respectively against S. aureus. Alkynyl chitosan coded as ACS1, ACS2, ACS3 and ACS4 were prepared by changing the molar ratio of chitosan 0 unit to 3-bromopropyne monomer as 1:0.5, 1:1, 1:1.5 and 1:2. Reprinted with permission from Ref [101].

Antimicrobial hydrogels formed by crosslinking polyallylamine with aldaric acid derivatives exhibited complete kill within 4 hours against Pseudomonas aeruginosa, Escherichia coli, Staphylococcus aureus and Candida albicans suspended in culture medium [105] and a facile strategy to fabricate antibacterial ultrathin hydrogel films via a layer-by-layer (LbL) technique and “click” chemistry was reported by Wang et al. [106]. This hydrogels consisted of poly[oligo(ethylene glycol)fumarate]-co-poly[dodecyl bis(2-hydroxyethyl)methylammonium fumarate] (POEGDMAM) containing multi-enes and poly[oligo(ethylene glycol)mercaptosuccinate] (POEGMS). These ultrathin films exhibited excellent antibacterial activity against both Staphylococcus aureus and Escherichia coli due to the presence of ammonium groups with long alkyl chains in the POEGDMAM.

Antibacterial properties can also be imparted to a hydrogel by doping in an exogenous antibiotic for eventual release [107]. In these delivery systems, the active agent is released from the polymer matrix over time. However, the material’s antibiotic activity is eventually exhausted with the remaining matrix being left inactive and the remaining vehicle may become a substrate for colonization by bacterial biofilms once the payload is depleted, which can become life threatening. For this reason, secondary surgeries are typically performed to remove these empty depots as a means of preventing this type of infection. To avoid this second surgery, a hydrogel drug delivery system in which the drug release rate of vancomycin and degradation rate of the hydrogel are linked via covalent incorporation of vancomycin in the hydrogel backbone was successfully developed [108].

However, many hydrogels themselves do not have any antimicrobial activity and therefore some fillers, and antimicrobial agents need to be incorporated by physical blending in order to produce antimicrobial materials [109]. Thus, graphene has emerged as a novel green broad-spectrum antimicrobial nanomaterial, with little bacterial resistance and tolerable cytotoxic effect on mammalian cells. It exerts its antibacterial action via physical damages through direct contact of its sharp edges with bacterial membranes and destructive extraction of lipid molecules. The antimicrobial activity of GO against two bacterial pathogens (Pseudomonas syringae and Xanthomonas campestris pv. undulosa) and two fungal pathogens (Fusarium graminearum and Fusarium oxysporum) showed that GO had a powerful effect on the reproduction of all these four pathogens because it killed nearly 90% of the bacteria and repressed 80% macroconidia germination along with partial cell swelling and lysis at 500 μgmL−1. The graphene-based nanocomposites have a wide range of biomedical applications, such as wound dressing due to its superior antimicrobial properties and good biocompatibility [110].

Another strategy to design hydrogels with desired antimicrobial performance consists of adding silver nanoparticles (Ag NPs). Hence, silver nanoparticles have emerged up with diverse medical applications ranging from silver based dressings, silver coated medicinal devices, such as nanogels, nanolotions, etc. [111].

Infections are also frequent and highly undesired occurrences after orthopedic procedures. Thus, for example, medicated hydrogels of hyaluronic acid derivatives have been developed [112] to address this problem. However, the growing concern caused by the rise in antibiotic resistance is progressively dwindling the efficacy of such drugs and the integration of silver nanoparticles in hydrogels has become a very promising alternative [113].

The combination of both previous strategies (graphene and Ag NPs) to design antimicrobial hydrogels with good water maintaining ability is of particular significance to promote the development of wound dressing. Thus, a series of hydrogels were synthesized by crosslinking of Ag/graphene composites with acrylic acid and N,N′-methylene bisacrylamide at different mass ratios. In this study, prepared hydrogel with an optimal Ag to graphene mass ratio of 5:1 exhibited much stronger antibacterial abilities than other hydrogels and showed excellent biocompatibility, high swelling ratio, and good extensibility at the same time. Besides, in vivo experiments indicated that this nanocomposite hydrogel could significantly accelerate the healing rate of artificial wounds in rats, and it helped to successfully reconstruct intact and thickened epidermis during 15 day of healing of impaired wounds [114]. In the same way, acrylic acid (AA) grafted onto poly(ethylene terephthalate) (PET) film through gamma-ray induced graft copolymerization with silver nanoparticles on the surface showed strong and stable antibacterial activity [115].

It is highly desired to have a hydrogel material bearing the excellent antifouling property/biocompataiblity to prolong the lifetime of implanted materials, switchable antimicrobial property to eliminate infection and inflammation, and good mechanical properties to avoid the failure of the implanted material. We hypothesize derivatives of zwitterionic carboxybetaine with hydroxyl group(s) can switch between the lactone form (anti-microbial) and the zwitterionic form (anti-fouling) and the intramolecular hydrogen bonds will enhance the mechanical property of the zwitterionic hydrogel. It is highly desired to have a hydrogel material bearing the excellent antifouling property/biocompataiblity to prolong the lifetime of implanted materials, switchable antimicrobial property to eliminate infection and inflammation, and good mechanical properties to avoid the failure of the implanted material. We hypothesize derivatives of zwitterionic carboxybetaine with hydroxyl group(s) can switch between the lactone form (anti-microbial) and the zwitterionic form (anti-fouling) and the intramolecular hydrogen bonds will enhance the mechanical property of the zwitterionic hydrogel.

On the other hand, the surface of hydrogels must be modified to make it resistant to protein adsorption and cell adhesion to avoid fouling. Thus, there is a need for coatings with antifouling properties that are able to improve the performances of implanted biomedical devices. Thus, the antifouling properties of poly(2-hydroxyethyl methacrylate-co-methyl methacrylate) hydrogels were improved by the surface grafting of a brush of poly(oligoethylene glycol methyl ether acrylate) [poly(OEGA)] [116]. Novel antifouling highly wettable hydrogels with superior mechanical and self-healing properties have also been developed by UV-initiated copolymerization of non-fouling zwitterionic carboxybetaine methacrylamide (CBMAA-3) and 2-hydroxyethyl methacrylate (HEMA) in the presence of uniformly dispersed clay nanoparticles (Laponite XLG) in water [117].

Therefore, it would be highly desired to have a hydrogel material bearing the excellent antifouling property/biocompataiblity to prolong the lifetime of implanted materials, switchable antimicrobial property to eliminate infection and inflammation, and good mechanical properties to avoid the failure of the implanted material. Thus, derivatives of zwitterionic carboxybetaine have been developed with hydroxyl group(s), which can switch between the lactone form (antimicrobial) and the zwitterionic form (anti-fouling) [118]. Besides, the intramolecular hydrogen bonds enhance the mechanical property of the zwitterionic hydrogel.

Nevertheless, the rapid emergence of antibiotic resistance in pathogenic microbes is becoming an imminent global public health problem because they are highly prone to develop resistance through mutation and the treatment with conventional antibiotics often leads to resistance development leaving the bacterial morphology intact. Therefore, much research is currently being done in the development of new antimicrobial hydrogels because they have been demonstrated to be very effective in preventing and treating multidrug-resistant infections.

Advertisement

7. Porosity

Porous polymers have received an increased level of research interest because of their potential to merge the properties of both porous materials and polymers [119]. Porous polymers have potential applications in many fields such as gas storage and separation materials [120, 121], drug delivery [122], catalysts [123], supports for electrochemical sensing [124], low-dielectric constant materials [125], packing materials in chromatography [126], scaffolds or three-dimensional porous matrices for tissue engineering in regenerative medicine [5, 41, 127, 128] and many others. These high value applications have driven much emphasis on development of reliable methods for preparation of porous polymers with designed pore architectures in the last decades (see Figure 6).

Figure 6.

Different porous polymers prepared by different preparation methods to obtain diverse pore architectures. Reprinted with permission from Ref [12, 41, 48, 128, 129, 130].

Tissue engineering holds great promise for regeneration and repair of diseased tissues, making the development of new porous supports as scaffolds for tissue regeneration a topic of great interest in biomedical research. Hydrogels have emerged as leading candidates for engineered tissue scaffolds due to their good biocompatibility and similarities to native extracellular matrix. However, precise control of hydrogel properties, such as high porosity, remains a challenge. Traditional techniques for creating bulk porosity in polymers have demonstrated success in hydrogels for tissue engineering. However, some problems related to direct cell encapsulation often occur. Thus, emerging technologies have demonstrated the ability to control porosity and morphology in hydrogels, creating engineered tissues with structure and function similar to native tissues [131]. The interconnection and geometry of pores, which depends on the tissue to regenerate, physicochemical properties, and mechanical resistance of the material play a major role in these biomedical applications. Thus, there are several methods to produce scaffolds, which include gas foaming [132], sintering fiber meshes [133], solvent casting [134], polymerization in solution [86, 135], porogen technique [129, 136], freeze-drying techniques [137, 138], electrospinning [139], 3D printing [140], 3D bioplotting of scaffold with cells [141], etc. For example, scaffolds with interconnected spherical pores and controlled hydrophilicity with interconnected porous structure were synthesized using a template of sintered PMMA microspheres of controlled size. In these scaffolds, the geometric characteristics (pore size, connectivity and porosity) and the physico-chemical properties of the resulting material can be controlled in an independent way. Copolymerization of hydrophobic ethyl acrylate (EA) and hydrophilic hydroxyethyl methacrylate comonomers in the free space of the template and subsequent solution of the PMMA microspheres gave rise to the scaffold with the designed pore architecture (see Figure 7) [129].

Figure 7.

Scanning electron micrographs of EA/HEMA copolymer scaffolds (30% HEMA) at different magnifications. Reprinted with permission from Ref. [129].

A novel preparation of gelatin-PHEMA porous scaffolds by freeze-drying technique was developed recently [138]. Their morphology was assessed by SEM and μ-CT (Figure 8). In this study, four types of novel hydrogels using different methacrylamide-modified gelatin/2-hydroxyethyl methacrylate ratios between 1/0 and 1/2 (w/w) (samples from C0 to C3) were prepared and the results indicated that the HEMA content in the initial polymerization mixtures modulate the architecture of the porous scaffolds from straightforward, top-to-bottom oriented channels for hydrogels possessing the lowest HEMA content to a complex and dense internal porosity of the channels the case of higher HEMA loaded materials. Besides, the covalently bound gelatin sequences significantly improve the biocompatibility of PHEMA based hydrogels, which is very desirable for tissue engineering purposes.

Figure 8.

Morphology of the gelatin-PHEMA porous scaffolds as obtained through μ-CT (panels (a)–(c): (a)—top view, (b)—bottom view, and (c)—side view) and SEM (panel (d)) analyses: panel (I)—C0; panel (II)—C1; panel (III)—C2; panel (IV)—C3. Reprinted with permission from Ref. [138].

Superporous scaffolds can be also prepared by the salt-leaching technique using NaCl or (NH4)2SO4 as a porogen [142] or with many other porogenic agents such as ammonium oxalate crystals [143].

By submitting carbon dioxide to supercritical conditions after certain time and then rapidly depressurized is also possible to fabricate porous structures that are related to the supercritical parameters and to the polymer blend composition [131]. The use of CO2 to create such scaffolds has received some attention in the past but many researchers believe that there is limited interconnectivity between the pores, which is required for tissue engineering. However, highly porous (greater than 85%) and well interconnected scaffolds with very promising applications for cartilage repair have been obtained in a blend of poly(ethyl methacrylate) and tetrahydrofurfuryl methacrylate [144].

Probably the most sophisticated techniques to produce scaffolds are electrospinning, 3D printing and bioprinting. Electrospinning is composed of a high-voltage DC power supply, an infusion pumps and a syringe with a needle tip usually with a diameter of 0.5 mm. For example, a three-dimensional aligned nanofibers-collagen type I hydrogel scaffold for controlled non-viral drug/gene delivery to direct axon regeneration in spinal cord injury treatment has been reported very recently [145].

3D printing promises to produce complex biomedical devices according to computer design using patient-specific anatomical data. This 3D printing technique has slowly evolved to create one-of-a-kind devices, implants, scaffolds for tissue engineering, and drug delivery systems among other important applications. However, several technological limitations, related to the kind of commercially printable materials available and other technical printing aspects such as printing speed, must still be overcome. The common 3D printing technologies are three-dimensional printing, fused deposition modeling, selective laser sintering, stereolithography, and 3D plotting/direct-write/bioprinting, and are still under deep research for the progress of each technology applied in tissue engineering. Bioprinting is the more advanced 3D printing technology because it consists of printing cells combined with custom 3D scaffolds for personalized regenerative medicine [140].

Mechanical resistance depends both on the material properties and on the interconnected pore structure of the scaffold. This problem is more important in polymer hydrogels, which exhibit even poorer mechanical properties when they are porous and hydrated. Therefore, it is usually necessary to enhance the mechanical properties of these porous structures by means of the methods, shown in chapter 2, with nanomaterials or other techniques. For example, the use of a hybrid hydrogel nanocomposite of silica/PHEA as scaffold material matrix greatly improves the mechanical properties [41].

Other modifications of scaffolds such as those of PHEMA with cholesterol methacrylate (CHLMA) and laminin have been developed in the presence of ammonium oxalate crystals to introduce interconnected superpores in the matrix in order to design superporous scaffolds that promote cell-surface interaction [146]. PHEMA has also been modified with laminin-derived Ac-CGGASIKVAVS-OH peptide sequences to construct scaffolds that promote cell adhesion and neural differentiation. With the same goal, nanofiber scaffolds of poly (L-lactide) (PLLA) prepared by electrospinning were treated with oxygen plasma and then simultaneously in situ grafted with hydrophilic acrylic acid to obtain PLLA-g-PAA with a modified surface, which significantly improved cell adhesion and proliferation [130].

Polysaccharide hydrogels have become increasingly studied as matrices in soft tissue engineering due to their known cytocompatibility. For example, cross-linkable dextran methacrylates and hyaluronan methacrylate hydrogels, which are candidates as matrices for soft tissue reconstruction, were synthesized showing that the in vitro degradation behavior of these types of hydrogels could be controlled by the polysaccharide structure and the cross-linking density. Furthermore, under in vitro conditions, these novel materials had no cytotoxic effects against fibroblasts and the use of composite gels improved the adherence of cells [147].

Therefore, great advances have been achieved so far in scaffold design of new advanced porous hydrogels for tissue engineering applications. Nevertheless, much research has to be conducted still in order to find new ways and methods capable of providing suitable materials able to fulfill all the necessary requirements of this biomedical field.

Advertisement

Acknowledgments

This work was supported by the 2017-231-001UCV grant from the Universidad Católica de Valencia “San Vicente Mártir”.

References

  1. 1. Buwalda SJ, Vermonden T, Hennink WE. Hydrogels for therapeutic delivery: Current developments and future directions. Biomacromolecules. 2017;18(2):316-330
  2. 2. Chirila T, Harkin D. Biomaterials and Regenerative Medicine in Ophthalmology. 2nd ed. Sawston, Cambridge: Elsevier; 2016
  3. 3. Shalaby S, Nagatomi S, Peniston S. Polymeric Biomaterials for Articulating Joint Repair and Total Joint Replacement. Polymers for Dental and Orthopedic Applications. Boca Raton: CRC Press; 2007
  4. 4. Kamoun EA, Kenawy E-RS, Chen X. A review on polymeric hydrogel membranes for wound dressing applications: PVA-based hydrogel dressings. Journal of Advanced Research. 2017;8(3):217-233
  5. 5. Van Blitterswijk C. De Boer J, Tissue Engineering. London: Academic Press; 2014
  6. 6. Ratner BD, Hoffman AS, Schoen FJ, Lemons JE. Biomaterials Science: An Introduction to Materials in Medicine. Canada: Academic Press; 2012
  7. 7. Stoy V, Climent C. In Hydrogels: Speciality Plastics for Biomedical and Pharmaceutical Applications. Basel: Technomic Publishers; 1996
  8. 8. Serrano-Aroca Á, Monleón-Pradas M, Gómez-Ribelles JL. Effect of crosslinking on porous poly(methyl methacrylate) produced by phase separation. Colloid & Polymer Science. 2008;286(2):209-216
  9. 9. Ramaraj B, Radhakrishnan G. Modification of the dynamic swelling behaviour of poly(2-hydroxyethyl methacrylate) hydrogels in water through interpenetrating polymer networks (IPNs). Polymer (Guildf). 1994;35(10):2167-2173
  10. 10. Serrano-Aroca Á, Monleón-Pradas M, Gómez-Ribelles JL. Plasma-induced polymerisation of hydrophilic coatings onto macroporous hydrophobic scaffolds. Polymer (Guildf). 2007;48(7):2071-2078
  11. 11. Serrano-Aroca Á, Gómez-Ribelles JL, Monleón-Pradas M, Vidaurre-Garayo A, Suay-Antón J. Characterisation of macroporous poly(methyl methacrylate) coated with plasma-polymerised poly(2-hydroxyethyl acrylate). European Polymer Journal. 2007;43(10):4552-4564
  12. 12. Serrano-Aroca Á, Monleón-Pradas M, Gómez-Ribelles JL, Rault J. Thermal analysis of water in reinforced plasma-polymerised poly(2-hydroxyethyl acrylate) hydrogels. European Polymer Journal. 2015;72:523-534
  13. 13. Gilbert JL, Ney DS, Lautenschlager EP. Self-reinforced composite poly (methy1 methacrylate ): Static and fatigue properties. 1995;16(14):1043-1055
  14. 14. Amerio E, Fabbri P, Malucelli G, Messori M, Sangermano M, Taurino R. Scratch resistance of nano-silica reinforced acrylic coatings. Progress in Organic Coatings. 2008;62(2):129-133
  15. 15. Kugler S, Kowalczyk K, Spychaj T. Progress in organic coatings hybrid carbon nanotubes/graphene modified acrylic coats. Progress in Organic Coatings. 2015;85:1-7
  16. 16. Cha C, Shin SR, Gao X, Annabi N, Dokmeci MR, Tang XS, et al. Controlling mechanical properties of cell-laden hydrogels by covalent incorporation of graphene oxide. Small. 2014;10(3):514-523
  17. 17. Faghihi S, Gheysour M, Karimi A, Salarian R. Fabrication and mechanical characterization of graphene oxide-reinforced poly (acrylic acid)/gelatin composite hydrogels. Journal of Applied Physics. 2014;115(83513)
  18. 18. Ha H, Shanmuganathan K, Ellison CJ. Mechanically stable thermally crosslinked poly(acrylic acid)/reduced graphene oxide aerogels. ACS Applied Materials & Interfaces. 2015;7(11):6220-6229
  19. 19. Gong JP, Katsuyama Y, Kurokawa T, Osada Y. Double-network hydrogels with extremely high mechanical strength. Advanced Materials. 2003;15(14):1155-1158
  20. 20. Sperling LH, Mishra V. The current status of interpenetrating polymer networks. Polymers for Advanced Technologies. 1996;7(4):197-208
  21. 21. Maity J, Ray SK. Enhanced adsorption of methyl violet and congo red by using semi and full IPN of polymethacrylic acid and chitosan. Carbohydrate Polymers. 2014;104(1):8-16
  22. 22. Merlin DL, Sivasankar B. Synthesis and characterization of semi-interpenetrating polymer networks using biocompatible polyurethane and acrylamide monomer. European Polymer Journal. 2009;45(1):165-170
  23. 23. Wang M, Pramoda KP, Goh SH. Mechanical behavior of pseudo-semi-interpenetrating polymer networks based on double-C 60 -end-capped poly ( ethylene oxide ) and poly (methyl methacrylate ). Chemistry of Materials. 2004;16:3452-3456
  24. 24. LQ X, Yao F, GD F, Kang ET. Interpenetrating network hydrogels via simultaneous “click chemistry” and atom transfer radical polymerization. Biomacromolecules. 2010;11(7):1810-1817
  25. 25. Burugapalli K, Koul V, Dinda AK. Effect of composition of interpenetrating polymer network hydrogels based on poly(acrylic acid) and gelatin on tissue response: A quantitative in vivo study. Journal of Biomedical Materials Research. Part A. 2004;68(2):210-218
  26. 26. Shams Es-haghi S, Weiss RA. Finite strain damage-elastoplasticity in double-network hydrogels. Polymer (United Kingdom). 2016;103:277-287
  27. 27. Dragan ES. Design and applications of interpenetrating polymer network hydrogels. A review. Chemical Engineering Journal. 2014;243:572-590
  28. 28. Lee J, Kim S, Kim S, Lee Y, Lee K. Synthesis and characteristics of interpenetrating polymer network hydrogel composed of chitosan and poly ( acrylic acid). Polymer (Guildf). 1998:113-120
  29. 29. Bajpai AK, Shukla SK, Bhanu S, Kankane S. Responsive polymers in controlled drug delivery. Progress in Polymer Science. 2008;33(11):1088-1118
  30. 30. Zhao Y, Kang J, Tan T. Salt-, pH- and temperature-responsive semi-interpenetrating polymer network hydrogel based on poly(aspartic acid) and poly(acrylic acid). Polymer (Guildf). 2006;47(22):7702-7710
  31. 31. Richter A, Paschew G, Klatt S, Lienig J, Arndt K-F, Adler H-JP. Review on hydrogel-based pH sensors and microsensors. Sensors. 2008;8(1):561-581
  32. 32. DeKosky BJ, Dormer NH, Ingavle GC, Roatch CH, Lomakin J, Detamore MS, et al. Hierarchically designed agarose and poly(ethylene glycol) interpenetrating network hydrogels for cartilage tissue engineering. Tissue Engineering. Part C, Methods. 2010;16(6):1533-1542
  33. 33. Stumpel JE, Gil ER, Spoelstra AB, Bastiaansen CWM, Broer DJ, Schenning APHJ. Stimuli-responsive materials based on interpenetrating polymer liquid crystal hydrogels. Advanced Functional Materials. 2015;25(22):3314-3320
  34. 34. Groover MP. Fundamentals of Modern Manufacturing_Materials, Processes, and Systems. 4th ed. United States of America: J. Wiley & Sons; 2010
  35. 35. Young CD, JR W, Tsou TL. High-strength, ultra-thin and fiber-reinforced pHEMA artificial skin. Biomaterials. 1998;19(19):1745-1752
  36. 36. Saheb DN, Jog JP. Natural fiber polymer composites: A review. Advances in Polymer Technology. 1999;18(4):351-363
  37. 37. Ku H, Wang H, Pattarachaiyakoop N, Trada M. A review on the tensile properties of natural fiber reinforced polymer composites. Composites. Part B, Engineering. 2011;42(4):856-873
  38. 38. Chen C, Li D, Hu Q, Wang R. Properties of polymethyl methacrylate-based nanocomposites: Reinforced with ultra-long chitin nanofiber extracted from crab shells. Materials and Design. 2014 Apr;56:1049-1056
  39. 39. Kokubo T. Design of bioactive bone substitutes based on biomineralization process. Materials Science and Engineering: C. 2005;25(2):97-104
  40. 40. Hench LL, West JK. The sol-gel process. Chemical Reviews. 1990;90(1):33-72
  41. 41. Rodríguez-Hernández JC, Serrano-Aroca Á, Gómez-Ribelles JL, Monleón-Pradas M. Three-dimensional nanocomposite scaffolds with ordered cylindrical orthogonal pores. Journal of Biomedical Materials Research - Part B Applied Biomaterials. 2008;84(2):541-549
  42. 42. Lin H-R, Ling M-H, Lin Y-J. High strength and low friction of a PAA-alginate-silica hydrogel as potential material for artificial soft tissues. Journal of Biomaterials Science. Polymer Edition. 2009;20(5-6):637-652
  43. 43. Novoselov KS, Geim AK, Morozov SV, Jiang D, Zhang Y, Dubonos SV, et al. Electric field effect in atomically thin carbon films. Science. 2004;306(5696):666-669
  44. 44. Geim AK, Novoselov KS. The rise of graphene. Nature Materials. 2007;6:183-191
  45. 45. Balandin AA, Ghosh S, Bao W, Calizo I, Teweldebrhan D, Miao F, et al. Superior thermal conductivity of single-layer graphene. Nano Letters. 2008;8(3):902-907
  46. 46. Lee C, Wei X, Kysar JW, Hone J. Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science. 2008;321(5887):385-388
  47. 47. Kalbacova M, Broz A, Kong J, Kalbac M. Graphene substrates promote adherence of human osteoblasts and mesenchymal stromal cells. Carbon N Y. 2010;48(15):4323-4329
  48. 48. Serrano-Aroca Á, Ruiz-Pividal JF, Llorens-Gámez M. Enhancement of water diffusion and compression performance of crosslinked alginate with a minuscule amount of graphene oxide. Scientific Reports. 2017;7:11684
  49. 49. Shen J, Yan B, Li T, Long Y, Li N, Ye M. Study on graphene-oxide-based polyacrylamide composite hydrogels. Composites. Part A, Applied Science and Manufacturing. 2012;43(9):1476-1481
  50. 50. Huang Y, Zeng M, Ren J, Wang J, Fan L, Xu Q. Preparation and swelling properties of graphene oxide/poly(acrylic acid-co-acrylamide) super-absorbent hydrogel nanocomposites. Colloids Surfaces A Physicochemical and Engineering Aspects. 2012;401:97-106
  51. 51. Shen J, Yan B, Li T, Long Y, Li N, Ye M. Mechanical, thermal and swelling properties of poly(acrylic acid)-graphene oxide composite hydrogels. Soft Matter. 2012;8(6):1831-1836
  52. 52. Xu Y, Wu Q, Sun Y, Bai H, Shi G. Three-dimensional self-assembly of graphene oxide and DNA into multifunctional hydrogels. ACS Nano. 2010;4(12):7358-7362
  53. 53. Serrano-Aroca Á, Deb S. Synthesis of irregular graphene oxide tubes using green chemistry and their potential use as reinforcement materials for biomedical applications. PLoS One. 2017 Sep 21;12(9):e0185235
  54. 54. Iijima S. Helical microtubules of graphitic carbon. Nature. 1991;354(6348):56-58
  55. 55. Ahadian S, Ramón-Azcón J, Estili M, Liang X, Ostrovidov S, Shiku H, et al. Hybrid hydrogels containing vertically aligned carbon nanotubes with anisotropic electrical conductivity for muscle myofiber fabrication. Scientific Reports. 2014;4:4271
  56. 56. Ogoshi T, Takashima Y, Yamaguchi H, Harada A. Chemically-responsive sol-gel transition of supramol single-walled carbon nanotube SWNT hydrogel made by hybrid of SWNTs & cyclodextrin. Journal of the American Chemical Society. 2007;129:4878-4879
  57. 57. Satarkar NS, Johnson D, Marrs B, Andrews R, Poh C, Gharaibeh B, et al. Hydrogel-MWCNT nanocomposites: Synthesis, characterization, and heating with radiofrequency fields. Journal of Applied Polymer Science. 2010;117(3):1813-1819
  58. 58. Bhattacharyya S, Guillot S, Dabboue H, Tranchant JF, Salvetat JP. Carbon nanotubes as structural nanofibers for hyaluronic acid hydrogel scaffolds. Biomacromolecules. 2008;9(2):505-509
  59. 59. Meng X, Stout DA, Sun L, Beingessner RL, Fenniri H, Webster TJ. Novel injectable biomimetic hydrogels with carbon nanofibers and self assembled rosette nanotubes for myocardial applications. Journal of Biomedical Material Research - Part A. 2013;101 A(4):1095-1102
  60. 60. Iwamoto S, Nakagaito AN, Yano H, Nogi M. Optically transparent composites reinforced with plant fiber-based nanofibers. Applied Physics A: Materials Science & Processing. 2005;81(6):1109-1112
  61. 61. Zhou C, Wu Q. A novel polyacrylamide nanocomposite hydrogel reinforced with natural chitosan nanofibers. Colloids Surfaces B Biointerfaces. 2011;84(1):155-162
  62. 62. Kokabi M, Sirousazar M, Hassan ZM. PVA-clay nanocomposite hydrogels for wound dressing. European Polymer Journal. 2007;43(3):773-781
  63. 63. Meneghetti P, Qutubuddin S. Synthesis, thermal properties and applications of polymer-clay nanocomposites. Thermochimica Acta. 2006;442(1-2):74-77
  64. 64. Zhao M, Forrester JV, McCaig CD. A small, physiological electric field orients cell division. Proceedings of the National Academy of Sciences of the United States of America. 1999;96:4942-4946
  65. 65. Yao L, Shanley L, Mccaig C, Zhao M. Small applied electric fields guide migration of hippocampal neurons. Journal of Cellular Physiology. 2008 Aug;216(2):527-535
  66. 66. Woo DG, Shim MS, Park JS, Yang HN, Lee DR, Park KH. The effect of electrical stimulation on the differentiation of hESCs adhered onto fibronectin-coated gold nanoparticles. Biomaterials. 2009;30(29):5631-5638
  67. 67. Sisken BF, Walker J, Orgel M. Prospects on clinical applications of electrical stimulation for nerve regeneration. Journal of Cellular Biochemistry. 1993;51(4):404-409
  68. 68. Zhao M. Electrical fields in wound healing-an overriding signal that directs cell migration. Seminars in Cell & Developmental Biology. 2009;20(6):674-682
  69. 69. Kirson ED, Dbalý V, Tovaryš F, Vymazal J, Soustiel JF, Itzhaki A, et al. Alternating electric fields arrest cell proliferation in animal tumor models and human brain tumors. Proceedings of the National Academy of Sciences. 2007;104(24):10152-10157
  70. 70. Fabbri P, Valentini L, Bittolo Bon S, Foix D, Pasquali L, Montecchi M, et al. In-situ graphene oxide reduction during UV-photopolymerization of graphene oxide/acrylic resins mixtures. Polymer (Guildf). 2012;53(26):6039-6044
  71. 71. Huang YL, Tien HW, Ma CC, Yang SY, SY W, Liu HY, et al. Effect of extended polymer chains on properties of transparent graphene nanosheets conductive film. Journal of Materials Chemistry. 2011;21(45):18236
  72. 72. Huang P, Chen W, Yan L. An inorganic-organic double network hydrogel of graphene and polymer. Nanoscale. 2013;5(13):6034-6039
  73. 73. Wan S, Hu H, Peng J, Li Y, Fan Y, Jiang L, et al. Nacre-inspired integrated strong and tough reduced graphene oxide-poly(acrylic acid) nanocomposites. Nanoscale. 2016;8(10):5649-5656
  74. 74. Rao CNR, Satishkumar BC, Govindaraj A. Nanotubes. Chemphyschem. 2001;2:78-105
  75. 75. Wang ZG, Wang Y, Xu H, Li G, Xu ZK. Carbon nanotube-filled Nanofibrous membranes electrospun from poly(acrylonitrile- co -acrylic acid) for glucose biosensor. Journal of Physical Chemistry C. 2009;113(7):2955-2960
  76. 76. Ramón-Azcón J, Ahadian S, Estili M, Liang X, Ostrovidov S, Kaji H, et al. Dielectrophoretically aligned carbon nanotubes to control electrical and mechanical properties of hydrogels to fabricate contractile muscle myofibers. Advanced Materials. 2013;25(29):4028-4034
  77. 77. Huang ZH, Qiu KY. The effects of interactions on the properties of acrylic polymers/silica hybrid materials prepared by the in situ sol-gel process. Polymer (Guildf). 1997;38(3):521-526
  78. 78. Ramanathan T, Abdala AA, Stankovich S, Dikin DA, Herrera-Alonso M, Piner RD, et al. Functionalized graphene sheets for polymer nanocomposites. Nature Nanotechnology. 2008;3(6):327-331
  79. 79. Fuhrer R, Athanassiou EK, Luechinger NA, Stark WJ. Crosslinking metal nanoparticles into the polymer backbone of hydrogels enables preparation of soft, magnetic field-driven actuators with muscle-like flexibility. Small. 2009;5(3):383-388
  80. 80. Dash R, Foston M, Ragauskas AJ. Improving the mechanical and thermal properties of gelatin hydrogels cross-linked by cellulose nanowhiskers. Carbohydrate Polymers. 2013;91(2):638-645
  81. 81. Roorda WE, Bouwstra JA, de Vries MA, Junginger HE. Thermal analysis of water in p(HEMA) hydrogels. Biomaterials 1988;9(6):494-499
  82. 82. Megeed Z, Cappello J, Ghandehari H. Thermal analysis of water in silk-elastinlike hydrogels by differential scanning calorimetry. Biomacromolecules. 2004;5(3):793-797
  83. 83. Serrano-Aroca Á, Campillo-Fernández AJ, Gómez-Ribelles JL, Monleón-Pradas M, Gallego-Ferrer G, Pissis P. Porous poly(2-hydroxyethyl acrylate) hydrogels prepared by radical polymerisation with methanol as diluent. Polymer (Guildf). 2004;45(26):8949-8955
  84. 84. Clayton AB, Chirila TV, Lou X. Hydrophilic sponges based on 2-Hydroxyethyl methacrylate. V. Effect of crosslinking agent reactivity on mechanical properties. Polymer International. 1997;44:201-207
  85. 85. Monleón-Pradas M, Gómez-Ribelles JL, Serrano-Aroca Á, Gallego-Ferrer G, Suay-Antón J, Pissis P. Interaction between water and polymer chains in poly(hydroxyethyl acrylate) hydrogels. Colloid & Polymer Science. 2001;279(4):323-330
  86. 86. Monleón-Pradas M, Gómez-Ribelles JL, Serrano-Aroca Á, Gallego Ferrer G, Suay Antón J, Pissis P. Porous poly (2-hydroxyethyl acrylate) hydrogels. Polymer (Guildf). 2001;42(10):4667-4674
  87. 87. Ahmed EM. Hydrogel: Preparation, characterization, and applications: A review. Journal of Advanced Research. 2015;6(2):105-121
  88. 88. Tanaka T. Gels. Scientific American. 1981;244(1):124-36, 138
  89. 89. Crank J. The Mathematics of Diffusion Second Edition. New York: Oxford University Press; 1975
  90. 90. Wang X, Zhang Y, Hao C, Dai X, Zhu F, Ge C. Ultrasonic synthesis and properties of a sodium lignosulfonate–grafted poly(acrylic acid-co-acryl amide) composite super absorbent polymer. New Journal of Chemistry. 2014;38(12):6057-6063
  91. 91. Gehrke SH, Biren D, Hopkins JJ. Evidence for Fickian water transport in initially glassy poly(2-hydroxyethyl methacrylate). Journal of Biomaterials Science. Polymer Edition. 1995;6(4):375-390
  92. 92. Wang J, Wu W. Swelling behaviors, tensile properties and thermodynamic studies of water sorption of 2-hydroxyethyl methacrylate/epoxy methacrylate copolymeric hydrogels. European Polymer Journal. 2005;41(5):1143-1151
  93. 93. Kavitha T, Kang IK, Park SY. Poly(acrylic acid)-grafted graphene oxide as an intracellular protein carrier. Langmuir. 2014;30(1):402-409
  94. 94. Krishna-Rao KSV, Ha CS. PH sensitive hydrogels based on acryl amides and their swelling and diffusion characteristics with drug delivery behavior. Polymer Bulletin. 2009;62(2):167-181
  95. 95. Zhang XZ, Yang YY, Wang FJ, Chung TS. Thermosensitive poly(N-isopropylacrylamide-co-acrylic acid) hydrogels with expanded network structures and improved oscillating swelling-deswelling properties. Langmuir. 2002;18(6):2013-2018
  96. 96. Shi L, Chen J, Teng L, Wang L, Zhu G, Liu S, et al. The antibacterial applications of Graphene and its derivatives. Small. 2016;12(31):4165-4184
  97. 97. Giano MC, Ibrahim Z, Medina SH, Sarhane KA, Christensen JM, Yamada Y, et al. Injectable bioadhesive hydrogels with innate antibacterial properties. Nature Communications. 2014;5(May):1-9
  98. 98. Zhao X, Wu H, Guo B, Dong R, Qiu Y, Ma PX. Antibacterial anti-oxidant electroactive injectable hydrogel as self-healing wound dressing with hemostasis and adhesiveness for cutaneous wound healing. Biomaterials. 2017;122:34-47
  99. 99. Jayakumar R, Prabaharan M, Nair SV, Tamura H. Novel chitin and chitosan nanofibers in biomedical applications. Biotechnology Advances. 2010;28:142-150
  100. 100. Jayakumar R, Prabaharan M, Sudheesh Kumar PT, Nair SV, Tamura H. Biomaterials based on chitin and chitosan in wound dressing applications. Biotechnology Advances. 2011;29:322-337
  101. 101. Ding F, Nie Z, Deng H, Xiao L, Du Y, Shi X. Antibacterial hydrogel coating by electrophoretic co-deposition of chitosan/alkynyl chitosan. Carbohydrate Polymers. 2013;98(2):1547-1552
  102. 102. Bauer AW, Kirby WMM, Sherris JC, Turck AM. Antibiotic susceptibility testing by a standardized single disk method. American Journal of Clinical Pathology. 1966;45:493-496
  103. 103. Zhou C, Li P, Qi X, Sharif ARM, Poon YF, Cao Y, et al. A photopolymerized antimicrobial hydrogel coating derived from epsilon-poly-l-lysine. Biomaterials. 2011;32(11):2704-2712
  104. 104. Li P, Poon YF, Li W, Zhu H-Y, Yeap SH, Cao Y, et al. A polycationic antimicrobial and biocompatible hydrogel with microbe membrane suctioning ability. Nature Materials. 2011;10(2):149-156
  105. 105. Andrews MA, Figuly GD, Chapman JS, Hunt TW, Glunt CD, Rivenbark JA, et al. Antimicrobial hydrogels formed by crosslinking polyallylamine with aldaric acid derivatives. Journal of Applied Polymer Science. 2011;119(6):3244-3252
  106. 106. Wang H, Zha G, Du H, Gao L, Li X, Shen Z, et al. Facile fabrication of ultrathin antibacterial hydrogel films via layer-by-layer “click” chemistry. Polymer Chemistry. 2014;5(22):6489-6494
  107. 107. Smith AW. Biofilms and antibiotic therapy: Is there a role for combating bacterial resistance by the use of novel drug delivery systems? Advanced Drug Delivery Reviews. 2005;57:1539-1550
  108. 108. Lakes AL, Peyyala R, Ebersole JL, Puleo DA, Hilt JZ, Dziubla TD. Synthesis and characterization of an antibacterial hydrogel containing covalently bound vancomycin. Biomacromolecules. 2014;15(8):3009-3018
  109. 109. He J, Söderling E, Lassila LVJ, Vallittu PK. Incorporation of an antibacterial and radiopaque monomer in to dental resin system. Dental Materials. 2012;28(8):e110-e117
  110. 110. Ji H, Sun H, Antibacterial QX. Applications of graphene-based nanomaterials: Recent achievements and challenges. Advanced Drug Delivery Reviews. 2016;105:176-189
  111. 111. Rai M, Yadav A, Gade A. Silver nanoparticles as a new generation of antimicrobials. Biotechnology Advances. Elsevier. 2009;27:76-83
  112. 112. Pitarresi G, Palumbo FS, Calascibetta F, Fiorica C, Di Stefano M, Giammona G. Medicated hydrogels of hyaluronic acid derivatives for use in orthopedic field. International Journal of Pharmaceutics. 2013;449(1-2):84-94
  113. 113. González-Sánchez MI, Perni S, Tommasi G, Morris NG, Hawkins K, López-Cabarcos E, et al. Silver nanoparticle based antibacterial methacrylate hydrogels potential for bone graft applications. Materials Science and Engineering: C. 2015;50:332-340
  114. 114. Fan Z, Liu B, Wang J, Zhang S, Lin Q, Gong P, et al. A novel wound dressing based on Ag/graphene polymer hydrogel: Effectively kill bacteria and accelerate wound healing. Advanced Functional Materials. 2014;24(25):3933-3943
  115. 115. Ping X, Wang M, Xuewu G. Surface modification of poly(ethylene terephthalate) (PET) film by gamma-ray induced grafting of poly(acrylic acid) and its application in antibacterial hybrid film. Radiation Physics and Chemistry. 2011;80(4):567-572
  116. 116. Bozukova D, Pagnoulle C, De Pauw-Gillet MC, Ruth N, Jérôme R, Jérôme C. Imparting antifouling properties of poly(2-hydroxyethyl methacrylate) hydrogels by grafting poly(oligoethylene glycol methyl ether acrylate). Langmuir. 2008;24(13):6649-6658
  117. 117. Kostina NY, Sharifi S, de los Santos Pereira A, Michálek J, Grijpma DW, Rodriguez-Emmenegger C. Novel antifouling self-healing poly(carboxybetaine methacrylamide-co-HEMA) nanocomposite hydrogels with superior mechanical properties. Journal of Materials Chemistry B 2013;1(41):5644
  118. 118. Cao B, Tang Q, Li L, Humble J, Wu H, Liu L, et al. Switchable antimicrobial and antifouling hydrogels with enhanced mechanical properties. Advanced Healthcare Materials. 2013;2(8):1096-1102
  119. 119. Wu D, Xu F, Sun B, Fu R, He H, Matyjaszewski K. Design and preparation of porous polymers. Chemical Reviews. American Chemical Society. 2012;112:3959-4015
  120. 120. Du N, Park HB, Robertson GP, Dal-Cin MM, Visser T, Scoles L, et al. Polymer nanosieve membranes for CO2-capture applications. Nature Materials. 2011;10(5):372-375
  121. 121. Shetty D, Jahovic I, Raya J, Ravaux F, Jouiad M, Olsen J-C, et al. An ultra-absorbent alkyne-rich porous covalent polycalix[4]arene for water purification. Journal of Materials Chemistry A. 2017;5:62
  122. 122. Kolanthai E, Dikeshwar Colon VS, Sindu PA, Chandra VS, Karthikeyan KR, Babu MS, et al. Effect of solvent; enhancing the wettability and engineering the porous structure of a calcium phosphate/agarose composite for drug delivery. RSC Advances. 2015;5(24):18301-18311
  123. 123. Pulko I, Wall J, Krajnc P, Cameron NR. Ultra-high surface area functional porous polymers by emulsion templating and hypercrosslinking: Efficient nucleophilic catalyst supports. Chemistry - A European Journal. 2010;16(8):2350-2354
  124. 124. Zhao C, Danish E, Cameron NR, Kataky R. Emulsion-templated porous materials (PolyHIPEs) for selective ion and molecular recognition and transport : Applications in electrochemical sensing. Journal of Materials Chemistry. 2007;17(23):2446
  125. 125. GD F, Yuan Z, Kang ET, Neoh KG, Lai DM, Huan ACH. Nanoporous ultra-low-dielectric-constant fluoropolymer films via selective UV decomposition of poly(pentafluorostyrene)-block-poly(methyl methacrylate) copolymers prepared using atom transfer radical polymerization. Advanced Functional Materials. 2005;15(2):315-322
  126. 126. Lv Y, Hughes TC, Hao X, Hart NK, Littler SW, Zhang X, et al. A novel route to prepare highly reactive and versatile chromatographic monoliths. Macromolecular Rapid Communications. 2010;31(20):1785-1790
  127. 127. Murphy AR, Laslett A, O’Brien CM, Cameron NR. Scaffolds for 3D in vitro culture of neural lineage cells. Acta Biomaterialia. 2017
  128. 128. Serrano-Aroca Á, Llorens-Gámez M. Dynamic mechanical analysis and water vapour sorption of highly porous poly(methyl methacrylate). Polymer (Guildf). 2017;125:58-65
  129. 129. Brígido-Diego R, Pérez-Olmedilla M, Serrano-Aroca ÁS, Gómez-Ribelles JL, Monleón-Pradas M, Gallego-Ferrer G, et al. Acrylic scaffolds with interconnected spherical pores and controlled hydrophilicity for tissue engineering. Journal of Materials Science. Materials in Medicine. 2005;40(18):4881-4887
  130. 130. Park K, Hyun JJ, Kim JJ, Ahn KD, Dong KH, Young MJ. Acrylic acid-grafted hydrophilic electrospun nanofibrous poly(L-lactic acid) scaffold. Macromolecular Research. 2006;14(5):552-558
  131. 131. Annabi N, Nichol JW, Zhong X, Ji C, Koshy S, Khademhosseini A, et al. Controlling the porosity and microarchitecture of hydrogels for tissue engineering. Tissue Engineering. Part B, Reviews. 2010;16(4):371-383
  132. 132. Arora KA, Lesser AJ, McCarthy TJ. Compressive behavior of microcellular polystyrene foams processed in supercritical carbon dioxide. Polymer Engineering and Science. 1998;38(12):2055-2062
  133. 133. Thomson RC, Wake MC, Yaszemski MJ, Mikos AG. Biodegradable polymer scaffolds to regenerate organs. Advances in Polymer Science. 1995;122:247-277
  134. 134. Andrianova GP, Pakhomov SI. Porous materials from crystallizable polyolefins produced by gel technology. Polymer Engineering and Science. 1997;37(8):1367-1380
  135. 135. Serrano-Aroca Á, Monleón-Pradas M, Gómez-Ribelles JL. Macroporous poly(methyl methacrylate) produced by phase separation during polymerisation in solution. Colloid & Polymer Science. 2007;285(7):753-760
  136. 136. Flynn L, Dalton PD, Shoichet MS. Fiber templating of poly(2-hydroxyethyl methacrylate) for neural tissue engineering. Biomaterials. 2003;24(23):4265-4272
  137. 137. Kang HW, Tabata Y, Ikada Y. Fabrication of porous gelatin scaffolds for tissue engineering. Biomaterials. 1999;20(14):1339-1344
  138. 138. Dragusin DM, Van Vlierberghe S, Dubruel P, Dierick M, Van Hoorebeke L, Declercq HA, et al. Novel gelatin–PHEMA porous scaffolds for tissue engineering applications. Soft Matter. 2012;8(37):9589
  139. 139. Agarwal S, Wendorff JH, Greiner A. Use of electrospinning technique for biomedical applications. Polymer (Guildf). 2008;49(26):5603-5621
  140. 140. Chia HN, Recent WBM. Advances in 3D printing of tissue engineering scaffolds. Journal of Biological Engineering. 2015;9(4):2-14
  141. 141. Derby B. Printing and prototyping of tissues and scaffolds. Science. 2012;338(6109):921-926
  142. 142. Horák D, Hlídková H, Hradil J, Lapčíková M, Šlouf M. Superporous poly(2-hydroxyethyl methacrylate) based scaffolds: Preparation and characterization. Polymer (Guildf). 2008;49(8):2046-2054
  143. 143. Kubinová Š, Horák D, Kozubenko N, Vaněček V, Proks V, Price J, et al. The use of superporous ac-CGGASIKVAVS-OH-modified PHEMA scaffolds to promote cell adhesion and the differentiation of human fetal neural precursors. Biomaterials. 2010;31(23):5966-5975
  144. 144. Barry JJA, Silva MMCG, Cartmell SH, Guldberg RE, Scotchford CA, Howdle SM. Porous methacrylate tissue engineering scaffolds: Using carbon dioxide to control porosity and interconnectivity. Journal of Materials Science. 2006;41(13):4197-4204
  145. 145. Nguyen LH, Gao M, Lin J, Wu W, Wang J, Chew SY. Three-dimensional aligned nanofibers-hydrogel scaffold for controlled non-viral drug/gene delivery to direct axon regeneration in spinal cord injury treatment. Scientific Reports. 2017;7(October 2016):42212
  146. 146. Kubinová Š, Horák D, Syková E. Cholesterol-modified superporous poly(2-hydroxyethyl methacrylate) scaffolds for tissue engineering. Biomaterials. 2009;30(27):4601-4609
  147. 147. Möller S, Weisser J, Bischoff S, Schnabelrauch M. Dextran and hyaluronan methacrylate based hydrogels as matrices for soft tissue reconstruction. Biomolecular Engineering. 2007;24(5):496-504

Written By

Ángel Serrano-Aroca

Submitted: 21 April 2017 Reviewed: 13 October 2017 Published: 01 August 2018