• No results found

Photo-crosslinked synthetic biodegradable polymer networks for biomedical applications

N/A
N/A
Protected

Academic year: 2021

Share "Photo-crosslinked synthetic biodegradable polymer networks for biomedical applications"

Copied!
31
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Full Terms & Conditions of access and use can be found at

https://www.tandfonline.com/action/journalInformation?journalCode=tbsp20

ISSN: 0920-5063 (Print) 1568-5624 (Online) Journal homepage: https://www.tandfonline.com/loi/tbsp20

Photo-crosslinked synthetic biodegradable

polymer networks for biomedical applications

Bas van Bochove & Dirk W. Grijpma

To cite this article: Bas van Bochove & Dirk W. Grijpma (2019) Photo-crosslinked synthetic biodegradable polymer networks for biomedical applications, Journal of Biomaterials Science, Polymer Edition, 30:2, 77-106, DOI: 10.1080/09205063.2018.1553105

To link to this article: https://doi.org/10.1080/09205063.2018.1553105

© 2019 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group.

Accepted author version posted online: 29 Nov 2018.

Published online: 12 Jan 2019. Submit your article to this journal

Article views: 1427

View related articles

View Crossmark data

(2)

REVIEW ARTICLE

Photo-crosslinked synthetic biodegradable polymer

networks for biomedical applications

Bas van Bochovea and Dirk W. Grijpmaa,b

a

Department of Biomaterials Science and Technology, Faculty of Science and Technology, Technical Medical Centre University of Twente, Enschede, The Netherlands;bDepartment of Biomedical Engineering, W. J. Kolff Institute, University Medical Centre, University of Groningen, Groningen, The Netherlands

ABSTRACT

Photo-crosslinked synthetic biodegradable polymer networks are highly interesting materials for utilization in biomedical applica-tions such as drug delivery, cell encapsulation and tissue engin-eering scaffolds. Varying the architecture, chemistry, degree of functionalization and molecular weight of the macromer precursor molecules results in networks with a wide range of physical- and mechanical properties, crosslinking densities, degradation charac-teristics and thus in potential applications. Photo-crosslinked net-works can easily be prepared and have the possibility to entrap a wide range of (biologically active) substances and cells. Additionally, spatial and temporal control over the crosslinking pro-cess when using additive manufacturing propro-cesses, allows for the preparation of network structures with complex shapes. Photo-crosslinked networks have been used to prepare drug delivery devi-ces, as these networks allow for drug delivery in a controlled way over a prolonged period of time. Furthermore, additive manufactur-ing techniques such as extrusion-based additive manufacturmanufactur-ing and stereolithography have been used to prepare photo-crosslinked tis-sue engineering scaffolds. This allows for the preparation of designed porous structures with precise control over the pore size and pore architecture and optimal mechanical properties. In par-ticular for stereolithography, a wide variety of resins based on bio-degradable photo-crosslinkable macromers has been developed.

ARTICLE HISTORY Received 20 September 2018 Accepted 25 November 2018 KEYWORDS Photo-crosslinking; bio-degradable polymer networks; biomedical applications; stereolithography

Photo-crosslinked synthetic biodegradable polymer networks

In polymer networks, the macromolecular chains are attached to each other by cova-lent bonds. In these materials viscous flow is not possible and creep is restricted. Especially in polymeric materials with low glass transition temperatures, such as

CONTACTDirk W. Grijpma d.w.grijpma@utwente.nl University of Twente, PO Box 217, 7500 AE Enschede, The Netherlands

ß 2019 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group.

This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives License (http://creativecommons.org/licenses/by-nc-nd/4.0/), which permits non-commercial re-use, dis-tribution, and reproduction in any medium, provided the original work is properly cited, and is not altered, trans-formed, or built upon in any way.

(3)

elastomers, it is important to prevent creep and so ensure form-stability of the mater-ial. Biodegradable elastomers have gained much attention in the biomedical field for application as flexible tissue engineering scaffolds and controlled drug delivery sys-tems [1]. Biodegradable materials are preferred, as these avoid a long-term foreign body response [2]. In tissue engineering, biodegradability allows for the growing tis-sue to slowly replace the scaffolding material and in drug delivery systems the drug release characteristics can be controlled by the degradation process [3]. Covalently crosslinked biodegradable elastomers have been prepared by reactions of end-func-tionalized polymers or oligomers. For example by free addition reactions or step poly-merization reactions [4]. In some cases, biodegradable polymers have been crosslinked by actinic radiation such as by gamma irradiation [5].

A very effective method to prepare such polymer networks is by photo-crosslink-ing oligomers that contain photo-polymerizable groups. Three specific photo-poly-merization reactions can be distinguished [6]:

1. [2þ 2] cyclo dimerization reactions using end-groups such as cinnamate-, cou-marin- or thymine end-groups[6–8],

2. radical recombination reactions leading to inter- and intramolecular crosslinking utilizing end-groups such as phenyl azide-, dithiocarbamate- and benzophenone end-groups, and

3. radical polymerization reactions using end-groups such as styryl-, fumarate- or (meth)acrylate end-groups (see Figure 1A) [6,9]. (Note that fumarate groups can be incorporated into the main chain [10].

Figure 1. (A) Photo-polymerizable (meth)acrylate- and fumarate end-groups. (B) Network formation by photo-crosslinking.

(4)

Radical photo-polymerizations have been used most often to polymerize the end-groups of these oligomers (macromers) and form a covalent network [9]. In the pro-cess of photo-crosslinking, a photo-initiator dissociates upon illumination and forms one or more radicals. These radicals can react with the double bonds of macromers, forming non-degradable carbon-carbon chains that act as multifunctional crosslin-kages [11]. This is schematically presented in Figure 1(B). (Although thermal-cross-linking would also be possible, the radical initiator is then formed upon heating, relatively high temperatures and reaction times are required [12]. Photo-crosslinking is relatively rapid and efficient and can be done at low temperatures, making it thus more advantageous compared to thermal-crosslinking).

Oligomers with fumarate (end-)groups are interesting materials for preparing bio-degradable networks. Fumaric acid is found in the human body and therefore it is expected that residual fumarate groups will be biocompatible and non-toxic [9,13]. However, compared to (meth)acrylate functionalized oligomers, the reactivity of fumarate-functionalized oligomers is relatively low and therefore the use of reactive diluents is required [14–16]. The use of such reactive diluents will lead to an increase in the non-degradable part of the networks. For applications in medicine where bio-degradability of the implant is desired the non-degradable content of the implant should be as low as possible [11]. Thus, for such applications (meth)acrylate function-alized oligomers are preferred.

Polymers and oligomers used in the preparation of photo-crosslinked synthetic biodegradable networks

There are many biodegradable polymers and oligomers that have been used to pre-pare photo-crosslinkable macromers for biodegradable polymer networks. Examples include poly(D,L-lactide) (PDLLA) [17], poly(e-caprolactone) (PCL) [18,19], poly(tri-methylene carbonate) (PTMC) [20], poly(ethylene carbonate) (PEC) [21], and block copolymers containing poly(ethylene glycol) (PEG), poly(propylene glycol) (PPG) or poly(tetramethylene glycol) (PTMG) and poly(glycolide) (PGA), PDLLA or PCL seg-ments [22,23]. Most of these polymers can readily synthesized by the ring opening polymerization of their cyclic monomers. The polymerization is usually initiated by an alcohol and catalyzed by stannous octoate [24] This reaction is usually performed in the melt at temperatures between 90 and 180C [1]. By adjusting the amount and the functionality of the hydroxyl-group terminated alcohols used as initiator, the molecular weight and architecture of the synthesized oligomers can be precisely con-trolled [25].

An overview the networks described below and their properties is given in Table 1.

Polyesters

Polyesters are a group of polymers that contain an ester group in their main chain. Biodegradable polyesters used to prepare synthetic biodegradable networks include PDLLA and PCL.

(5)

High molecular weight (HMW) non-crosslinked PDLLA has a glass transition tem-perature (Tg) of approximately 55C and an elasticity modulus close to 3 GPa. Photo-crosslinked networks prepared form methacrylate-functionalized PDLLA were described by Melchels et al. [17]. In this study, the effect of the molecular architecture of the macromers on the thermal and mechanical properties of the networks was investigated. For this, multifunctional alcohols were used to initiate the ring opening polymerization and obtain branched macromers with 2, 3 or 6 arms of different lengths. It was shown that the Tg of the networks increased with decreasing arm length. Networks prepared from macromers with the highest molecular weights and arm lengths had a Tgsimilar to that of HMW PDLLA. Networks prepared from mac-romers with arm lengths of only 0.6 kg/mol had a much higher Tg of approximately 76C. As a result of lower crosslink densities, the degree of swelling in good solvents was found to increase with increasing arm length. A significant effect on the

Table 1. Overview of photo-crosslinked, biodegradable networks, their properties and biomedical applications. Polymeric Precursor Photo-crosslinked Polymer Network Tg(C) Tensile E

modulus (MPa) Applications References

Polyesters PDLLA 55–76a 2.6 103 Orthopedic and

bone tissue engineering

[17,26,27,28]

PCL 62– 46a 6.7–15.4b,c

498–825d Tissue engineering,drug delivery [29,30,18,31]

PGS 30 0.048–1.375 Cell encapsulation,

tissue adhesives [32–35]

CABEse – 0.04–276 Soft tissues,

ortho-pedic tissue engineering, drug delivery, tissue adhesives

[36]

Polycarbonates PEC 15–20 0.5– 7b Drug delivery,

tis-sue engineering [21]

PTMC 16–7.6a 4–315b Cartilage and bone

tissue engineering

[37,38,39]

PEG-based materials

PEG-MA 55 – Cell encapsulation [30,40]

Combinatorial net-work 1f 68, 47, 24, 33h 1.49 Tissue engineering [30,41] Combinatorial net-work 2g 70, 22, 35h 1.24 Tissue engineering [30,41]

Semi synthetic pol-ymers based on natural polymers

GelMA – 2–30i Neural-, cartilage-,

bone-, muscle-and organ engineering

[42–44]

Chitosan – – Tissue engineering,

drug delivery, tissue adhesives

[45–48]

a

Depends on the molecular weight of the macromer used. The Tgincreased with decreasing molecular weight. bDepends on the molecular weight of the macromer used. The modulus decreased with increasing

molecu-lar weight.

cCrosslinked in the bulk. d

Crosslinked in a non-reactive diluent.

eProperties of CABEs are highly dependent on the polymers composition. For a more detailed overview, see Tran

et al. [36].

fCombination of PTMC 4 kg/mol, PDLLA 4 kg/mol, PCL 4 kg/mol, PEG 4 kg/mol and PEG 10 kg/mol macromers. g

Combination of PTMC 4 kg/mol, PDLLA 4 kg/mol, PEG 4 kg/mol, PTMC 10 kg/mol and PEG 10 kg/mol macromers.

hT

g’s are the Tg’s of the individual components. i

(6)

mechanical properties of the networks was not observed, all networks having proper-ties similar to those of HMW PDLLA.

Poly(e-caprolactone) is a semi-crystalline, highly biocompatible polymer with a low Tgof approximately 60C, a melting point close to 65C and an elasticity modulus of approximately 260 MPa [19,25,49]. The thermal properties of photo-crosslinked networks prepared from methacrylated PCL have been described by Elomaa et al. and Zant et al. [29,30]. Interestingly, the PCL networks were found to be amorphous. For networks prepared from macromers with a low molecular weight (below 4 kg/mol) the Tgis 10–15C higher than that of their respective macromers. For networks pre-pared from macromers with molecular weights of 4 kg/mol and higher, Tg is similar to that of linear PCL. Elomaa et al. further evaluated the swelling ratios and mechan-ical properties of the prepared PCL networks [29]. As can be expected, the swelling ratio of PCL networks in good solvents increases with increasing molecular weight of the macromer used to prepare the networks. The networks behaved in a rubber-like manner and showed elastic deformation. With increasing molecular weights, the elas-tic modulus of the networks decreased while their elongation at break increased.

Amsden and coworkers synthesized a series of poly(e-caprolactone-co-D,L-lactide) macromers [50–53] and prepared the corresponding networks by photo-crosslinking. The glass transition temperature of networks prepared from poly(e-caprolactone-co-D,L-lactide) macromers with a 50:50 molar ratio composition were close to 3C and independent of the molecular weight of the macromers [50].

Poly(glycerol sebacate) (PGS) is a semicrystalline polymer with low Tg between 52 and 18C [54]. At 37C it is completely amorphous [54,55]. Photo-crosslinked networks of PGS have been prepared by crosslinking acrylated PGS [32–35]. The mechanical properties of these networks a linearly depend on the degree of acryla-tion [32]. With increasing degree of acrylation the modulus increases, as does the ultimate tensile strength. The moduli varied from 0.048 MPa to 1.375 MPa with the highest ultimate tensile strength of 0.498 MPa. Increasing the macromer molecular weight results in increasing strains at break [34]. Potential applications include cell encapsulations devices [33] and tissue adhesives [35].

Citrate based biodegradable elastomers (CABEs) are polyesters prepared by the reaction of a diol with citric acid [36]. Polymers with a wide range of properties can be obtained by varying the diol length, chemical composition of the diol, and partial replacement of the citric acid with other diacids. Acrylate- and fumerate functional-ized CABEs have been developed to allow for photo-crosslinking, which resulted in strengthened networks and allowed fine tuning of the mechanical and degradation properties of the materials [56–58]. Proposed applications of these type of networks include coatings, films, and devices for tissue engineering and drug delivery.

Polycarbonates

Polycarbonates are a group of polymers that contain a carbonate ester group in their main chain. Polycarbonates that have been used to prepare synthetic biodegradable networks are PEC and PTMC.

Poly(ethylene carbonate) networks have been prepared by the thermal degradation of high molecular weight PEC, subsequent acrylation an photo-crosslinking [21].

(7)

These networks had a Tgof approximately 20C meaning the networks were rubbery at room temperature. The elastic modulus of the networks decreased with increasing macromer molecular weight. Potential applications include drug delivery and tissue engineering scaffolds.

Poly(trimethylene carbonate) is an amorphous polymer with a low Tg of approxi-mately 16C [37,59]. The mechanical properties of PTMC are strongly dependent on its molecular weight [5]. Non-crosslinked, low molecular weight (LMW) PTMC is soft and gummy, and has very low modulus and tensile strength. As a result, TMC was mainly used as comonomer to reduce the modulus of lactide and glycolide poly-mers. Non-crosslinked, high molecular weight (HMW) PTMC is tough, flexible and to some extent shows rubber-like recovery after mechanical deformation. By prepar-ing networks from methacrylate-functionalized PTMC oligomers, creep resistant net-works with excellent mechanical properties could be obtained [37].

Figure 2 shows stress-strain curves of PTMC networks prepared from macromers (methacrylate end-functionalized) of different molecular weights. Networks prepared from macromers with molecular weights lower than 1800 g/mol were rigid and brittle. In contrast, networks prepared from macromers with molecular weights higher than 10 kg/mol were rubber-like with elastic moduli of approximately 5 MPa. The max-imum tensile strengths and elongations at break of the networks increased with increasing molecular weight of the macromers used. As was the case for PCL, the swelling ratios of the networks in a good solvent increased with increasing molecular weights. Interestingly, the Tg values of networks prepared from very low molecular weight macromers were relatively high (the Tg of networks prepared from a macro-mer with Mn of 1 kg/mol was 7.6C). With an increase in the molecular weight of

Figure 2. Stress-strain curves of PTMC networks prepared by photo-crosslinking PTMC macromers (methacrylate end-functionalized) of different molecular weights. In the figure, the molecular weights of the macromers used to prepare the networks are shown with the corresponding stress-strain curves. Reprinted with permission from [37].

(8)

the macromers, the Tg of the corresponding networks approached the Tg value of HMW PTMC.

PTMC networks have been investigated for a variety of medical applications, which include cartilage tissue engineering [60], annulus fibrosus tissue engineering [61,62], meniscus tissue engineering [38], preparation of microvascular networks [63] and orbital floor implants [39].

To allow tuning of the mechanical- and degradation properties, copolymer net-works of TMC with DLLA and/or e-CL have been extensively investigated [7,9,50,64]. Copolymerizing TMC and DLLA, subsequent functionalization with methacrylate end-groups to yield poly(trimethylene carbonate-co-D,L-lactide) macromers, and photo-polymerization allows the formation of copolymer networks in which the glass transition temperature depends on the ratio of the co-monomers [9,16]. In this way networks with a wide range of mechanical properties can be obtained. For example, Sharifi et al. used such networks to prepare structures with shape memory behavior: the temporary shape of the structure is fixed at temperatures below Tg of the copolymer, it then returns to its original permanent shape upon heating to body temperature. Surgically implantable devices prepared from these photo-crosslinked poly(trimethylene carbonate-co-D,L-lactide) macromers, can be used in minimal invasive surgery [65,66]. An example of such an implant is shown inFigure 3.

PTMC degrades without the formation of acidic degradation products [51,67]. Therefore, preparing biodegradable networks from functionalized TMC and DLLA copolymers instead of from DLLA homo-polymers may be beneficial in applications such as drug delivery or bone tissue engineering [68].

Copolymerizing TMC and e-CL to obtain poly(trimethylene carbonate-co-e-caprolactone) macromers results in networks with low glass transition tempera-tures ranging from 23 to 50C, depending on the e-CL content [9,51,69]. These networks are rubbery and amorphous at room temperature, with relatively low elastic moduli [51,69]. Copolymer networks of poly(TMC-co-e-caprolactone-co-D,L-lactide) macromers were prepared as well [64]. These networks were investigated for drug release purposes and were able to show sustained release for more than 10 days.

Figure 3. Shape recovery of a 3D structure prepared from photo-crosslinked P(DLLA-co-TMC) mac-romers. (A) Temporary shape of the structure at 0C. (B) Transient shape of the structure during heating at 37C. (C) Completely recovered structure at 37C. Reprinted with permission from [66].

(9)

PEG-based materials

In contrast to natural hydrogels, synthetic hydrogels can readily be prepared, proc-essed and tailored. Hydrogels are interesting for biomedical applications, as they pro-vide highly swollen 3D environments which are similar to soft tissues [70]. Furthermore, they allow easy diffusion of nutrients, waste products and drugs.

One of the major hydrophilic synthetic polymers used to prepare hydrogels for biomedical applications is poly(ethylene glycol) (PEG) [70]. PEG is a biocompatible, non-toxic and water-soluble polymer [23]. As a result, PEG is used in wide range of biomedical applications including drug delivery, tissue engineering and implant sur-face modification.

The end groups of PEG oligomers and polymers are hydroxyl groups. These can be reacted to yield other functional end groups, which include (meth)acrylates that allow photo-crosslinking into hydrogel networks. Photo-crosslinking is the most com-mon method to make PEG hydrogel networks [70]. Hydrogels prepared from PEG-methacrylate have a low Tg of approximately 55C, and high water uptake of 900–1700 wt% [30]. Networks prepared from (meth)acrylated PEG are not readily degradable in vivo [71], but below a molecular weight of approximately 30 kg/mol it can be excreted from the body via the renal pathway [72]. However, by using PEG as initiator in the ring opening polymerization reaction of polyesters or TMC, biodegrad-able hydrogel networks containing high amounts of PEG can be prepared [23,73]. By carefully adjusting the molecular weight of the PEG and the composition of the TMC/ ester (co)polymer segment, the degradation behavior and drug release profile of such networks can be controlled.

Basic PEG hydrogels have resistance to protein adsorption [74], and in general a lack of cell specific adhesion [70]. For tissue engineering purposes, the lack of cell adhesion is a major limitation. Modification of PEG hydrogels with cell adhesive pep-tides (CAPs) derived from the extracellular matrix have been researched extensively to overcome these limitations. Photo-polymerization of PEG diacrylates and acrylated peptides, peptide monoacrylates [75] and peptide diacrylates [76] resulted in networks with significantly higher cell adhesion and spreading compared to basic PEG hydrogels.

In a combinatorial approach, Zant et al. used mixtures of homo-polymeric macro-mers based on methacrylated DLLA, TMC, CL and PEG to prepare 255 different photo-crosslinked networks in solution [30]. After extraction and drying, these mixed-macromer networks were evaluated with regard to their physical and biological characteristics in a high throughput manner. Two macromer combinations consisting of i) PTMC 4 kg/mol, PDLLA 4 kg/mol, PCL 4 kg/mol, PEG 4 kg/mol and PEG 10 kg/ mol and ii) PTMC 4 kg/mol, PDLLA 4 kg/mol, PEG 4 kg/mol, PTMC 10 kg/mol and PEG 10 kg/mol, showed interesting properties. These networks had high water uptake (approximately 190%), showed excellent cell adhesion, and at the same time possessed excellent mechanical properties. The elastic moduli where up to 1.49 MPa and the networks were very resistant to tearing. Porous structures prepared from these macro-mer combinations could be compressed up to 80% without failure. It was hypothe-sized that the excellent properties of these networks were due to phase separation of the different macromers [41]. Phase separation was shown by the presence of glass

(10)

transition temperatures that corresponded to the individual macromer components in DSC, as well as by AFM and XRD.

Semi-synthetic polymers

Semi synthetic polymers are chemically modified natural polymers. Some examples include modified gelatin and chitosan.

Gelatin is a substance consisting of denatured and partially hydrolyzed natural col-lagen [77]. Gelatin is an attractive material for tissue engineering and drug delivery as it is biocompatible, biodegradable, can easily be manipulated and can be used at low cost [78,79]. However, gelatin is instable at body temperature and therefor it needs to be covalently crosslinked for use as biomaterial [80]. Gelatin modified with photo-crosslinkable side groups, gelatin-methacryloyl (gelMA), allows for crosslinking with a high degree of control of the properties of the obtained hydrogel network [78]. The crosslinking of gelMA leads to hydrogel that is stiffer as compared to its non-cross-linked counterpart [81]. The mechanical properties of photo-crosslinked gelMA were shown to be directly proportional to the degree of methacryloyl substitution (DS) and the gelMA mass/volume ratio [42–44]. When the DS increased from approximately 50 to 73%, the compression modulus of gelMA hydrogels increased from 2.0 to 4.5 kPa [43]. Likewise, for gelMA with a DS of 54% the compression moduli of work with a w/v ratio of 5, 10 and 15% were 2, 10 and 22 kPa respectively while net-works prepared from gelMA with a DS of 81% and the same w/v ratios had moduli of 3, 16 and 30 kPa [44]. GelMA hydrogels have been shown to be highly adaptable and have been researched for a wide range of biomedical applications including neural, cartilage, bone, muscle and organ engineering [42].

Polysaccharide-based hydrogels are materials with interesting properties for bioma-terials as they are non-toxic, low cost in use, biocompatible and biodegradable [82]. One of these polysaccharides, chitosan, is a well-researched polysaccharide obtained from the alkaline hydrolysis of chitin [83,84]. On the chitosan chain, many amine and hydroxyl groups are available for reaction with photo-crosslinkable functional groups to prepare photo-crosslinked hydrogel networks [85,86]. These networks have been used in tissue engineering and drug delivery systems [45]. Methacrylated glycol chitosan could be crosslinked with blue light initiators [46]. These materials had bet-ter mechanical properties with increasing crosslinking time and the stability and deg-radation were depending on the mechanical properties. An injectable, photo-crosslinkable, chitosan based hydrogel was developed based on chitosan, PEGDA and N,N-dimethylacrylamide (DMMA) [47]. This material had excellent mechanical prop-erties, was thermally stable, showed no cytotoxicity and was able to promote cell adhesion and proliferation. A photo-crosslinkable chitosan adhesive showed superior strength compared to fibrin glue, potentially due to covalent binding with tissue pro-teins [48].

Degradation and erosion of synthetic biodegradable networks

To successfully apply the previously described networks in the biomedical field, it is essential to understand the degradation and erosion behavior of the networks.

(11)

Degradation is defined as the process in which polymer chains are cleaved, while ero-sion is defined as the loss of material mass as a result of dissolution and diffuero-sion of the soluble low molecular weight compounds that are formed upon degradation [87].

Degradation can occur by a variety of mechanisms, including hydrolysis, thermoly-sis and mechanical or oxidative stress [12]. Hydrolyzable bonds such as ester-, anhyd-ride-, amide- and carbonate bonds can be found in the main chains of many synthetic biodegradable polymers. These bonds can be cleaved upon reaction with water, either enzymatically or non-enzymatically. Factors that influence the rate of degradation are glass transition temperature, hydrophilicity, crosslinking density, pH, presence of proteins, nature of the labile bond and accessibility of the bonds to water or enzymes.

Biodegradable polymers and polymer networks can be categorized as surface- or bulk eroding materials [12]. Erosion is a complex process that depends on polymer degradation, polymer molecular weight, swelling, and diffusion of water, monomers and oligomers [88]. Surface eroding polymers lose material from the surface only [12]. Therefore, the rate of the loss of mass and the change in dimensions of the polymeric device depend on its surface area. As the molecular weight of the remain-ing polymer remains essentially the same, the strength of the material essentially remains unchanged. This is shown inFigure 4(A). In bulk degradation, the mass and the dimensions of the material remain unchanged for relatively long times. However, the molecular weight of the material decreases significantly [12]. Upon reaching a critical low molecular weight the material loses it mechanical strength, potentially with dramatic mechanical failure of the implant as a result. Rapid release of degrad-ation products then also occurs. This is shown inFigure 4(B).

Although most biodegradable polymers and polymer networks degrade by bulk erosion, surface eroding materials are to be preferred [12]: in medical implants and

Figure 4. Schematic illustration of the processes of surface erosion (A) and bulk erosion (B). The effect of degradation on strength, molecular weight and mass of the remaining material is shown. Reprinted with permission from [89].

(12)

tissue engineering scaffolds the mechanical properties and structural integrity of the implants are maintained during the functional life time of the implant. For drug delivery devices, the predictability of the surface erosion process allows for well-trolled release. The sequential release of the bioactive component takes place at a con-stant release rate (zero-order release).

The degradation and erosion behavior of photo-crosslinked networks has been studied extensively [26,52,64,90]. PDLLA networks degraded hydrolytically in approximately 40 weeks via bulk erosion [26]. The networks were form-stable and showed very little mass loss in the first 6 months. The mechanical properties remained unchanged for approximately 15 weeks, then the materials failed catastroph-ically with near complete mass loss in a very short time.

Interestingly, the degradation mechanism of copolymeric poly( e-caprolactone-co-D,L-lactide) networks appeared to depend on the crosslink density [52]. Networks prepared from end-functionalized macromers with low molecular weights (i.e. high crosslink density) degraded via surface erosion, while networks prepared form higher molecular weight macromers degraded via bulk erosion.

While the degradation of many polyester networks such as PDLLA and PCL have been extensively researched, the degradation products of the networks have not been extensively analyzed. However, polyester degradation is in general characterized by the formation of acidic compounds [91,92]. These acidic compounds may be harmful in applications such as bone tissue engineering [39]. Analyses of the degradation products of these PDLLA networks showed that the ester bonds with the poly(lactide) chains were much more prone to hydrolyses than the ester bonds between the lactide and the methacrylates [26]. As a result, the degradation products are lactic acid, low molecular weight oligomers and poly(lactic acid methacrylate) chains. The average length of these poly(lactic acid methacrylate) chains was between 1.1 and 3.5 kg/mol, falling within the range of effective and fast renal clearance.

PGS degrades via surface erosion [54,93], primarily due to cleavage of the ester linkages [55]. PGS networks degrade at a slower rate, indicating that the alkyl cross-links formed by the acrylate groups are less susceptible to degradation [32]. No evi-dence of inflammation or necrosis was observed upon PGS implantation and degradation in vivo [34]. CABEs degrade through hydrolysis of the ester linkages [36,56] where shorter, more hydrophilic diols used in the polymerization resulted in faster degradation [36,57].

PEC networks degrade in vivo via surface erosion [21]. Macrophages are heavily involved in the degradation. It was shown that increased crosslinking densities decreased the degradation rates.

PTMC networks degrade by enzymatic surface erosion. The degradation rate of networks prepared from PTMC macromers was found to depend on the molecular weight of the macromers used to prepare these networks [90]. Other studies showed that networks prepared by photo-crosslinking linear HMW PTMC in mixtures with low molecular weight PTMC macromers as a cross-linker also degrade via surface erosion [94,95]. In vivo, the surface erosion of PTMC may be mediated by macro-phages. It was shown that after culturing macrophages on PTMC network films, pits had formed on the surface and loss of mass was observed [95]. Degradation

(13)

experiments on linear high molecular weight PTMC showed that PTMC degraded without producing acidic degradation products [67].

An oxidative degradation mechanism was proposed for PTMC [51] and PEC [21]. This degradation proceeds by the formation of ionic end groups via a nucleophilic attack of superoxide ions. This is then followed by a chain unzipping reaction of which TMC and EC monomers are among the degradation products. In vivo, the degradation products of PTMC may also include 1,3-propanediol and CO2[51].

Poly(TMC-co-e-caprolactone-co-D,L-lactide) macromers were used to prepare net-works that had higher degradation rates than poly(trimethylene carbonate-co-D,L-lac-tide) networks, but released minimal amounts of acidic degradation products [64].

Gelatin is prepared from collagen extracted from bovine- or porcine skin under acidic or basic conditions [78]. The bioactive sequences for cell attachment and met-alloproteinase (MMP) sensitive degradation sites in the collagen remain present in the gelatin backbone. In vitro, enzymatic degradation of GelMA networks is observed [42,96,97]. The degradation products are particularly oligomethacrylates [78,98]. GelMA hydrogels prepared from gelatin obtained under basic conditions showed good biocompatibility. In contrast, GelMA hydrogels prepared from gelatin prepared under acidic condition showed inflammatory reactions, possibly due to high concentrations of endotoxins in this type of gelatin. Polysaccharides such as chitosan degrade into non-toxic oligosaccharides which can be excreted or incorporated into glycosaminoglycans and glycoproteins [82].

Biomedical applications of synthetic biodegradable networks

Photo-crosslinked biodegradable networks form an interesting group of materials for biomedical applications [99]. This interest relates to:

1. the ease of preparation (also in vivo),

2. the possibility to entrap a wide range of substances and even cells in the net-works [100], and

3. the spatial and temporal control over the polymerization process which allows for the preparation of network structures with complex shapes [101].

As a result, photo-crosslinked biodegradable networks have been studied for a var-iety of applications such as drug delivery [102] and tissue engineering [3].

Drug delivery devices

Controlled and sustained delivery greatly improves the therapeutic efficacy and safety of drugs. [103]. Ideally, implantable drug delivery devices are biodegradable as they will not need to be removed after the drug has been delivered [104].

Photo-crosslinked biodegradable polymer networks are an interesting group of materials for application in drug delivery devices [105]. Through photo-crosslinking, drugs can easily be entrapped in the networks by dissolving or dispersing the drugs into the macromer solution prior to crosslinking [73,106]. This allows for large

(14)

amounts of drugs to be loaded into the devices at high efficiencies. As photo-cross-linking is fast and can be performed with minimal heat generation, heat-sensitive compounds such as proteins can be incorporated as well. Detrimental reactions of proteins with free radicals [107] are avoided, as in the photo-crosslinking the macro-mers act as free radical scavengers [108].

Photo-crosslinked biodegradable polymer networks allow control over the rate of release of the incorporated compounds by variation of the crosslink density and com-position of the networks [64,68,73,105,106,109]. Different studies showed that less densely crosslinked networks released incorporated compounds faster than more densely crosslinked networks [106]. Furthermore, several studies showed that more hydrophilic networks lead to more rapid release [105,106,109]. In block-copolymeric hydrogel networks, variation of the hydrophilicity of the networks by varying the chemical composition if the hydrophobic segment allowed good control of drug release profiles [73].

In addition, poly(ester anhydride) networks prepared from PCL show great versa-tility for drug delivery [18,110–113]. Hydroxyl group-terminated PCL oligomers are acid-functionalized with an anhydride compound to yield acid-terminated oligomers. Subsequent methacrylate-functionalization results in photo-crosslinkable methacrylate functionalized oligomers (macromers). Regular PCL networks degrade very slowly. The addition of the labile anhydride bond results in much faster degradation. PCL-anhydride networks prepared from low molecular weight precursors degrade in 48 hours [110]. The addition of the alkenyl chain in the anhydride bond result in slower degradation of 64 to 72 hours for alkenyl chains containing 12 or 18 carbons respectively [111]. The degradation can be further slowed down by increasing the molecular weight of the PCL precursors [113]. The control over the degradation time of these types of polymer networks results in control over the drug release time. These networks have been investigated for the controlled release of drugs and peptides.

For drug delivery devices, it is important that the formation of large amounts of acidic degradation products is avoided [64,68]. In PLGA release systems, is was shown that the degradation of such systems resulted in a drop of the pH within the device due to acidic monomeric and oligomeric degradation products [114,115]. This can result in the denaturation of acid labile proteins such as VEGF in about 7–10 days [68]. Cleland et al. reported that VEGF released from PLGA lost approxi-mately 13% of its heparin binding affinity in 8 days [116] and Kim et al. showed that released VEGF lost 25% if its bioactivity [117].

Tissue engineering scaffolds

In tissue engineering, biodegradable scaffolds are used in combination with cells and/ or biologically active compounds to induce the (re)generation of tissuesin vitro or in vivo [118]. Scaffolds are porous implants intended to provide temporary support for cells and the formed tissues. Such scaffolds ideally have a high porosity, good pore interconnectivity and optimal pore sizes for an intended application [119–121]. The scaffolds need to be biocompatible, biodegradable at a rate which matches the tissue

(15)

replacement, and have mechanical properties that are compatible with those of the tissues that are to be regenerated [119,122,123].

Conventional techniques used to fabricate tissue engineering scaffolds include solv-ent casting, particulate porogen leaching, phase separation, membrane lamination, melt molding, injection molding and freeze drying [119,120,124]. Several of these techniques have also been used to prepare photo-crosslinked porous struc-tures [125–127]. For example, porous tubular scaffolds for vascular tissue engineering have been prepared by photo-crosslinking a mixture of photo-crosslinkable PTMC macromers and salt particles, followed by leaching of the salt [125]. Porous photo-crosslinked scaffolds have also been prepared by employing temperature-induced phase separation [126,127]. Upon cooling macromer solutions ethylene carbonate (a crystallizable solvent), subsequent photo-crosslinking of the matrix and extraction of the dispersed ethylene carbonate crystals with water, a porous photo-crosslinked structure is obtained.

Scaffolds fabricated by these conventional techniques often result in inhomogen-eous structures with irregular pore sizes and wide pore size distributions, poor pore connectivity and inferior mechanical properties [27,29]. Additive manufacturing tech-niques, on the other hand, allow for the preparation of designed porous structures with precise control over pore size and pore architecture, and optimal mechanical properties [38,128]. Furthermore, additive manufacturing allows the preparation of complex structures, shapes and patient-specific tissue engineering scaffolds [101,129]. In Figure 5 a comparison is made between a designed porous structure prepared by SLA (an additive manufacturing method) and a scaffolding structure prepared by salt-leaching.

Figure 5. Overview a scaffold with a complex porous architecture prepared by stereolithography and a scaffold prepared by salt-leaching. From the photos and themCT visualization it is clear that preparing tissue engineering scaffolds by a 3D printing technique such as stereolithography results in scaffolds with much higher control over pore architecture. Reprinted with permission from [128].

(16)

Cell encapsulation devices

A most interesting use of photo-crosslinked networks is in the preparation of cell encapsulation devices. In such devices, cells are encapsulated in a support structure during its formation rather than seeded onto prefabricated tissue engineering scaf-folds [130]. Although the number of photo-crosslinkable biomaterials suited for cell encapsulation is limited due to the required cytocompatibility of the encapsulation process, use of cell encapsulation devices can be highly advantageous. First, injectable systems with cells suspended in liquid precursor solutions can be used, and second, by curing the materialin situ, enhanced adhesion of the implant to the tissues can be achieved without the use of glues or sutures.

Hydrogels are attractive materials for this application as they provide a highly hydrated tissue-like environment for cells and tissues. In addition, they are easy to handle and can be formed in situ. Several studies aiming at engineering cartilage tis-sue have made use of PEG-based hydrogels to encapsulate the cells [100,131]. Uniform cell seeding was easy to achieve and chondrocyte cell viability could readily be maintained in these hydrogels [132]. It has been shown that the mechanical prop-erties of the hydrogels and the incorporation tissue-specific molecules can have an effect on ECM production [100], chondrocyte metabolism and gene expression [133].

The illumination needed to initiate the photo-crosslinking process can be done with ultraviolet (UV) and visible light (VIS). While the use of UV light is not a prob-lem as long as no biological content is involved, when cells and/or proteins are incor-porated into the material prior to photo-crosslinking caution with UV light is necessary [134]. The free radicals formed upon irradiation not only react into the net-work, but can also react with proteins, cell membranes and DNA. Reactive oxygen species that indirectly cause cellular damage can be formed as a result of this inter-action of the radicals with the cells. Therefore, careful selection of the wavelength of the light and a photo-initiator that reacts to that wavelength is essential.

Other applications

Other biomedical applications of photo-crosslinked networks include tissue adhesives, tissue barriers and dental composites.

Photo-crosslinkable tissue adhesives have been developed from natural materials such as chitosan and mussel proteins, and from synthetic methacrylate-functionalized block copolymers containing PEG and DLLA or TMC segments [135–137]. Upon irradiation with light, such synthetic adhesives not only crosslink but at the same time also adhere to the tissue [136] as the (meth)acrylate groups can covalently bind to amine groups present in the tissue [138].

Photo-crosslinkable hydrogels have been investigated for use as resorbable tissue barriers to prevent postoperative adhesions [139,140]. These systems were based on PEG and lactide block copolymers that are end-functionalized with methacrylic acid. In situ photo-crosslinking allows the formation of the barriers that prevent adhe-sions [139]. These barriers could also be loaded with drugs [140].

Low molecular weight, multifunctional (meth)acrylates have been used in the prep-aration of photo-crosslinkable resins for dental applications [141–143]. While metal

(17)

alloys were the standard, drawback such as mercury content and its associated tox-icity, high costs and poor aesthetics prompted the need for other systems [142]. These materials were mixtures of low molecular weight acrylates and methacrylates and could be crosslinked by UV- and visible light. The exact composition of the mix-tures is an important factor in determining the crosslinking density of the networks [141,142]. High degrees of crosslinking need to be achieved to prevent shrinkage of the network and leaching of unreacted (meth)acrylates.

Additive manufacturing

Additive manufacturing is a very interesting method to prepare photo-crosslinked networks for most of the aforementioned application as it allows for the preparation of highly complex, designed 3D structures with optimized properties and patient-spe-cific shapes [38]. Photo-crosslinking has been employed in several additive manufac-turing techniques. While the most widely used additive manufacmanufac-turing technique to prepare photo-crosslinked structures and tissue engineering scaffolds is stereolithog-raphy (SLA) [17,144], extrusion-based additive manufacturing [91,94,145] and a com-bination of these methods [146] have also been done.

Extrusion-based additive manufacturing

Extrusion-based additive manufacturing methods are interesting for the preparation of designed structures. These methods are based on the extrusion of a material at pre-defined locations in a layer-by-layer manner to form 3D structures with specific internal and external geometries [91]. A commonly used extrusion-based additive manufacturing technique is fused deposition modeling (FDM) [147,148]. Aliphatic polyesters such as PLLA and PCL are very well suited for FDM, as they flow in the melt at elevated temperatures and readily solidify after extrusion. Polymers that do not crystallize or only slowly solidify are more difficult to process as they will not be form-stable [37,59,91]. An example of such a polymer is PTMC, which is amorphous and has a low glass transition temperature. Nevertheless, this polymer could be proc-essed by an extrusion-based additive manufacturing method when the polymer was dissolved in a crystallizable solvent. Using low-temperature extrusion-based additive manufacturing (LTEAM) [91,94] the materials was extruded at a temperature above the melting point of the solvent. By cooling the material after extrusion, a scaffold that was form stable until photo-crosslinking could be performed was obtained.

A new approach to prepare porous TE scaffolds is melt electrospinning writing (MEW) [145,149–151]. MEW is essentially applying an extrusion-based additive man-ufacturing approach to melt electrospinning [152]. An electrified polymer melt is extruded through a nozzle onto a grounded, translating and/or rotating platform. As the electrified molten jet rapidly cools in the air and on the platform, well-defined porous structures can be prepared. Furthermore, polymer fibers with diameters smaller than 1mm can be prepared [150]. Chen et al. used MEW to prepare scaffolds from poly(L-lactide-co-e-caprolactone-co-acryloyl carbonate) macromers which were photo-crosslinked to prevent creep and a decrease in the elasticity modulus upon

(18)

hydration [145]. Furthermore, crosslinking increased the average elasticity modulus of the fibers and improved their toughness. The crosslinked scaffolds could be exposed to cyclic strains of 10% elongation for 200,000 cycles without failure, whereas 4 out of 6 non-crosslinked scaffolds failed under the same conditions.

Stereolithography

Stereolithography makes use of a light source to photo-crosslink a polymer resin in a layer-by-layer manner [101,153]. As can be seen inFigure 6, a 3D design of an implant (for example a patient-specific meniscus implant based on a render from CT imaging data) is virtually sliced into 2D layers. The thickness of these layers corresponds to the thickness of the layers in the additive manufacturing process. The data is then uploaded to the control computer and the structures are fabricated by SLA. Of all 3D printing techniques, SLA is the most accurate additive manufacturing technique allowing build-ing of designed structures at the highest resolution. Whereas commercially available SLA setups allow building constructs with details of 20mm in size, other additive manufacturing techniques allow building structures with details in the range of 50–200 mm [101]. High resolution micro-SLA setups have been developed that allow building at resolutions even lower than 20mm, with details from 5 mm [154–156] to close to 40–500 nm in size [157–161]. More recently, an apparatus in which SLA and extrusion are combined has been developed by Shanjani et al. [146].

In general, two types of SLA systems are used to prepare designed tissue engineering structures [38,162]: laser-based SLA and digital light processing SLA (DLP-SLA). In laser-based SLA, a layer of a photo-crosslinkable resin is illuminated at the surface using

Figure 6. From a 3D design towards a porous meniscus implant manufactured by stereolithog-raphy (SLA). A 3D design based on a 3D render from CT imaging data with a gyroid porous net-work architecture is made. 2D slices with thicknesses corresponding to the build layers are then made and converted into pixel patterns. The structure is then manufactured by SLA in a layer-by-layer manner.

(19)

a computer controlled laser beam. The structures are prepared layer-by-layer by moving the build platform down into the resin. In DLP-SLA, a UV or blue light pattern of pixels is projected into the resin through a transparent and non-adherent resin container from below. In this case, the build platform moves upwards and out of the resin.

3D bioprinting

A novel approach to the use of additive manufacturing is 3D bioprinting. 3D bio-printing is based on the layer-by-layer manufacturing of 3D structures consisting of natural and synthetic polymers, living cells, drugs, growth factor and genes with spa-tial control over the placement of these components [163,164]. In 3D bioprinting, tis-sues, organs and other biological systems are prepared in vitro [165]. These constructs can be used ad 3D models to replace 2D in vitro cell culture and animal models. In addition, bone and cartilage tissues have been prepared for musculoskel-etal injury repair [164].

The main additive manufacturing methods used for 3D bioprinting are extrusion-based additive manufacturing, inkjet printing, laser assisted bioprinting and stereolithog-raphy [163,164]. Of these techniques, extrusion-based additive manufacturing [166–169], stereolithography [170], and a combination of these two [146] have been used to pre-pare 3D bioprinted structures by photo-crosslinking. The concerns with regards to the adverse effect of the irradiation wavelength on the cell viability [171] were in a few cases addressed by using visible light [170] and a water filter [146]. Cell viabilities between 80 and 95% were reported in these studies [146,166–168,170].

Photo-crosslinkable systems for additive manufacturing

As additive manufacturing is such an important method to prepare photo-crosslinked biodegradable polymer networks, it is important to discuss the photo-crosslinkable sys-tems that have been developed for use in additive manufacturing. Photo-crosslinkable systems based on fumarate-functionalized oligomers were developed [10,129,172–175]. These systems were based on PDLLA, PCL or poly(propylene fumarate) (PPF). The disadvantage of these materials is that they need a reactive diluent such as N-vinyl-2-pyrrolidone (NVP) [173,175] and diethyl fumarate [129,172,174]. As described earlier, this increases the non-degradable content of the resulting polymer networks. Resins based on (meth)acrylated macromers are therefore perhaps a more suited alternative [176,177], as these are more reactive. In the case of photo-crosslinkable systems for stereolithography (resins), non-reactive diluents can be used to adjust the viscosity of the resin to allow its processing [17,38]. (Note that this non-react-ive diluent needs to be extracted from the built structure.)

Photo-crosslinkable systems based on poly(D,L-lactide)

Methacrylate-functionalized PDLLA oligomers were one of the early photo-reactive compounds used for the preparation of biodegradable tissue engineering scaffolds by SLA [17]. As non-crosslinked HMW PDLLA was already used in bone tissue

(20)

engineering, Melchels et al. proposed to prepare scaffolds for bone tissue engineering by SLA [17]. Using a non-reactive diluent, they developed resins based on such photo-crosslinkable PDLLA macromers. The mechanical properties of the networks and porous scaffolds prepared by SLA were similar to those prepared using HMW PDLLA. In the further development of PDLLA bone tissue engineering scaffolds, nano-hydroxyapatite was incorporated into the resin. As the chemistry of hydroxy-apatite is similar to the calcium phosphate mineral phase present in hard tissues, this may lead to composite scaffolds that enhance bone formation [27,28]. The incorpor-ation of nano-hydroxyapatite into the polymer matrix also resulted in increasing the compressive- and tensile moduli of the networks [27,28].

Photo-crosslinkable systems based on fumarate-functionalized PDLLA oligomers containing 35 wt% NVP reactive diluent were developed for SLA [173]. Highly porous structures, closely matching the designs could be prepared.

Photo-crosslinkable systems based on poly(e-caprolactone)

Designed tissue engineering scaffolds have also been prepared using SLA resins based on methacrylated PCL macromers with relatively low molecular weights [29,178]. In this case, no diluents were required as the resins had sufficiently low viscosity at room temperature or after moderate heating to allow their processing. The scaffolds could be manufactured very accurately, closely resembling the geometry, porosity and pore architecture of the designs [29].

Liquid coumarin end-functionalized copolymers based on CL and TMC were used to prepare microstructured films and surfaces by stereolithography [7,8]. Multilayered films containing three different copolymers in one single construct were prepared with these materials. The polymers crosslinked upon UV irradiation.

A photo-crosslinkable system for SLA based on divinyl-fumarate poly( e-caprolac-tone) was used to prepare tissue engineering scaffolds [175]. Initially, resin were pre-pared with either a non-reactive solvent (N-methyl-2-pyrrolidone, NMP) or with NVP as a reactive diluent. As the best crosslinking was achieved with the reactive diluent under UV light the tissue engineering scaffolds were prepared using that resin. The obtained scaffolds closely resembled the designs.

Photo-crosslinkable systems based on poly(trimethylene carbonate)

For processing in extrusion-based additive manufacturing, PTMC was dissolved in ethylene carbonate (melting point 37C) and used in (LTEAM) [91,94]. After extru-sion of the fibers at 60C, the ethylene carbonate was crystallized at a temperature below the melting temperature of the solvent. This provided the required form-stabil-ity when building the structure. The prepared structures were then photo-crosslinked, and after subsequent extraction of ethylene carbonate, the manufactured structures remained form-stable. Interestingly, this use of crystallizing ethylene carbonate resulted in porous scaffolds with an additional micro-porosity. It has been suggested that these micro-pores may have a beneficial effect on the regenerative capacity of the scaffolds [91].

(21)

Methacrylate-functionalized PTMC macromers have been extensively used in the preparation of designed structures and tissue engineering scaffolds by SLA [38,39,60–63,179]. Non-reactive diluents were used to decrease the viscosity and allow processing. It was shown that the mechanical properties, and especially the moduli, of porous scaffolds strongly depended on porosity and not on the molecular weight of the macromer [179] or on pore size [38]. Furthermore, it was shown that incorporat-ing nano hydroxyapatite into the resin to create composites resulted in increased ten-sile strength and toughness [39].

Photo-crosslinkable systems based on poly(ethylene glycol)

Photo-crosslinkable systems based on meth(acrylated) PEG have been developed as well [124,180,181]. These resins could also contain cells, making the preparation of cell encapsulating scaffolds by SLA possible [40]. As described previously, PEG-based networks are only biodegradable when PEG is block co-polymerized with a bio-degradable component. Scaffolds using a resin based on tri-block copolymer of PDLLA and PEG were prepared by SLA [182]. These scaffolds were hydrogels and highly flexible and the structures matched their design precisely. Furthermore, the mechanical properties of these scaffolds in compression experiments were similar to the properties of soft tissues. In a combinatorial approach PEG was one of the com-ponents in hydrogel mixtures which further included PTMC, PDLLA and PCL to prepare mixed-macromer network scaffolds by SLA[41]. These scaffolds had compres-sion moduli up to 170 kPa.

For 3D bioprinting, a photo-crosslinkable system based on methacrylated PEG was developed [146]. The cells were encapsulated in the PEG subsequently crosslinked after deposition while non-crosslinkable PCL was used as a rigid component.

Photo-crosslinkable systems based on semi-synthetic polymers

A stereolithography resin based on GelMA was developed by Gauvin et al [183]. The resin was prepared to contain calcium carbonate micrioparticles and further con-tained photo-initiator, UV absorber and a solution quencher. Resins based on GEL-MA modified with PEG or methoxy-PEG has been evaluated in preparing designed 3D structures at high resolutions. These resins contained reactive diluents [177] or aqueous solutions of co-monomers [184].

Photo-crosslinkable systems based on gelMA and cells (bioinks) have been widely utilized as systems for 3D bioprinting [166–170]. GelMA was mixed with alginate, photo-initiator and cells [167]. The alginate was used to obtain physically crosslinked fibers that acted as a structural template to prevent the collapsing of the printed structures. After printing the GelMA ws crosslinked by UV irradiation. A bioink with the same components but also containing a 4-arm PEG-tetra-acrylate (PEGTA) was also developed [166]. The addition of PEGTA resulted in increased compression moduli of the obtained structures. Similarly, a GelMA based bioink was developed with PEG, but without the alginate [170]. In addition, bioinks based on GelMA con-taining methacrylated hyaluronic acid have been investigated [168,169].

(22)

Photo-crosslinkable systems based on other polymers

Photo-crosslinkable systems based on other polymeric biomaterials have been investi-gated as well.

Resins for m-SLA based on vinylesters and vinylcarbonates have been investigated [156,185]. These resin contained mixtures of monomers, photo-initiator and a UV-absorber. Using these resins, highly accurate designed 3D structures could be prepared.

To prepare designed 3D ceramic structures, photo-crosslinkable composite resins have been developed as well [186,187]. In these resins, a ceramic powder is dispersed in a solution of acrylate-based monomers. After fabrication of the designed structures by SLA, heating of the green body to elevated temperatures leads to decomposition of the polymer network phase and sintering of the ceramic particles.

Conclusions

In this review, a short overview is given of the field of photo-crosslinked synthetic biodegradable polymer networks. These materials can be prepared form a wide range of materials such as polyesters, polycarbonates, PEG-based materials and semi-syn-thetic polymers based on natural polymers. Photo-crosslinking these materials results in enormously versatile materials as varying in materials, crosslink density and mac-romer molecular weight results in networks with diverse mechanical properties, deg-radation properties and applications. These applications include scaffolds for tissue engineering, drug delivery devices and cell encapsulation.

An interesting method to prepare photo-crosslinked networks for use in these applica-tions is additive manufacturing. Additive manufacturing allows for the 3D preparation of photo-crosslinked constructs at a high resolution and in complex structures. Additive man-ufacturing is a relatively new technique which has prospect in many fields from medicine and engineering to life sciences and has become much more affordable and accessible in recent years. Combined with the high versatility of photo-crosslinked networks, and for example the recent developments in 3D bioprinting, this may result in significant break-throughs in the improvement of current, and development of novel generation constructs that deliver drugs at a controlled rate and replace and repair damaged tissues.

Disclosure statement

No potential conflict of interest was reported by the authors.

Funding

This work was supported by the Netherlands Organization for Scientific Research (NWO) under Stichting voor de Technische Wetenschappen project number 12410.

References

[1] Amsden B. Curable, Biodegradable elastomers: emerging biomaterials for drug delivery and tissue engineering. Soft Matter. 2007;3:1335–1348.

(23)

[2] Kaihara S, Vacanti JP. Tissue engineering: toward new solutions for transplantation and reconstructive surgery. Arch Surg. 1999;134:1184–1188.

[3] Ifkovits JL, Burdick JA. Review: photopolymerizable and degradable biomaterials for tissue engineering applications. Tissue Eng. 2007;13:2369–2385.

[4] Pereira MJN, Ouyang B, Sundback CA, et al. A highly tunable biocompatible and multifunctional biodegradable elastomer. Adv Mater. 2013;25:1209–1215.

[5] Pego AP, Grijpma DW, Feijen J. Enhanced mechanical properties of 1,3-trimethylene carbonate polymers and networks. Polymer. 2003;44:6495–6504.

[6] Matsuda T, Magoshi T. Preparation of vinylated polysaccharides and photofabrication of tubular scaffolds as potential use in tissue engineering. Biomacromolecules. 2002;3: 942–950.

[7] Matsuda T, Mizutani M, Arnold SC. Molecular design of photocurable liquid bio-degradable copolymers. 1. Synthesis and photocuring characteristics. Macromolecules. 2000;33:795–800.

[8] Mizutani M, Matsuda T. Liquid photocurable biodegradable copolymers: in vivo deg-radation of photocured poly(epsilon-caprolactone-co-trimethylene carbonate). J Biomed Mater Res. 2002;61:53–60.

[9] Grijpma DW, Hou Q, Feijen J. Preparation of biodegradable networks by photo-cross-linking lactide, epsilon-caprolactone and trimethylene carbonate-based oligomers func-tionalized with fumaric acid monoethyl ester. Biomaterials. 2005;26:2795–2802. [10] Cooke MN, Fisher JP, Dean D, et al. Use of stereolithography to manufacture

critical-sized 3D biodegradable scaffolds for bone ingrowth. J Biomed Mater Res. 2003;64: 65–69.

[11] Jansen J, Ghaffar A, van der Horst TN, et al. Controlling the kinetic chain length of the crosslinks in photo-polymerized biodegradable networks. J Mater Sci: Mater Med. 2013;24:877–888.

[12] Bat E, Zhang Z, Feijen J, et al. Biodegradable elastomers for biomedical applications and regenerative medicine. Regen Med. 2014;9:385–398.

[13] Shin H, Temenoff JS, Mikos AG. In vitro cytotoxicity of unsaturated oligo[poly(ethy-lene glycol) fumarate] macromers and their cross-linked hydrogels. Biomacromolecules. 2003;4:552–560.

[14] Domb AJ, Mathiowitz E, Ron E, et al. Polyanhydrides. 4. Unsaturated and cross-linked polyanhydrides. J Polym Sci A Polym Chem. 1991;29:571–579.

[15] He S, Timmer MD, Yaszemski MJ, et al. Synthesis of biodegradable poly(propylene fumarate) networks with poly(propylene fumarate)-diacrylate macromers as crosslink-ing agents and characterization of their degradation products. Polymer. 2001;42: 1251–1260.

[16] Storey RF, Warren SC, Allison CJ, et al. Methacrylate-endcapped poly(d,l-lactide-co-trimethylene carbonate) oligomers. Network formation by thermal free-radical curing. Polymer. 1997;38:6295–6301.

[17] Melchels FPW, Feijen J, Grijpma DW. A poly(D,L-lactide) resin for the preparation of tissue engineering scaffolds by stereolithography. Biomaterials. 2009;30:3801–3809. [18] Helminen AO, Korhonen H, Seppala JV. Crosslinked poly(ester anhydride)s based on

poly(epsilon-caprolactone) and polylactide oligomers. J Polym Sci A Polym Chem. 2003;41:3788–3797.

[19] Turunen MPK, Korhonen H, Tuominen J, et al. Synthesis, characterization and cross-linking of functional star-shaped poly(epsilon-caprolactone). Polym Int. 2002;51: 92–100.

[20] Matsuda T, Kwon IK, Kidoaki S. Photocurable biodegradable liquid copolymers: Synthesis of acrylate-end-capped trimethylene carbonate-based prepolymers, photocur-ing, and hydrolysis. Biomacromolecules. 2004;5:295–305.

[21] Cornacchione LA, Qi B, Bianco J, et al. Photo-cross-linked poly(ethylene carbonate) elastomers: synthesis, in vivo degradation, and determination of in vivo degradation mechanism. Biomacromolecules. 2012;13:3099–3107.

(24)

[22] Kim BS, Hrkach JS, Langer R. Biodegradable photo-crosslinked poly(ether-ester) net-works for lubricious coatings. Biomaterials. 2000;21:259–265.

[23] Sawhney AS, Pathak CP, Hubbell JA. Bioerodible hydrogels based on photopolymer-ized poly(ethylene glycol)-co-poly(alpha-hydroxy acid) diacrylate macromers. Macromolecules. 1993;26:581–587.

[24] Storey RF, Hickey TP. Degradable polyurethane networks based on d,l-lactide, glyco-lide, e-caprolactone, and trimethylene carbonate homopolyester and copolyester triols. Polymer. 1994;35:830–838.

[25] Korhonen H, Helminen A, Seppala JV. Synthesis of polylactides in the presence of co-initiators with different numbers of hydroxyl groups. Polymer. 2001;42:7541–7549. [26] Melchels FPW, Velders AH, Feijen J, et al. Photo-cross-linked poly(DL-lactide)-based

networks. structural characterization by HR-MAS NMR spectroscopy and hydrolytic degradation behavior. Macromolecules. 2010;43:8570–8579.

[27] Ronca A, Ambrosio L, Grijpma DW. Preparation of designed poly(D,L-lactide)/nano-sized hydroxyapatite composite structures by stereolithography. Acta Biomater. 2013;9: 5989–5996.

[28] Ronca A, Ambrosio L, Grijpma DW. Design of porous three-dimensional PDLLA/ nano-hap composite scaffolds using stereolithography. JABFM. 2012;10:249–258. [29] Elomaa L, Teixeira S, Hakala R, et al. Preparation of poly(epsilon-caprolactone)-based

tissue engineering scaffolds by stereolithography. Acta Biomater. 2011;7:3850–3856. [30] Zant E, Grijpma DW. Synthetic biodegradable hydrogels with excellent mechanical

properties and good cell adhesion characteristics obtained by the combinatorial syn-thesis of photo-cross-linked networks. Biomacromolecules. 2016;17:1582–1592. [31] Zant E, Grijpma D. Tough biodegradable mixed-macromer networks and hydrogels by

photo-crosslinking in solution. Acta Biomater. 2016;31:80–88.

[32] Nijst CLE, Bruggeman JP, Karp JM, et al. Synthesis and characterization of photocura-ble elastomers from poly(glycerol-co-sebacate). Biomacromolecules. 2007;8:3067–3073. [33] Gerecht S, Townsend SA, Pressler H, et al. A porous photocurable elastomer for cell

encapsulation and culture. Biomaterials. 2007;28:4826–4835.

[34] Ifkovits JL, Padera RF, Burdick JA. Biodegradable and radically polymerized elasto-mers with enhanced processing capabilities. Biomed Mater. 2008;3:034104.

[35] Mahdavi A, Ferreira L, Sundback C, et al. A biodegradable and biocompatible gecko-inspired tissue adhesive. Proc Natl Acad Sci USA. 2008;105:2307–2312.

[36] Tran RT, Yang J, Ameer GA. Citrate-based biomaterials and their applications in regenerative engineering. Annu Rev Mater Res. 2015;45:277–310.

[37] Schuller-Ravoo S, Feijen J, Grijpma DW. Flexible, elastic and tear-resistant networks prepared by photo-crosslinking poly(trimethylene carbonate) macromers. Acta Biomater. 2012;8:3576–3585.

[38] van Bochove B, Hannink G, Buma P, et al. Preparation of designed poly(trimethylene carbonate) meniscus implants by stereolithography: challenges in stereolithography. Macromol Biosci. 2016;16:1853–1863.

[39] Geven MA, Varjas V, Kamer L, et al. Fabrication of patient specific composite orbital floor implants by stereolithography. Polym Adv Technol. 2015;26:1433–1438.

[40] Lu Y, Mapili G, Suhali G, et al. A digital micro-mirror device-based system for the microfabrication of complex, spatially patterned tissue engineering scaffolds. J Biomed Mater Res A. 2006;77: 396–405.

[41] Przeradzka MA, van Bochove B, Bor TC, et al. Phase-separated mixed-macromer hydrogel networks and scaffolds prepared by stereolithography. Polym Adv Technol. 2017;28:1212–1218.

[42] Yue K, Trujillo-de Santiago G, Alvarez MM, et al. Synthesis, properties, and biomed-ical applications of gelatin methacryloyl (GelMA) hydrogels. Biomaterials. 2015;73: 254–271.

[43] Chen Y, Lin R, Qi H, et al. Functional human vascular network generated in photo-crosslinkable gelatin methacrylate hydrogels. Adv Funct Mater. 2012;22:2027–2039.

(25)

[44] Nichol JW, Koshy ST, Bae H, et al. Cell-laden microengineered gelatin methacrylate hydrogels. Biomaterials. 2010;31:5536–5544.

[45] Zhou Y, Ma G, Shi S, et al. Photopolymerized water-soluble chitosan-based hydrogel as potential use in tissue engineering. Int J Biol Macromol. 2011;48:408–413.

[46] Hu J, Hou Y, Park H, et al. Visible light crosslinkable chitosan hydrogels for tissue engineering. Acta Biomaterialia. 2012;8:1730–1738.

[47] Ma G, Yang D, Li Q, et al. Injectable hydrogels based on chitosan derivative/polyethyl-ene glycol dimethacrylate/N,N-dimethylacrylamide as bone tissue engineering matrix. Carbohyrdate Polymers. 2010;79:620–627.

[48] Ono K, Saito Y, Yura H, et al. Photocrosslinkable chitosan as a biological adhesive Katsuaki. J Biomed Mater Res. 2000;49:289–295.

[49] Engelberg I, Kohn J. Physico-mechanical properties of degradable polymers used in medical applications: a comparative study. Biomaterials. 1991;12:292–304.

[50] Amsden BG, Misra G, Gu F, et al. Synthesis and characterization of a photo-cross-linked biodegradable elastomer. Biomacromolecules. 2004;5:2479–2486.

[51] Chapanian R, Tse MY, Pang SC, et al. The role of oxidation and enzymatic hydrolysis on the in vivo degradation of trimethylene carbonate based photocrosslinkable elasto-mers. Biomaterials. 2009;30:295–306.

[52] Amsden BG, Tse MY, Turner ND, et al. In vivo degradation behavior of photo-cross-linked star-poly(epsilon-caprolactone-co-D,L-lactide) elastomers. Biomacromolecules. 2006;7:365–372.

[53] Ilagan BG, Amsden BG. Macroporous photocrosslinked elastomer scaffolds containing microposity: preparation and in vitro degradation properties. J Biomed Mater Res A. 2010;93:211–218.

[54] Wang Y, Ameer GA, Sheppard BJ, et al. A tough biodegradable elastomer. Nat Biotechnol. 2002;20:602–606.

[55] Rai R, Tallawi M, Grigore A, et al. Synthesis, properties and biomedical applications of poly(glycerol sebacate) (PGS): A review. Prog Polym Sci. 2012;37:1051–1078. [56] Zhao H, Ameer GA. Modulating the mechanical properties of poly(diol citrates) via

the incorporation of a second type of crosslink network. J Appl Polym Sci. 2009;114: 1464–1470.

[57] Gyawali D, Tran RT, Guleserian KJ, et al. Citric-acid-derived photo-cross-linked bio-degradable elastomers. J Biomater Sci Polym Ed. 2010;21:1761–1782.

[58] Wang Y, Kibbe MR, Ameer GA. Photo-crosslinked biodegradable elastomers for con-trolled nitric oxide delivery. Biomater Sci. 2013;1:625–632.

[59] Zhu KJ, Hendren RW, Jensen K, et al. Synthesis, properties, and biodegradation of poly(1,3-trimethylene carbonate). Macromolecules. 1991;24:1736–1740.

[60] Schuller-Ravoo S, Teixeira SM, Feijen J, et al. Flexible and elastic scaffolds for cartilage tissue engineering prepared by stereolithography using poly(trimethylene carbonate)-based resins. Macromol Biosci. 2013;13:1711–1719.

[61] Blanquer SB, Sharifi S, Grijpma DW. Development of poly(trimethylene carbonate) network implants for annulus fibrosus tissue engineering. JABFM. 2012;10:177–184. [62] Blanquer SBG, Gebraad AWH, Miettinen S, et al. Differentiation of adipose stem cells

seeded towards annulus fibrosus cells on a designed poly(trimethylene carbonate) scaf-fold prepared by stereolithography. J Tissue Eng Regen Med. 2017;11:2752–2762. [63] Schuller-Ravoo S, Zant E, Feijen J, et al. Preparation of a designed poly(trimethylene

carbonate) microvascular network by stereolithography. Adv Healthcare Mater. 2014;3: 2004–2011.

[64] Chapanian R, Amsden BG. Combined and sequential delivery of bioactive VEGF165 and HGF from poly(trimethylene carbonate) based photo-cross-linked elastomers . J Control Release. 2010;143:53–63.

[65] Sharifi S, Blanquer S, Grijpma DW. Polymeric microstructures with shape-memory properties for biomedical use built by stereolithography. JABFM. 2012;10:280–286.

Referenties

GERELATEERDE DOCUMENTEN

Consultation with fellow scientists in the area that have a keen interest in amphibian species in the Southern Cape (pers. W Matthee, NMU) and information gleaned from

of three di fferent TMD surfaces (Figure S2); height of the depression for dark defects (Figure S3); types of dark defects (Figure S4); STS of the bright defect, dark defect,

”Exact Plant Inversion of Flexible Motion Systems with a Time- Varying State-To-Output Map”, European Control Conference, pp. ”Composite jerk feedforward and disturbance observer

The aims of this systematic review were to determine the therapeutic validity and to assess the effectiveness of physiotherapeutic exercise interventions for joint and muscle

In the burial field, five seals made of some other type of stone than steatite were recovered, three of which are possibly Persian Gulf style or Proto-Dilmun style.. Neither of

Therefore, we sought to develop a model incorporating clinical risk factors for predicting 5-year recurrent ischemic stroke in patients with ischemic stroke and to determine

We report the electrical top-gating of a 2-dimensional electron gas (2DEG) formed at the LaAlO 3 /SrTiO 3 interface, using electron-beam evaporated Au gate electrodes.. In

The analysis of educational data can result in an understanding of what learners are doing (which is what EDM predominantly aims at), changing the behaviour of the learner (through