PDF - UWA Research Portal
Transcription
PDF - UWA Research Portal
Development of a drug delivery platform using multifunctional polymeric scaffold for scar therapy Vipul Agarwal, MApplSc This thesis is presented for the degree of Doctor of Philosophy at The University of Western Australia School of Chemistry and Biochemistry 2015 Contents Abbreviations .............................................................................................................. iv Abstract ......................................................................................................................vii Acknowledgement ...................................................................................................... ix Statement of candidate contribution ........................................................................... xi 1 Introduction and literature review........................................................................ 1 1.1 Tissue engineering.............................................................................................. 2 1.2 Cellular scaffolds................................................................................................ 3 1.3 Physical features of the cell microenvironment ................................................. 4 1.4 Design concepts and strategies ........................................................................... 5 1.4.1 Physical properties ...................................................................................... 6 1.4.2 Mechanical properties ................................................................................. 7 1.4.3 Surface properties ........................................................................................ 8 1.5 Skin, injury and wound healing.......................................................................... 9 1.5.1 Haemostasis and inflammation ................................................................. 11 1.5.2 Reepithelialisation ..................................................................................... 13 1.5.3 Formation of granulation tissue ................................................................. 16 1.6 Role of TGFβ in fibrosis and scarring .............................................................. 17 1.7 Mechanisms of activation of extracellular TGFβ ............................................. 19 1.8 Exogenous mannose-6-phosphate act as an antagonist of TGFβ1 activation ... 20 1.9 Structural requirements for mannose-6-phosphate recognition to M6P/IGFII receptor ...................................................................................................................... 22 1.10 Mannose-6-phosphate analogues ..................................................................... 23 1.11 Scarring ............................................................................................................ 27 1.12 Current treatments for skin wound healing ...................................................... 28 1.13 Cell based therapies .......................................................................................... 29 1.14 Matrix based therapies ..................................................................................... 30 i 1.14.1 Amniotic membrane ................................................................................ 30 1.14.2 Human cadaver allografts and xenografts ............................................... 30 1.14.3 Tissue engineered skin ............................................................................ 32 1.15 Nanofibrous scaffolds for skin tissue engineering ........................................... 34 1.16 Skin regeneration using electrospun scaffolds ................................................. 37 1.17 Electrospun hybrid materials ............................................................................ 41 1.18 Summary .......................................................................................................... 44 1.19 Hypotheses and Aims ....................................................................................... 46 2 Introduction to the series of papers .................................................................... 48 2.1 Development of a universal multifunctional scaffold ......................................... 48 2.2 Evaluation of mannose-6-phosphate analogues as potential anti-scarring agents ... .......................................................................................................................... 51 2.3 Delivery of the lipophilic mannose-6-phosphate analogue PXS64 using an electrospun PGMA scaffold ...................................................................................... 53 3 Series of papers ..................................................................................................... 56 A Functional Reactive Polymer Nanofiber Matrix .................................................... 57 Enhancing the Efficacy of Cation-Independent Mannose 6-Phosphate Receptor Inhibitors by Intracellular Delivery ........................................................................... 62 Inhibiting the activation of transforming growth factor-β using a polymeric nanofiber scaffold ...................................................................................................... 67 4 Conclusions and future work ............................................................................... 71 4.1 Tissue engineering of a nanoscaffold ............................................................... 71 4.2 Scar therapy and mannose-6-phosphate analogues .......................................... 73 4.3 Scaffold based delivery of analogue 2 ............................................................. 74 4.4 Future recommendations .................................................................................. 75 4.5 Final remarks .................................................................................................... 77 A Supporting information for papers .................................................................... 79 Supporting Information for ‘A Functional Reactive Polymer Nanofiber Matrix’ .... 80 ii Supporting information for ‘Enhancing the Efficacy of Cation-Independent Mannose 6-Phosphate Receptor Inhibitors by Intracellular Delivery’ ...................................... 87 Supporting Information for ‘Inhibiting the activation of transforming growth factorβ using a polymeric nanofiber scaffold’ .................................................................... 98 B Published work not directly included in the thesis ......................................... 103 References................................................................................................................ 147 iii Abbreviations ANOVA Analysis of variance bFGF Basic fibroblasts growth factor CDMPR Cation dependent mannose-6-phosphate receptor CIMPR Cation independent mannose-6-phosphate receptor DBM Demineralized bone matrix DCFH-DA 2’, 7’-dichlorodihydrofluorescein diacetate ECM Extracellular matrix EGF Epidermal growth factor ES Electrospun FTIR Fourier transform infra-red GAG Glycosaminoglycan HCAS Human cadaver allograft skin HDF Human dermal skin fibroblasts HPLC High pressure liquid chromatography HSF Human scar fibroblasts LAP Latency associated peptide LCST Lower critical solution temperature LTGFβ Latent transforming growth factor β M6P Mannose-6-phosphate M6P/IGF II Mannose-6-phosphate/Insulin-like growth factor II MMP Matrix metalloproteinase iv MRI Magnetic resonance imaging NMR Nuclear magnetic resonance NP Nanoparticle Pa Pascal PBS Phosphate buffer saline PCL Poly(ε-caprolactone) Pd Palladium PEG Poly(ethylene glycol) PGA Poly(glycolic acid) PGMA Poly(glycidyl methacrylate) PLA Poly(lactic acid) PLGA Poly(lactic-co-glycolic acid) PNIPAM Poly (N-isopropyl acrylamide) PU Polyurethane ROS Reactive oxygen species RT-qPCR Real time quantitative-polymerase chain reaction SEM Scanning electron microscopy SQUID Superconducting quantum interference device TE Tissue engineering TEM Transmission electron microscopy TGFβ Transforming growth factor β TSP Thrombospondin UCNP Upconverting nanoparticles v VEGF Vascular endothelial growth factor vi Abstract Tissue engineering is a multidisciplinary approach and has been used to promote tissue regeneration, wound repair and to enhance drug delivery. In the case of burn injury, despite the advances in treatment leading to reduced mortality, the problem of permanent, disfiguring scar formation following healing is far from being resolved. Scarring is the result of an imbalance of fibroblast activity resulting in an excess of architecturally disorganised extracellular matrix protein deposition and is predominantly mediated by the TGFβ pathway. Whilst there are a number of promising therapeutic targets identified through our increasing understanding of scar formation, there has been limited success in the clinical translation. This can largely be attributed to difficulties with delivery, stability and efficacy of treatments tested to date. In this thesis the problems of drug delivery and stability have been addressed using a combinatorial approach. First, a scaffold was developed to provide a platform for drug delivery and stability. A novel multifunctional polyglycidyl methacrylate (PGMA) scaffold was developed using electrospinning. The multifunctionality of this polymeric scaffold was demonstrated for various applications. These include surface functionalization of electrospun PGMA with poly (N-isopropyl acrylamide) for stimuli response surfaces and development of multifunctional nanocomposites with (NaGdF4:Yb, Er); Pd and Fe3O4 nanoparticles for upconversion fluorescence imaging, sensing, and magneto-responsive properties. This was followed by exploration of modified analogues of a potential therapeutic target as anti-scarring agents. Mannose-6-phosphate (M6P) has been shown to ameliorate scarring by inhibiting the activation of TGFβ1, a necessary step required for its receptor recognition and function. However, therapeutic delivery of M6P to a wound is currently ineffective due to the low stability of M6P and limited ability to maintain M6P concentration at the site of injury. Therefore whilst targeting TGFβ activity through M6P has significant potential in burn therapy, stability and delivery issues must be addressed before this can become a therapeutic reality. Two M6P analogues, vii PXS25 (analogue 1) and PXS64 (analogue 2), have been developed in collaboration with Pharmaxis Ltd., to overcome the metabolic vulnerability of M6P whilst retaining the receptor recognition function. Initial work was carried out to investigate the biocompatibility and cytotoxicity capabilities of both analogues compared to M6P in human dermal skin fibroblasts. Subsequently they were investigated for their proficiency in regulating the expression of the critical fibrotic marker, Collagen I. Analogue 2 was shown to significantly inhibit TGFβ1 mediated up-regulation of collagen I gene expression. However, the lipophilic analogue 2 has limited bioavailability. The final chapter addresses this specific problem by incorporating analogue 2 into the multifunctional scaffold and testing the efficacy of the combined drug/scaffold therapy on human dermal skin fibroblasts. The tissue engineering approach presented herein demonstrated the potential of combinatorial scaffold mediated drug delivery method to progress some of the existing therapeutic targets into clinical therapies. The future work using porcine wound healing model is a necessary extension to establish the efficacy and potential of this combinatorial approach in viii vivo before its clinical translation. Acknowledgement I would like to begin by thanking my supervisors for their consistent support throughout my candidature. I am really grateful to Prof Swaminathan Iyer for offering me PhD scholarship and introduce me to a completely new stream of research, to keep me on track and motivate me especially during the difficult times and moulding me into a better researcher; Prof Fiona Wood for support and guidance and brining clinical prospective into my work; Dr Mark Fear for bringing biological insight, enabling me to resolve complex experimental problems and improving my overall understanding. In addition, I would extend my gratitude towards Prof Iyer and Dr Fear for going above and beyond to help me with my grant application, you faith and support only project me to improve and become an accomplished scientist. I would also like to thank my colleagues Dr Cameron Evans, Dominic Ho, Diwei Ho, Dr Faizah Yasin, Alaa Munshi, Ruhani Singh, Ivan Lozic, and Michael Bradshaw, Dr Rahi Varsani, Dr Tristan Clemons and Michael Challenor for making me welcome and accepting me within the group with all the humour and pranks and also to support me where possible in the lab. I would like to acknowledge Dr Tristan Clemons, Dr Nicole Smith, Sumi Shrestha, Callum Ormonde, Diwei Ho and Dr Marck Norret to proof reading this thesis. I would like to thank the staff at CMCA especially Lyn Kirilak and Prof Paul Rigby for their extended support with microscopy instrumentation and experiments. I would like to also thank Pharmaxis Ltd and the team for inviting me into their lab and welcoming me within their group, and also to consistently support me during my candidature. Because of the multidisciplinary aspect to my project, I had the privilege to work with some really good people including Prof Charlie Bond, Dr Bernard Callus, Megan Finch, Benjamin Gully, Dr Foteini Hassiotou, Dr Ben Corry, Dr Natalie Smith, Dulharie Wijeratne and Dr Lindsay Byrne, without whom this thesis would not have been possible. For the motivation, persistent support and having faith in me to pursue my research ambitions, I would like to thank Dr Karen Stack, Dr Stephen Newbery, Dr Manab Sharma, Sinu Sharma and Murray Frith. For funding, I would like to thank Australian Research Council (ARC), National Health & Medical ix Research Council (NHMRC), Australian Nanotechnology Network (ANN), Australian Academy of Science and The University of Western Australia. Finally, I would like to thank my family, my girlfriend and my friends for their faith and selfless support enabling me to pursue my goals. x Statement of candidate contribution This thesis contained published work and work prepared for publication, some of which has been co-authored. The bibliographical details of the work and author contributions are outlined below. Refereed journal articles included in the series of papers 1. Agarwal, V., Ho, D., Ho, D., Galabura, Y., Yasin, F. M.D., Gong, P., Ye, W., Singh, R., Munshi, A., Saunders, M., Woodward, R. C., St. Pierre, T., Wood, F.M., Fear, M., Lorenser, D., Sampson, D. D., Zdyrko, B., Smith, N.M., Luzinov, I., Iyer, K.S., A Functional Reactive Polymer Nanofiber Matrix, RSC Advances (Submitted) VA and DH developed the initial concept of PGMA electrospinning. VA collaborated with WY and DH to synthesize upconverting nanoparticles, AM to synthesize palladium nanoparticles and acquired TEM images on nanoparticles, RH to synthesize magnetite nanoparticles and acquired TEM images on fibers, PG to perform NIR Room Temperature Emission Spectroscopy, FMDY to perform hydrogen gas sensing analysis and RCW to perform squid analysis. YG, BZ, IL provided the PGMA; remaining authors supervised the work. Contribution by VA: 75% 2. Agarwal, V., Toshniwal, P., Smith, N. E., Smith, N. M., Li, B., Clemons, T. D., Byrne, L. T., Hassiotou, F., Wood, F. M., Fear, M., Corry, B., and Iyer, K. S., Enhancing the Efficacy of Cation-Independent Mannose 6-Phosphate Receptor Inhibitors by Intracellular Delivery, Angewandte Chemie International Edition (Submitted) xi PT and TDC carried out protein expression studies. NES and BC performed docking experiments; rest all the authors supervised the work. Contribution by VA: 90% 3. Agarwal, V., Wood, F. M., Fear, M., and Iyer, K. S., Inhibiting the activation of transforming growth factor-β using a polymeric nanofiber scaffold, Nanoscale (submitted) FMW, MF and KSI supervised the work. Contribution by VA: 90% Refereed journal articles included in the appendix 1. Agarwal, V., Tjandra, E.S., Iyer, K. S., Humfrey, B., Fear, M., Wood, F. W., Dunlop, S. and Raston, C. L., Evaluating the effects of nacre on human skin and scar cells in culture, Toxicology Research, 3, 223-227 (2014) EST performed the viability and reactive oxygen species assay on HaCaTs; BH provided the pearl shells; remaining authors supervised the work. Contribution by VA: 90% 2. Eroglu, E., Chen, X., Bradshaw, M., Agarwal, V. , Zou, J., Stewart, S.G., Duan, X., Lamb, R.N., Smith, S.M., Raston, C. and Iyer, K. S., Biogenic production of palladium nanocrystals using microalgae and their immobilization on chitosan nanofibers for catalytic applications, RSC Advances, 3, 1009-1012 (2013) VA performed and optimised protocol for electrospinning, performed characterisation and contributed to the manuscript. Contribution by VA: 25% 3. Eroglu, E., Agarwal, V., Bradshaw, M., Chen, X., Smith, S. M., Raston, C. and Iyer, K. S., Nitrate removal from liquid effluents using microalgae immobilized on chitosan nanofiber mats, Green chemistry, 14 (10), 2682-2685 (2012) VA performed and optimised protocol for electrospinning, performed characterisation and contributed to the manuscript. Contribution by VA: 40% xii Conference presentations 1. Agarwal, V., Schilter, H., Ho, D., Guo, L., Hassiotou, F., Jarolimek, W., Wood, F. M., Fear, M., Iyer, K.S., An Innovative Tissue Engineering Approach Towards Scar Therapy, Fifth International NanoMedicine Conference, Sydney, Australia, 30 June- 2 July 2014 (Oral presentation) 2. Agarwal, V., Ho, D., Schilter, H., Guo, L., Hassiotou, F., Jarolimek, W., Wood, F. M., Fear, M., Iyer, K.S., Novel Scaffold Approach Towards Scar Therapy, International Conference on Nanotechnology and Nanoscience (ICONN2014), Adelaide, Australia 2014 (Poster presentation) 3. Agarwal, V., Schilter, H., Guo, L., Jarolimek, W., Wood, F. M., Fear, M., Iyer, K.S., Scarless wound healing: A new approach, Sixth International Conference on Advanced Materials and Nanotechnology (AMN-6), Auckland, New Zealand, 1115 February 2013 (Poster presentation) 4. Agarwal, V., Wood, F. M., Fear, M., Iyer, K.S., 6. Development of scarless wound healing platforms, School of Chemistry and Biochemistry Research Forum, University of Western Australia, Perth, Australia, 2 November 2012 (Poster presentation award) 5. Agarwal, V., Wood, F. M., Fear, M., Iyer, K.S., Development of Hybrid Hydrogel for Scarless Wound Healing, First International Conference on BioNano Innovation (ICBNI 2012), Brisbane, Australia, 18-20 July 2012 (Oral presentation) 6. Agarwal, V., Wood, F. M., Fear, M., Iyer, K.S., Regulating the Migratory Behaviour of Fibroblasts on a Hydrogel Scaffolds, International Conference on xiii Nanotechnology and Nanoscience (ICONN2012), Perth, Australia, 5-9 February 2012 (Oral presentation) 7. Agarwal, V., Eroglu, E., W., Wood, F. M., Fear, M., Iyer, K.S., In vitro evaluation of electrospun pluronic F-127 dimethacrylate copolymer towards wound healing in burn injuries, Australian Nanotechnology Network Early Career Symposium, Sydney, Australia, 21–22 November 2011 (Oral presentation) Vipul Agarwal Prof Swaminathan Iyer Candidate Co-ordinating Supervisor xiv Chapter 1 Introduction and Literature Review Each year 100 million people develop scars in the developed world alone,1 with causes ranging from elective surgery to severe trauma. Scars result not only in aesthetic deficits but also long-term functional and psychological problems.2 According to WHO estimates, 6 million people suffer from burn injury each year of which over 300,000 succumb to their injuries.3 Further, chronic skin ulcers contribute an additional 6 million patients annually.4 In attempts to address the functional and aesthetic deficits, the field of tissue engineering has seen a surge of interest and applications in recent years. However, current clinical treatment of scarring still centres on a surgical approach. Small molecules such as mannose-6phosphate (and many other biological factors) have been shown to have significant potential to promote healing and reduce scarring both in elective surgery and after injury. However, the potential of these factors to deliver improvements in the clinic has been significantly hampered by issues of stability and delivery. This thesis will describe the preparation of a multifunctional universal nanofibrous scaffold with the potential to be used as a delivery vehicle for wound repair. Subsequent work will assess the potential of more stable mannose-6-phosphate analogues which may overcome some of the limitations of the endogenous ligand that has previously been trialled. Finally, a combinatorial approach will be investigated using the scaffold/analogue combined. In this chapter the literature detailing the application of tissue engineering and the key considerations for tissue engineering approaches will be discussed. The pathophysiology of skin injury and wound healing will also be reviewed. Finally, the cell therapy and tissue engineering approaches to skin repair will be discussed and the concept of multifunctional scaffolds, electrospinning and their potential for drug delivery and enhanced healing will be introduced. 1 1.1 Tissue engineering Tissue engineering (TE) has evolved as an interdisciplinary science combining principles from nature and integrating them with material and engineering science with the goal of developing functional substitutes for damaged tissues and organs. 5 The materials serve as temporary scaffolds and promote the organisation of cells to form a functional tissue with the cells providing a source for repopulation.6 The general strategy behind TE is to seed cells within a scaffold that mimics the architecture of the replacement tissue while providing environmental cues that promote tissue regeneration. Tissue engineered skin equivalents have been in clinical use since 1997.7 In tissue engineering, a scaffold is usually required to provide a platform or niche that promotes the desired behaviour of cells (for example proliferation of epidermal cells to promote wound coverage).8 The scaffolds are intended to replicate or enhance the natural ECM environment in order to retain cell viability and control cellular behaviour. For example, demineralized bone matrix (DBM, bone from which mineral and cells have been removed, leaving only proteinaceous material) have been implanted in the muscle to initiate bone formation in the neighbouring muscle tissue. This led to the commercial production of recombinant DBM from cadavers for implantation in bone defects.7 There are three important criteria to be considered for the development of tissue engineering scaffolds: i) the 3-dimensional micro-structure of the scaffold such as inter-connectivity of the network pores and subsequent porosity allowing the exchange of gases, nutrients and waste and facilitating cell adhesion, spreading and tissue formation, ii) mechanical parameters catering to specific tissue type such as scaffold morphology like linearity, plasticity, flexibility or anisotropy, and iii) successful delivery of cells, drugs, growth factors and/or cytokines.8 2 1.2 Cellular scaffolds Cellular scaffolds have been developed using natural and synthetic materials and their composites, harnessing the specific advantages of each component and amalgamating them to achieve the desired application. Usually, natural materials are based on purified ECM components (eg. collagen, gelatin, laminin, hyaluronic acid)9-11, often containing multiple components to create a composite substrate (Matrigel®, Integra®).12 Other natural materials are also commonly used, mainly derived from plant or animal sources including silk, agarose and chitosan.13 The main advantages associated with the use of natural materials are their biological activity and biocompatibility. In contrast, synthetic materials are employed to overcome the shortfalls experienced in the use of natural materials, for example manufacturing and process variability and an inability to control their physicochemical properties.8 Single component scaffolds lack the complexity and often functionality of biomaterial scaffolds, such as mechanical properties, electrical activity or cues for cell-matrix interactions14.8 Composite materials are therefore generally preferred. For example, bone is made up of collagen and ceramic like hydroxyapatite, based on which polymer-ceramic composites have been widely used in bone tissue engineering15-19.8 Hydrogels have also been explored as they can imitate the crosslinked architecture of ECM through their cross-linked network of monomers, oligomers or polymers.20 However, the major limitation in their use is their lack of mechanical strength necessary for tissue engineering applications. 8 Gelatin methacrylate hydrogels for example are known to promote cell adhesion and spreading but lack the required mechanical strength.10 Carbon nanotubes have been used to reinforce the mechanical properties of gelatin methacrylate by significantly enhancing the compressive modulus, while retaining the porosity and cell adhesiveness21.8 In designing a scaffold, one of the most desirable features is biodegradability; they must gradually degrade over time and simultaneously get substituted by the naturally deposited ECM and newly formed tissue.8 Consequently, linear aliphatic polyesters such as poly(lactic acid), poly(glycolic acid) and poly(ε-caprolactone) have been 3 extensively investigated owing to their biodegradability, achieved through the hydrolysis of their ester bonds and the ability to control their degradation rate.8 Scaffold structure is another important feature that needs to be considered during fabrication of cellular scaffolds. Initially the emphasis was on macroporous structures to facilitate mass transfer of vital molecules, developed using microspheres, salt leaches or gas foams.8, 22 Such micron-scale scaffolds however do not recapitulate the nano-scale dimensions of the ECM.23 In order to generate such nanometer-scale dimensions techniques such as electrospinning,24 molecular selfassembly,25 soft lithography,26 and phase separation27 have been employed.8 1.3 Physical features of the cell microenvironment The important physical properties to consider towards the cell’s microenvironment are substrate mechanics and surface topography.8 Mechanical properties of the tissue are dependent on their anatomical location. For example, the elastic modulus of the soft tissue of the brain (0.5 kPa) is low compared to intermediate muscles and skin (~ 10 kPa), and hard precalcified bone (>30 kPa).8, 28 One of the major limitations in recreating the native cellular environment is to reproduce the intricate physical features found in tissues which are viscoelastic with non-linear, anisotropic and heterogeneous mechanical properties. It is an important consideration while designing scaffolds because the optimal scaffold should completely replicate the replacement tissue.8 For example, the biological nanostructure of the heart influences its biochemical, electrical and mechanical functions and its ECM comprised of dense elastic fibrillar collagen, elastin bundles and proteoglycans.5, 29 It promotes mechanical coupling between cardiomyocytes resulting in aligned and anisotropic cell bundles promoting not only intercellular interactions but also with neighbouring capillaries and nerves.5 This bundled elongated cellular structure is critical for the function of the cardiac syncytium, enabling the muscles with exclusive rhythmic contraction and mechanical and electrical properties to facilitate the blood pumping function of the heart.5 Cardiomyocytes have been shown to adopt random morphology on flat surfaces as they lose their elongated morphology post 4 isolation. This is most likely because they lose the cues required to instruct their formation and interactions with their environment.5 Strategies adopted to address this problem of cell alignment in engineered cardiac tissue include applying mechanical stretching,30 interstitial fluid flow, electrical stimulation31 and microcontact printing,32 with limited success.5 Alternatively, a model cadiac tissue was developed by controlling the nanotopography of the scaffold that mimics the ECM matrix of the native myocardial tissue.5, 33 Polyethylene glycol hydrogel was fabricated to yield grooved arrays with ridge widths ranging from 15 to 800 nm which when cultured with cardiac cells resulted in their self-assembly aligned to the direction of the scaffold.5 It resulted in elongation of cadiac cells to form anisotrophic cell array which could result in rhythmic contraction necessary for normal heart function.33 Despite the progress made, the challenge of patterning such nano-arrays within 3D scaffolds towards anisotropically aligned tissues still remains.5 1.4 Design concepts and strategies Three-dimensional (3D) scaffold constructs have more demanding requirements for efficient cell motility compared to the 2-D monolayer cultures to achieve tissue uniformity and to avoid heterogeneous tissue growth.34 The supply of oxygen and other nutrients along with waste removal and cell motility reply on mass transfer properties of the scaffold.35 Considering the primary mechanism behind mass transport is diffusion, scaffolds need to be designed after carefully considering its diffusion characteristics.36 Cells in vivo reside within 100 µm of a capillary for efficient nutrient supply.35 Neotissue growth on engineering scaffolds has been reported to be preferentially limited to the peripheral regions (100-200 µm) of the scaffold because of the limited oxygen supply.35, 37, 38 This is still a major limitation in the engineering of thick 3D tissues because it limits the cellular distribution and bioavailability of nutrients especially in the interiors of the scaffold as the cells preferentially localise to the periphery. 35 Therefore scaffolds for tissues with low metabolic activity have been easier to synthesize. For example, cartilage, an 5 avascular tissue known to have low metabolic activity, has been successfully fabricated into a thick 3D construct.39 Similarly, other structural tissues known for their mechanical function such as blood vessels, heart valves, ligaments, tendons and skin may have lower demand for high oxygen and nutrient transport and therefore be easier to replicate.35, 40 1.4.1 Physical properties One of the main features required facilitating nutrient and oxygen transport is the pore size of the engineered construct. Pore size is critical as it influences tissue ingrowth and cellular adhesion especially in the internal surface area of the matrix. For example, small pores are impeded by the growing tissue preventing tissue ingrowth and ECM production. Alternatively, big pores lack cellular recognition of the surface topography preventing neovascularisation.35 Although the pore size is susceptible to change in an in vivo environment,35, 39 it has been demonstrated that optimal size varies with differing cell types and their respective architecture. For example, where 5 µm is optimal for neovascularisation, fibroblasts ingrowth demands 5-15 µm whereas hepatocytes require 20 µm, keratinocytes need 20-125 µm, and 100-700 µm size pores are required for bone regeneration.35, 39, 41, 42 Ma et al. studied the effects of pore size in a 3-D polyethylene terephthalate nonwoven fibrous matrix on long-term tissue development of human trophoblast ED27 cells in terms of cell morphology and spatial organisation.43 They showed that human trophoblast ED27 cells with average diameter of 14 µm were not able to bridge the gap between fibres with a pore diameter of 20 µm. Instead they formed large aggregates and started to differentiate in larger pores. On the contrary, cells rapidly bridged the gap on the matrix with low porosity, a pore diameter of around 15 µm and promoted cell spreading and proliferation. Further, the cell differentiation was inhibited, highlighting the importance of pore size on cellular behaviour and response, in terms of cell morphology, differentiation and proliferation, towards scaffold design features. In yet another similar study, canine microvascular epithelial cells formed a thin endothelial lining only when grown on scaffolds with average pore size of 90 µm, whereas vascular smooth muscle cells preferred the pore size of 6 107 µm in order to form uniform tissue.35 Dermal fibroblasts, on the other hand, did not respond to any pore size.35, 44 Similarly, higher pore size greater than 250 µm has been shown to promote angiogenesis with faster cellular ingrowth when compared to smaller pore sizes.35 Despite the elaborate evidence on the impact of pore size and their distribution on angiogenesis, inflammation and tissue infiltration in vivo, the optimal pore size and range is still fairly unpredictable, especially in the context of regenerating complex tissues with multiple different cell types45, 46.35 In addition to pore size, cell transport mechanisms like diffusion, attachment and migration is dependent on total porosity, pore interconnectivity and scaffold surface area.35, 41 In order to achieve uniform cellular penetration and ingrowth, scaffolds are required to be highly porous with open interconnected geometry and a large surface area: volume ratio for example, scaffolds with over 90% porosity have been shown to have optimal surface area for cellular-matrix interactions warranted for efficient diffusion.35, 41, 47, 48 Highly porous scaffolds, however, lack the mechanical integrity required for tissue regeneration.36, 49, 50 Therefore, a fine balance between porosity and mechanical strength need to be established for the development of an optimal replacement scaffold.35 1.4.2 Mechanical properties Studies determining the mechanical properties of various tissue types revealed that most tissues are heterogeneous, viscoelastic, nonlinear and anisotropic materials51.35 Scaffolds are required to be mechanically strong to carry out its function while withstanding the hydrostatic or pulsatile pressures encountered in vivo.35, 41, 42 The bulk properties of the constituents have been used to predict the mechanical properties of the designed scaffolds. In addition, there has also been a direct correlation between scaffold architecture and structural features such as pore distribution, fibre diameter and orientation and mechanical strength of the scaffold52.35 This is of particular interest in the case of scaffolds with high total porosity and low material content, for example in hydrogels and electrospun scaffolds. It has also been demonstrated that the cellular response, such as cell contractility, motility, adhesion, spreading and differentiation, can be influenced by 7 the physical properties of the scaffold like softness and hardness53-55.35 Engler et al. reported that matrix properties such as softness and elasticity has a profound influence on stem cell lineage and their commitment to a particular phenotype.55 It was shown that human mesenchymal stem cells display a more neurogenic phenotype on soft substrates (0.1-1 kPa), a myogenic phenotype on moderately hard matrices (8-17 kPa) and an osteogenic phenotype on hard substrates (25-40 kPa).35 In another study, soft substrates with Young’s moduli less than 1-1.5 kPa have been reported to enhance the differentiation of neural progenitor cells.56 This property of the substrates can be exploited to achieve tailored cell proliferation and differentiation56.35 1.4.3 Surface properties Cellular adhesion and growth has been shown to be correlated to the surface properties of scaffolds including topology and chemical characteristics.57, 58 Chemical properties correspond to the cellular interactions including adhesion and protein interactions at the material surface. Tamada and Ikada investigated 13 different polymeric surfaces for fibroblast adhesion, growth and collagen synthesis, with a range of different surface energies using goniometer (contact angle measurements).59 It was concluded that proliferation was independent of surface chemistry, whereas cell adhesion which relies on cell protein-substrate interactions was correlated to surface wettability. A positive correlation was observed between the surface charge and density of adsorbed proteins which corresponded to better cellular adhesion.35, 60 Furthermore, cellular behaviour was altered corresponding to the surface topography from angular edges, abrupt grooves or other surface indentations to smooth surfaces.61-64 A number of studies have reported the directional motility of cells along the fibres and ridges fabricated on the substrate surface which is explained by a phenomenon called “contact guidance”.35, 65 As highlighted above, an ideal scaffold is an amalgamation of chemical, mechanical and structural properties. It should be biodegradable but structurally strong, should be porous but also flexible. Therefore, an ideal scaffold should be multifunctional, incorporating cues for cellular processes like adhesion, extracellular matrix 8 production and migration, but also have potential applications such as drug delivery and multimodal imaging capabilities. 1.5 Skin, injury and wound healing Skin is the largest organ on the human body. Adult skin consists of two main tissue layers: a thin top layer called epidermis which is mainly composed of keratinocytes while the lower epidermal layer is made up of melanocytes, cells responsible for skin pigmentation and an underlying supporting layer called dermis which is mainly composed of collagen synthesising cells, fibroblasts (Figure 1).66, 67 The main role of epidermis is protection against microbes and to help regulate body temperature.68 The layer sandwiching these two layers is a 20 nm thick membrane called basement membrane. It is made up of hemidesmosomal structures and its main role is to mechanically stabilize the interaction between the epidermis and dermis.67 9 Figure 1: This diagram of human skin shows the two main layers of skin — the upper epidermal barrier layer and the lower, much thicker, dermis. Figure and caption taken from MacNeil 2007. 69 The main role of dermis which is 2-5 mm thick is to provide support to the overlaying epidermis.67 It provides considerable tensile strength and elasticity to skin mediated by the intertwined arrangement of collagen fibers, has specialized components and structures.67 Collagenous architecture is composed of varying amounts of interwoven elastin fibers, proteoglycans, fibronectin and other components.67, 68 To avoid infection, injury to the skin demands rapid intervention where the healing time is dependent on extent and depth of the injury. Epidermal lesions mostly heal within a week.66 Wound healing is a dynamic and interactive process involving synergistic interactions between blood cells, ECM, and parenchymal cells.70 Wound healing is accomplished by three successive but overlapping stages: inflammation, reepithelialisation and remodelling (Table 1). Failure or delays in healing can lead to significant scarring and potentially fatal sepsis and therefore interventions to promote healing are critical to reduce morbidity and mortality associated with extensive skin injuries. 10 Table 1: Phases of wound healing, major types of cells involved in each phase, and selected specific events. Taken from Falanga 2005.71 1.5.1 Haemostasis and inflammation An insult to the skin causes the disruption of blood vessels and extravasation of blood constituents. As the blood clots it re-establishes haemostasis and provides a provisional ECM (Figure 2).70 As the platelets aggregate they not only facilitate the formation of a hemostatic fibrin mesh but also secrete several mediators of wound healing, such as platelet derived growth factor, a chemoattractant that recruits and activate macrophages and fibroblasts.70 Coagulation of platelets generates numerous vasoactive mediators and chemotactic factors and activates complementary pathways by injured and activated parenchymal cells. This then results in the recruitment of inflammatory leukocytes to the injury site.70 11 Figure 2: A Cutaneous Wound Three Days after Injury. Growth factors thought to be necessary for cell movement into the wound are shown. TGF-β1, TGF-β2, and TGF-β3 denote transforming growth factor β1, β2, and β3, respectively; TGF-α transforming growth factor α; FGF fibroblast growth factor; VEGF vascular endothelial growth factor; PDGF, PDGF AB, and PDGF BB platelet-derived growth factor, platelet-derived growth factor AB, and platelet-derived growth factor BB, respectively; IGF insulin-like growth factor; and KGF keratinocyte growth factor. Reproduced with permission from A. J. Singer and R. A. F. Clark, N. Engl. J. Med., 1999, 341, 738-746. Copyright Massachusetts Medical Society. Next, neutrophils infiltrate the wounded area to initiate cleaning the injury site of foreign particles and bacteria, which are then extruded with the eschar or phagocytosed by macrophages.70 Chemoattractants such as fragments of ECM proteins, transforming growth factor β (TGFβ) and monocyte chemoattractant protein 1 (MCP1), attract monocytes to the injury site. These monocytes become activated macrophages that release growth factors and additional cytokines to initiate the formation of granulation tissue.70 Interaction of macrophages with the ECM through their integrin receptors, stimulate them to phagocytose microorganisms and fragments of ECM and simultaneously, stimulate monocytes to differentiate into pro-inflammatory or reparative macrophages 72 70 . This interaction also stimulates both monocytes and macrophages to express colony-stimulating factor 1, a prosurvival cytokine; tumor necrosis factor α, an inflammatory cytokine; and platelet12 derived growth factor, an important chemoattractant and mitogen for fibroblasts.70 Other key cytokines secreted by monocytes and macrophages include transforming growth factor α, interleukin-1, transforming growth factor β, and insulin like growth factor 173.70 Macrophages are an integral part of the wound healing process critical to wound repair, with macrophage deficient animals shown to have markedly defective wound healing.74, 75 Figure 3: Schematic diagram of wound reepithelialization models. (a) Basal KCs at the leading edge of the wound that are firmly attached to surrounding basal and suprabasal KCs actively migrate to close the wound in the “sliding” model of reepithelialization. Arrow indicates direction of movement of basal KC. (b) With the basal KCs firmly attached to the BMZ, suprabasal KCs roll onto the wound matrix in the “rolling” model of reepithelialization. Arrow indicates movement of suprabasal KCs as they tumble over basal KCs. Arrowheads indicate initial cut edge. Figure and caption taken from Usui 2005.76 1.5.2 Reepithelialisation Two mechanisms behind reepithelialisation have been proposed in the literature: the “rolling” model and the “sliding” model (Figure 3). The “rolling” model postulates that basal keratinocytes remain strongly attached to the basement membrane, while suprabasal keratinocytes at the wound margin are activated to roll over into the wound site77, 78.76 The “sliding” mechanism on the other hand, postulates that basal keratinocytes are the principal cells mediating the migration and wound closure. Both basal and suprabasal keratinocytes remain strongly attached to the leading edge and basal keratinocytes are passively dragged along as a sheet79, 80.76 A variant is also proposed where a large population of suprabasal cells migrate out of the wound (Figure 4).76 13 Figure 4: Schematic diagram of a new model of reepithelialization. Both basal and suprabasal KCs are activated to respond to the complex microenvironment created by a wound. Suprabasal KCs undergo dramatic changes in response to injury and are possibly the primary source of cells available for wound closure. Figure and caption taken from Usui 2005.76 Reepithelialisation is initiated spontaneously after injury.70 Epidermal cells from the skin appendages initiate the cleaning process by removing the clotted blood and damaged stroma.70 Simultaneously, cells alter their phenotype which includes retraction of intracellular tonofilaments;78 dissolution of most intercellular desmosomes, important for physical connections between the cells; and formation of peripheral cytoplasmic actin filaments, crucial for cell motility81, 82.70 Furthermore, the loss of basement membrane and hemidesmosomal links disrupts the interactions between juxtaposed epidermal and dermal cells, allowing the lateral movement of epidermal cells.70 Epidermal cells interact with a variety of ECM proteins via their integrin receptors. These include proteins such as fibronectin and vitronectin which are interspersed with stromal collagen I at the wound margin and interwoven with the fibrin clot in the wound space83-85.70 The epidermal cells start to migrate and dissect the wound to separate desiccated eschar from viable cells.70 Collagenase as well as the activation of plasmin by plasminogen activator is produced by migrating epidermal cells in order to degrade the ECM required for their migration between the collagenous dermis and the fibrin eschar86.70 Plasminogen activator is also 14 known to activate collagenase (matrix metalloproteinase I) which propels the degradation of collagen and ECM proteins. Figure 5: A Cutaneous Wound Five Days after Injury. Blood vessels are seen sprouting into the fibrin clot as epidermal cells resurface the wound. Proteinases thought to be necessary for cell movement are shown. The abbreviation u-PA denotes urokinase-type plasminogen activator; MMP-1, 2, 3, and 13 matrix metalloproteinases 1, 2, 3, and 13 (collagenase 1, gelatinase A, stromelysin 1, and collagenase 3, respectively); and t-PA tissue plasminogen activator. Reproduced with permission from A. J. Singer and R. A. F. Clark, N. Engl. J. Med., 1999, 341, 738-746. Copyright Massachusetts Medical Society. Epidermal cell proliferation begin few days after injury.70 Local release of growth factors, such as epidermal growth factor, transforming growth factor α and keratinocyte growth factor;87 and absence of the neighbouring cells on the wound margin are thought to be the principal reasons behind this epidermal response.70 The reepithelialisation results in the synthesis of basement membrane proteins laid down in an orderly sequence from the margins of the wound inwards, mimicking the original architecture in a zipper like fashion88.70 Epidermal cells return to their normal phenotype, establishing juxtaposition to the re-established basement membrane and underlying dermis once the epidermal barrier has been restored.70 This also acts as a trigger for next step of inflammation and matrix remodelling. 15 1.5.3 Formation of granulation tissue Four days post injury, granulation tissue starts to invade the wound space (Figure 5). Macrophages, fibroblasts and blood vessels simultaneously start to infiltrate into the wound space.70 Angiogenesis and fibroplasia is constantly supplemented by the growth factor secreted by macrophages; while fibroblasts lay down ECM necessary to support cell ingrowth and blood vessels carry oxygen and nutrients crucial to sustain cell metabolism.70 Fibroblasts, nurtured by growth factors, especially platelet-derived growth factor and transforming growth factor β1 (TGFβ1), in addition to the ECM proteins, begin to proliferate and start to express integrin receptors for their migration into the wound.70 Platelet derived growth factor, along with many other similar factors, have been shown to accelerate healing of chronic pressure sores and diabetic ulcers when applied exogenously to these wounds. 89, 90 Newly synthesised provisional ECM promotes granulation tissue formation by providing the conduit for cell migration88.70 It has been postulated that the fibronectin architecture and fibronectin specific integrin receptors on fibroblasts control the rate of granulation tissue formation.70 These fibroblasts are responsible for the synthesis, deposition, and remodelling of the ECM. Cell migration into the blood clot of crosslinked fibrin, or tightly woven ECM requires fibroblast-derived enzyme proteases like plasminogen activator, collagenases, gelatinase A, and stromelysin that can cleave the matrix to facilitate their migration91.70 Post migration into the wound, fibroblasts initiates ECM synthesis.92, 93 It is believed that transforming growth factor β1 stimulates the replacement of the provisional ECM with a collagenous scar-type matrix92, 93.70 Once an abundant collagen matrix has been deposited, fibroblasts stop their excessive collagen production triggering the replacement of granulation tissue by a relatively acellular scar.70 16 1.6 Role of TGFβ in fibrosis and scarring The TGFβ superfamily contains over 30 proteins found in both vertebrates and invertebrates and encompassing a wide range of functions throughout the lifetime of the animal.94 The three mammalian TGFβ gene isoforms are TGFβ1, 2 and 3 which share 60-80% sequence homology and are encoded by three different genes.95, 96 The three mammalian isoforms of TGFβ are expressed in a cell specific and developmentally regulated manner.97, 98 For example, TGFβ1 and TGFβ3 are expressed during the morphogenesis stage of early development followed by TGFβ2 which is expressed later in mature and differentiating epithelium.98 All three isoforms are highly conserved in mammals, but differ in their binding affinity towards TGFβ receptors.98, 99 Altered and different phenotypes were resulted as a result of the deletion of individual isoforms in mice.98-100 TGFβ1-null mice were mostly embryonically lethal due to abnormal development of the yolk sac, and those that did survive developed multi-organ autoimmunity and multi-focal inflammation disease and died within 3 weeks of birth101.102 TGFβ2-null mice showed perinatal mortality due to cyanotic heart disease and various other developmental defects103.102 TGFβ3-null mice die due to cranial bone defects especially cleft palate104.102 Figure 6: Schematic representation of latent TGF-β. The putative transglutaminase-mediated covalent bonding between the N-terminus of LTBP and the ECM is indicated by a question mark (?). Disulfide bonds between LAP monomers, between TGF-β monomers and between LAP and LTBP are indicted by thin lines. Carbohydrate residues on LTBP are not shown. Abbreviations are in the text. Figure reproduced, and caption taken from Munger 1997.105 17 Newly synthesised TGFβ is secreted in an inactive latent form. The non-covalent association of TGFβ with a latency associated peptide (LAP) is the reason behind the latency (Figure 6). LAP is a homodimer formed from the propeptide region of TGFβ.105 This association of LAP and TGFβ is termed as latent TGFβ (LTGFβ). To exert its function, TGFβ needs to be cleaved from its latent complex with LAP.105 Two of the N-linked carbohydrate residues contain mannose-6-phosphate (M6P) groups. There are 3 cysteines in LAP-1 which form interchain disulfide bonds to dimerise LAP monomers.105 This cysteine interaction is critical to maintain latency because their replacement with serines at positions 223 and 225 of LAP-1results in the secretion of active TGFβ1.106 LAP is also known to form a disulfide linkage to another protein called latent TGFβ binding protein (LTBP).105 LTBP link latent TGFβ to the ECM. TGFβ has three high-affinity membrane receptors where type I and II are transmembrane serine/threonine kinases that coordinate to facilitate each other’s signalling,98, 107-109 while type III receptor, a transmembrane proteoglycan, has no specific signalling function because of its highly conserved cytoplasmic domain.98, 110, 111 TGFβ interact with these different receptors to mediate specific functions. For example, interaction with receptor type I promotes secretion and deposition of ECM, while interaction with type II receptor mediates cell growth and proliferation. In order to activate the type II receptor, TGFβ binds either to a type III receptor which then facilitates its binding to type II receptor or it can bind directly to a type II receptor.112 Upon activation by TGFβ, type II receptors recruit, bind and transphosphorylate type I receptors, thereby promoting their protein kinase activity and initiating its downstream signalling cascade.112 TGFβ signalling within the cells is mediated by the Smad family of transcriptional activators and has been extensively reviewed elsewhere.113-118 18 1.7 Mechanisms of activation of extracellular TGFβ Downstream signalling of TGFβ is regulated by local activation of the LTGFβ complex in vivo.102 There are various mechanisms of LTGFβ activation which have been widely studied.2, 102, 119 M6P/IGFII receptor mediated activation of LTGFβ has been illustrated as one of the most critical mechanism of activation. One mechanism of activation involves integrins. Integrins are transmembrane receptors that conjugate the cell cytoskeleton to the ECM and are critical for cell adhesion, proliferation, migration and differentiation120-122.102 Integrins are made up of α and β subunits.102 αv integrins recognise and bind to both TGFβ1 and LAP via their RGD sequence. This RGD mediated interaction between αvβ6 integrin and LAP-1 cause the activation of LTGFβ1 by altering its conformation.102 However, β6 integrin is required to interact with the actin cytoskeleton of the cell to ascertain LTGFβ1 activation.102, 123 This alteration in the LTGFβ1 structural conformation produces mature TGFβ which can interact with the TGFβ II receptor and therefore allow it to mediate its function124.102 Thrombospondin-1 (TSP-1), a 300 kDa protein found in the α-granules in platelets and ECM has also been shown to activate both small LTGFβ (LAP-TGF-β complex) and large LTGFβ (LTBP-LAP-TGF-β) complexes in a non-proteolytic mechanism both in vitro and in vivo125-127.102 It has been postulated that the TSP-1 interaction with LAP induces a conformational change in relation to mature TGFβ thereby unveiling it to the TGFβ receptor recognition site and facilitating TGFβ binding to its receptor128.102 The proteases such as plasmin, matrix metalloproteinase (MMP) 2 and 9 have also been reported to induce LTGFβ activation in vitro129, 130 102 . Plasmin conditioned media has also been shown to generate active TGFβ in fibroblasts and Chinese hamster ovary cultures131.102 Alternatively, the activation of LTGFβ can be inhibited in vitro by plasmin inhibitors or prolonged by neutralising antibody to plasminogen activator inhibitor-1 as demonstrated in co-culture studies of endothelial cells and pericytes or smooth muscle cells132.102 19 Another mode of in vivo activation is via the binding of the M6P residue present on the latency associated peptide (LAP) to the M6P/IGFII receptor. This invokes a conformational change in TGFβ bound to M6P/IGFII receptor, thus allowing the proteolytic cleavage of active TGFβ from its latent complex. In vitro activation of TGFβ has been investigated in detail in activated macrophage cultures and in cocultures of endothelial and smooth muscle cells where latent TGFβ is activated by a complex process involving recognition and binding of the latent form to the M6P/IGFII receptor and the concerted action of both transglutaminase and serine protease plasminogen/plasmin.133, 134 Once activated TGFβ suppresses the inflammatory response and promotes the formation of granulation tissue. TGFβ promotes keratinocytes migration by upregulating fibronectin expression and various integrins necessary for keratinocyte adhesion, whilst simultaneously inhibiting keratinocyte proliferation thereby regulating reepithelialisation stage of wound healing.135 It has been reported that overexpression of TGFβ1 in the epidermis results in delayed re-epithelialisation response in transgenic mice model.136-138 Despite having an inhibitory effect on the proliferation of endothelial cells, TGFβ has been reported to promote angiogenesis in vivo.139, 140 In the dermis, TGFβ activates fibroblasts which then produce higher levels of matrix molecules including collagen, fibronectin and glycosaminoglycans (GAG), matrixdegrading proteases and protease inhibitors thereby resulting in an increase in the production of matrix proteins and decreasing their proteolysis.141 Whilst these changes are critical for rapid wound repair, excessive and/or extended matrix deposition and excess TGFβ stimulation can lead to poor healing and scar outcomes. 1.8 Exogenous mannose-6-phosphate act as an antagonist of TGFβ1 activation Purchio et al., used radiolabeled [32P] glycopeptide and demonstrated the presence of endogenous M6P as a carbohydrate unit on LAP (TGFβ precursor).142 Binding 20 studies demonstrated that TGFβ precursor has high binding affinity towards purified M6P/IGFII receptor which can be hindered by exogenous M6P.142Miyazono and Heldin studied the importance of carbohydrate in TGFβ1 activation and demonstrated that interaction between endogenous M6P present on LTGFβ1 precursor and M6P/IGFII receptor can be result in TGFβ1 activation. In a model to study potential anti-fibrotic agents Gosiewska et al. used macrophage-based system for TGFβ1 activation and demonstrated that exogenous M6P inhibits the activation of LTGFβ.143 Bates et al., used a rabbit flexor tendon in vitro and in vivo models to investigate the role of exogenous decorin and M6P in TGFβ activation. It was concluded that both decorin and M6P inhibits the stimulatory effects of TGFβ on collagen production while even a single low intraoperative dose of M6P significantly improved the range of motion of the operated tendon.144 Xia et al., studied inhibitory potential of M6P in three different cell types; sheath fibroblasts, epitenon tenocytes and endotenon tenocytes from rabbit flexor tendon post TGFβ stimulation. They reported that TGFβ induced collagen production was significantly downregulated by the addition of exogenous M6P in a dose dependent manner in all three cell types.145 It is clear that exogenous M6P is a potent inhibitor of TGFβ signalling in multiple systems. 21 Figure 7: Chemical structure of a) mannose-6-phosphate showing the carbon numbers, b) an example of isosteric phosphonate analogue and c) an example of non-isosteric analogue. Image taken from Vidal 2000.146 1.9 Structural requirements for mannose-6-phosphate recognition to M6P/IGFII receptor Tong et al., first elucidated the structural requirements of M6P recognition by both the cation dependent mannose-6-phosphate receptor (CD/MPR) and cation independent mannose-6-phosphate receptor (CI/MPR, also known as M6P/IGFII receptor).147, 148 They investigated various ligands like M6P, pentamannose phosphate, β-galactosides and a high mannose oligosaccharide with two phosphomonoesters, and inferred that certain structural features are determinant in binding to M6P/IGFII receptor such as: a) The hydroxyl group at the C2-position (C2-OH) of the pyranose ring must be axial (Figure 7a) to allow hydrogen bonding interaction between hydroxyl group and Gln-348 and Arg-391 amino acids located on β-strands 3 and 7 of the receptor.149 22 Similar strong interaction was observed in case of fructose-1-phosphate (F1P) while glucose-6-phosphate, with C2-OH in equatorial position is weakly recognised (2epimer of M6P).150 b) In addition to the orientation, C2-OH is also inherently required as its absence results in diminished interaction with the receptor like in the case of 2-deoxy glucose-6-phosphate (2dG6P).147 c) The substitution at anomeric centre does not influence the interaction of the analogue to the receptor as F1P is recognised in similar propensity as M6P. Furthermore, replacement of the hydroxyl group at anomeric position with a paranitrophenoxy group, slightly improves its recognition towards the M6P/IGFII receptor. It is due to the lipophilic interactions between the aromatic moiety of the ligand and the binding pocket of the receptor.150 d) Substitution at 5-position of the pyranoside ring does not influence the interaction of analogue to the receptor.147 e) The distance between the negative charge (phosphate group) and C5 of the pyranose ring should be four atoms.146 f) Only a single negative charge is necessary for the binding to M6P/IGFII receptor since isosteric M6-phosphonate analogues (example in Figure 7b) are very well recognised compared to non-isosteric M6-phosphonate analogues (example in Figure 7c) which are very weakly recognised.151 1.10 Mannose-6-phosphate analogues Due to the importance of M6P in inhibiting TGFβ and its potential to therefore ameliorate scarring M6P has been studied in a double blind, placebo controlled, randomised, phase 2 efficacy clinical trial (Renovo UK, Juvidex®).152, 153 The study was carried out in almost 200 male and female subjects administering different doses of the drug to split thickness skin graft donor sites. Intradermal delivery of the drug was shown to significantly accelerate the rate of wound healing. However, no significant improvement in the extent of scar formation was observed.154 This is in part due to the subjective measurement and difficulties in measuring scar improvement. However, the metabolic vulnerability of M6P against phosphatases 23 has also been postulated to be a major limitation in its clinical translation. Therefore attempts have been made to develop M6P analogues with greater stability against enzymatic degradation. Vidil et al., reported the synthesis of M6P analogues with the aim of developing analogues with enhanced affinity towards the M6P/IGFII receptor.155 It was postulated that replacing the P-O bond at C6 position with P-C bond, as is the case in phosphonate analogues, would enhance the stability of the analogues towards hydrolases. They showed that the isosteric analogue of M6-phosphonate binds to the M6P/IGFII receptor with higher affinity compared to the non-isosteric derivative (with shorter chain length (1 atom) between phosphate group and pyranose ring) and with similar affinity to M6P.155 Berkowitz et al., reported the synthesis of three mono and bivalent ligands bearing M6P surrogates (malonyl ether, malonate, and phosphonate) and studied their binding affinity towards the M6P/IGFII receptor. These surrogates were hydrolase resistant phosphates (with a methylene bridge at C6 position bridging the pyranose ring and respective functional groups) and mimic the M6P locked in the αconfiguration.156 It was concluded that phosphonate analogues have greater binding affinity than malonyl ether and malonate analogues.156 In another example, Jeanjean et al., developed two sulfonate (M6S) and one unsaturated phosphonate analogue of M6P in an attempt to study their stability in human serum and their binding affinity towards the M6P/IGFII receptor.157 The two isosteric sulfonated analogues were synthesised with the view that higher chemical stability of conjugated (unsaturated) analogues compared to M6S and unconjugated (saturated) analogue would have elevated binding efficiency to the receptor. Conjugation studies were carried out to decipher the binding of these analogues to the M6P/IGFII receptor. Binding studies confirmed the vulnerability of M6P in human serum which dropped 5 fold while the binding affinity of the three new analogues remained intact.157 24 Figure 8: Chemical structure of carboxylate analogues of M6P. Image taken from Jeanjean 2006. 158 In a similar study, four new carboxylate analogues of M6P were investigated for their receptor binding affinity and serum stability (Figure 8).158 Importance of a negative charge on the M6P analogue was emphasised by the receptor binding assay based on the ligand dependent elution of the receptor from PMP affinity columns.159 It was demonstrated that unsaturated isosteric carboxylate analogue had similar binding affinity as M6P while non-isosteric carboxylate analogues had slightly weaker affinity to the receptor.158 This difference in binding efficiency was explained by structural proximity of the double bond, which in the case of unsaturated isosteric analogue, could stabilise the analogue and its geometry would promote its interaction with M6P binding sites on the receptor, in comparison to the non-isosteric (saturated) analogue. The three analogues with greater binding affinity were also shown to have prolonged stability against hydrolases. Analogues were shown to retain their recognition potential even after 2 days incubation in human serum, while 6 hour incubation was shown to reduce M6P binding affinity by 32%. Christensen et al., synthesised a series of glycopeptide derivatives of M6P, containing two 6′-O-phosphorylated mannose disaccharides linked either α(l → 2) or α(l → 6) and 3-5 amino acids, in order to study the influence of structural variation in mannose disaccharides on their binding affinity towards M6P/IGFII receptor.160 The analogue comprised of two 6’-O-phosphorylated α(1→2)-linked mannose disaccharide showed higher binding affinity compared to α(1→6)- linked analogue,160, 161 It was concluded that analogue receptor affinity was dependent on terminal phosphorylated mannose disaccharides and was found independent of the 25 variation in invariant saccharide core from tri-, tetra- to penta-peptides. However, removal of aminobenzoyl group bound to the lysine tail in one of the analogue was shown to significantly diminish their receptor binding affinity.159 This demonstrates that structural rigidity is required to maintain strong affinity of the peptide towards the receptor160.159 It can be concluded that there are a number of key structural requirements to be considered in M6P analogue design. These structural constraints are important for binding affinity and specificity and also for stability in vivo. However, despite the design of a number of analogues to date with promising data suggesting increased stability and binding affinity, no clinical treatment targeted at the M6P inhibition of TGFβ signalling currently exists. Therefore it is likely that other delivery considerations may also be important. 26 Figure 9: An illustrated representation of raised dermal scar types, as commonly observed after a midsternal incision post‐cardiac surgery. (a) Fine line scar, (b) Hypertrophic scar, (c) Intermediate raised dermal scar, (d) Keloid scar. Image and caption taken from Sidgwick 2012.162 1.11 Scarring There are a broad range of scar outcomes found in humans. Clinically scars are quantified using scar assessment scales.163-165 Normal scars are usually fine line scars which can be extended with mechanical stretch applied on the scar (Figure 9a).2, 162 Hypertrophic scars are raised dermal scars which are the result of aggressive proliferation leading to excessive healing and matrix deposition162 (Figure 9b). Contracted scars as the name suggests result in contracture of the granulation scar tissue around the joints following burn injury or surgery162 (Figure 9c). Finally, keloid scars are similar to hypertrophic scars but crucially extend 27 beyond the boundaries of the initial wound and are commonly considered to be more similar to a benign tumour than a normal scar (Figure 9d).166-169 1.12 Current treatments for skin wound healing The etiology of skin damage is diverse, from genetic disorders such as epidermolysis bullosa through to acute trauma, tumours, chronic wounds (ulcers) and even surgical intervention.170 One of the main factors behind a major loss of skin is burn injury.171 It is estimated that each year over 300,000 people die from burn-related injuries worldwide.172 Many more people suffer from burn-related disabilities and disfigurements.172 With the advent and development of new aggressive surgical interventions the rate of patient mortality due to burn related injuries has receded considerably in recent years. This has had a profound effect on patients, with the increase in survival being mirrored by an increase in poor extensive scar outcomes. Treatment of skin wounds, including burn injury, is dependent on the extent of the injury. While superficial or incisional wounds heal with little or no intervention, deeper wounds often require clinical intervention. The current gold standard for replacement skin is via a split-skin graft - an autograft where skin is harvested from an uninjured donor site on the patient, meshed to cover a larger area and patched onto the injured site of the same patient. This approach is utilised preferentially in full thickness and deep partial thickness wounds where both epidermal and some of the dermal layer of the skin is harvested from an uninjured site and grafted on the damaged site.173 The major limitation to this approach is that as the surface area of the injury increases the availability of donor tissue decreases. Additionally, tissue from different body sites is not equivalent and this can lead to poor aesthetic outcomes (for example pigmentation). Finally, the use of donor sites creates an additional wound which needs to be healed and can lead to significant donor site morbidity.173, 174 In order to address these problems it is essential to develop new 28 interventions. Synthetic skin substitutes and engineered replacements are employed to restrict the burden of disease by reducing the need for graft donor sites.175 1.13 Cell based therapies Suspended keratinocytes cell based therapies are emerging as a leading therapeutic strategy, aimed at utilising cells as replacement therapy to repair severely damaged tissues176.173 In 1975, the first subcultures of human keratinocytes was developed and manipulated to obtain epithelial sheets for subsequent grafting177.173 Their first major clinical application was reported in 1981 where cultured autologous epithelial sheets were used for the treatment of extensive third degree burns178.173 The time to culture these sheets, approximately 3-5 weeks, was a major limitation leaving patients susceptible for prolonged periods after the injury179-181.173 Some of the other limitations included high production costs,182, fragility, variable engraftment rates, difficult handling, storage and preservation of viable sheets.173, 183-186 It was the development of an automated membrane bioreactor which brought down the delay time to 2 weeks187.173 An alternative innovative approach was later developed, to further reduce fabrication time, where cells extracted from a skin biopsy were not cultured but rather directly sprayed onto the lesions (ReCell®; Avita Medical, Australia and Spray®XP; Grace, USA)183, 188, 189.3 The biggest advantage in using cultured keratinocytes, or non-cultured sprayed cells, was the coverage of larger surface areas from relatively small biopsies in the range of 2-5 cm2 from an uninjured site of the patient. It is unsuitable to be used as a sole treatment for patients with deep dermal or full thickness injuries, as it predominantly caters to the epidermal cell population while deeper injuries inevitably also demand dermal support for efficient healing response.183 Nevertheless, autologous cells are in widespread use today for the treatment of partial thickness burns. More recent research has involved further developments of cell-based treatments, with the potential to enhance these treatments through novel delivery mechanisms or the addition of growth factors or other biologicals to promote better healing.3, 29 190-195 However, these advanced cell therapies remain in the research stage and are not currently in widespread use for burn treatment. 1.14 Matrix based therapies 1.14.1 Amniotic membrane Since 1910, allogenic amnion has been used as a biological wound dressing.173 It draws parallel to normal human skin allografts and was considered the most effective dressing ever used, especially in preserving the healthy excised wound bed and providing protection from pathogenic contamination196-199.173 Amniotic membrane is derived from the innermost layer of the fetal membrane, it is a semitransparent tissue consisting of avascular stroma, thick continuous basement membrane with a full complement of collagen IV and V, laminin and also contains several protease inhibitors200.173 Advantages of human amniotic membranes include promotion of epithelial regeneration by reduction in loss of proteins, electrolytes and fluids, drug delivery and reducing the need and frequency of dressing changes and pain associated with the process.173 The key disadvantages are its fragility, technically it is difficult to handle, the risk of contamination and transmission of disease and it is inadequate in deep dermal injuries where it disintegrates before healing occurs201, 202.173 1.14.2 Human cadaver derived allografts and xenografts Human cadaver allograft skin (HCAS) can be used as a temporary dressing in cases where the availability of patient donor sites are limited203-205.173 Similar to amniotic membranes, HCAS may also be used as wound dressing to cover widely meshed autografts in massive burns.173, 206 There are serious problems associated with HCAS use including limited supply, variable quality, possible contamination and immune rejection.204, 206, 207 On the other hand, xenografts are derived from various animal species including rabbit, dog and pig and have been used as a temporary replacement 30 skin for a long time.173 The major limitation in their use include immunologic disparities and predetermined rejection over time, despite the obvious benefits of biologically active dermal matrix they provide.203 These alternatives continue to be considered strictly as only temporary wound cover with therapeutic strategies to promote healing focused on living autograft cells and tissue. Figure 10: (a), Cells are isolated from the patient and may be cultivated (b) in vitro on two-dimensional surfaces for efficient expansion. (c), Next, the cells are seeded in porous scaffolds together with growth factors, small molecules, and micro- and/or nanoparticles. The scaffolds serve as a mechanical support and a shapedetermining material, and their porous nature provides high mass transfer and waste removal. (d), The cell constructs are further cultivated in bioreactors to provide optimal conditions for organization into a functioning tissue. (e), Once a functioning tissue has been successfully engineered, the construct is transplanted on the defect to restore function. Image and caption taken from Dvir 2011. 5 31 1.14.3 Tissue engineered skin In this approach skin cells are expanded in the laboratory before their application in wound healing (Figure 10).69 In addition, for deeper injuries a number of dermal scaffolds have also been developed that can be used in conjunction with the cell based approaches to restore both dermal and epidermal layers. All tissue engineered materials needs to have some essential characteristics including low disease risk, promotion of healing, possess physical properties similar to normal skin, it must attach well to the wound bed, be compatible to the developing new vasculature, non-immunogenic and be convenient to use.69, 171 To replace the dermal layer, predominantly collagen based matrices have been used, with Integra the first demonstration of a bilayered and biocompatible dermal scaffold used in the treatment of extensive burn injury to successfully induce the synthesis of neodermis.141, 208 Epidermal tissue grafts consist of in vitro differentiated keratinocytes forming a stratified epidermal layer.209 Keratinocytes and fibroblasts are cultured on various biocompatible substrates as implantable scaffolds including both natural and synthetic materials.209 Epicel® (Genzyme Corporation Offices, Cambridge, MA) is the first commercialized epidermal autograft which is composed of cultured keratinocytes extracted from 2 x 6 cm biopsy taken from the patient, which are expanded in the laboratory and then cocultured in the presence of proliferation arrested murine fibroblasts to form grafting sheets.209, 210 It is a front line treatment for full thickness burns in US and Europe.209 The clinical outcome of epidermal transplants fall short in cases with inadequate dermal tissue, which is required to provide support for grafted epidermal sheets211. Alternatively, biocompatible and biodegradable materials, such as benzyl-esterified analogues of hyaluronic acid (Hyalograft 3D, Fidia Advanced Biopolymers, Padua, Italy) and polyglycolic acid (Dermagraft, Shire Regenerative Medicine, San Diego, CA) have been used.209 Other examples include Transcyte (formerly Dermagraft-TC, Shire Regenerative Medicine) which is made up of porcine dermal collagen coated nylon mesh seeded with neonatal human fibroblast and externally supported by a silicone membrane212,209 Dermagraft (Shire Regenerative Medicine), utilise a biodegradable 32 scaffold coated with an allogenic human-derived dermal fibroblasts culture, and capable of producing several growth factors to stimulate angiogenesis, tissue growth and reepithelialisation from the wound edge even after cryopreservation and subsequent thawing213,209 and Apligraf (Organogenesis, Canton, MA) is an example of a clinically approved allogenic dermoepidermal product (combinatorial therapy), composed of cultured keratinocytes and a fibroblast dermal layer on collagen I matrix, for the treatment of venous and diabetic ulcers214.209 Clinical trials revealed a rate of healing of 63% in cases of ulcers as compared to 49% for compression therapy and a vast improvement in healing times from 181 to 61 days under Apligraf.215 Guo et al. developed a bilayer dermal equivalent by loading plasmid DNA (pDNA) encoding vascular endothelial growth factor-165 (VEGF165)/N,N,N-trimethyl chitosan chloride into collagen-chitosan/silicon membrane scaffold and tested for regenerative properties in a porcine full-thickness wound model216.3 Full-thickness burn wounds in the pig model when treated with N,N,Ntrimethyl chitosan chloride/pDNA-VEGF dermal substitute showed the best response towards dermal regeneration with highest density of newly-formed and mature vessels compared to a blank scaffold and scaffolds loaded with naked pDNA-VEGF and TMC/pDNA-eGFP.216 In another study, wound healing potential of five acellular dermal skin substitutes (Integra®, Integra Life Sciences, NJ, USA; ProDerm®, UDL Laboratories Inc., IL, USA; Renoskin®, Symatese, Ivry-le-Temple, France; Matriderm® 2 mm, Ideal Medical Solutions, Wallington, UK; and Hyalomatrix® PA, Addmedica, Paris, France) were investigated in a porcine fullthickness wound model in a two-step process217.3 In the first step, skin substitutes were implanted followed by the reconstruction of the epidermis using an autologous split-thickness skin graft or cultured epithelial autograft (after 21 days, current clinical treatment regime).3 Significant differences between the skin substitutes in terms of dermis incorporation and early wound contraction was observed. However, no significant long-term differences were in scar qualities between different dermal substitutes and the control group.217 Dermoepidermal substitutes mimic both epidermal and dermal layers of the skin and are considered the most advanced substitutes available in the clinic.3 They made up of both keratinocytes and fibroblasts incorporated into the scaffold which can provide temporary cover, promote simultaneous dermal and epidermal regeneration and also resemble the normal skin structure.3 The main limitations, however, are the high production costs 33 and their limited efficiency in terms of permanent wound closure due to allogeneic cell rejection, as well as limited evidence that the tissue substitutes ultimately improve the final outcome170.3 1.15 Nanofibrous scaffolds for skin tissue engineering Tissue engineering (TE) and regenerative medicine is the amalgamation of cellular components of a living tissue and functional biomaterials to develop functional living tissue of sufficient size for clinical translation.218 Recently, cell patterning, migration, proliferation and differentiation have been the point of focus for both TE and regenerative medicine.218 Development of scaffold as ECM substitutes are aimed to promote cells proliferation and differentiation for skin tissue regeneration in vivo.218 Cells respond to ECM signalling in a multistep process which is initiated by cell receptor-ECM ligand interactions, followed by the sequestration of growth factors by ECM, spatial cues and the mechanical force transduction219.218 Nanofabrication techniques have been employed to develop complex multifunctional porous, nanometer-sized fibrous scaffolds with surface morphology that can determine and influence cell fate, regulate the expression of specific proteins and encourage cell-specific scaffold remodelling.220.218 Techniques including nanoscale surface pattern fabrication such as lithography, electrospinning and self-assembly has been explored to fabricate scaffold topography down to nanoscale221.218 They can incorporate a myriad biological cues such as growth factors, angiogenic factors, cell surface receptors and spatial cues.218 These approaches can be used to influence cell proliferation, migration, differentiation and 3D organisation including ingrowth.3, 218 Electrospinning has gained considerable amount of interest because of its operational simplicity, versatility, ability to process variety of materials and propensity in producing nanofibers mimicking the ECM, for the fabrication of nanoscaffolds for skin regeneration.3, 218 Electrospinning utilise electric field to produce a highly impermeable, non-woven matrix of sub-micron to nanoscale fibers by pushing a highly charged polymer liquid jet through a small diameter nozzle of a syringe24, 222.223 Conventionally, a Taylor 34 cone is formed when electrostatic charging of the fluid pulls the polymer solution at the tip of the nozzle, ejecting a single fluid jet from the apex of the cone. 223 This polymer jet accelerates and thins in the electric field to form a nanofibrous scaffold. Electrospun scaffolds with different morphologies from aligned to random have been fabricated with fiber diameters down to the 100 nm range with various pore sizes224.209 Electrospun scaffold with high porosity and large surface area to volume ratio facilitate efficient mass transfer into the 3D structures of the scaffold.209 Nanofibrous scaffolds serve as a combinatorial approach towards regeneration of various tissues such as skin, bone, cartilage, vascular and neural tissues and subsequent drug delivery platform.209 Figure 11: Various fibrous structures fabricated via an electrospinning technique. (a) Core-shell structure, (b) random/align directional fiber and (c) micro-/nano-sized fiber. Image and caption taken from Rim 2013. 225 Nanofibers can be electrospun in various patterns including random, aligned and core-shell (Figure 11).209 Each type has their own advantages; for example random 35 fibers are used in tissue engineering approaches, while aligned ones have been explored in solar cells and shown to promote their efficiency225.209 In an another approach, emulsion electrospinning is used to fabricate polymer fibers with an integral core-sheath structure, promoting loading efficiency of the cargo, avoiding the initial burst release and protect and help maintain bioactivity of the loaded protein by avoiding the exposure of the cargo to the degrading physiological environment226.209 These fibrous scaffolds can be further functionalised using surface modification techniques such as plasma treatment, the wet chemical method, surface graft polymerization, co-electrospinning, and the immobilization of bioactive ligands227.209 In order to achieve controlled drug release the payload like drugs, enzymes and cytokines are physically or chemically immobilised within the fibers228.209 As a dermal substitute, highly porous 3D collagen scaffolds were fabricated using a electrospinning and were evaluated in a keratinocyte/fibroblast co-culture in vitro model for skin tissue engineering229.209 It resulted in well dispersed fibroblast ingrowth, while keratinocytes migrated through the pore structure and differentiated on the scaffold surface to form a stratum corneum and a 3D dispensed scaffold similar to normal skin.229 A well interconnected porous network is required in 3D scaffolds to achieve guided cell adhesion, cell growth favouring tissue development, cell migration and subsequent transport of solutes.209 This highlights the importance of mechanical properties of the engineered electrospun scaffolds towards their functions.209 Parenteau-Bareil et al crosslinked collagen, derived from various sources, with chitosan to produce ECM mimicking tissue engineering scaffolds230.209 Crosslinking of collagen increased mechanical strength and facilitated a similar cellular response to normal ECM whilst maintaining the biocompatibility of the scaffold. Nanofibrous scaffolds provide a high surface area allowing oxygen permeability and avoid fluid accumulation at the wound site, while preventing microbial infection by restricting their penetration through its small inter-fiber pores, thereby making them ideal candidates for wound dressings.209 The aim of using an electrospun scaffold based TE approach is to facilitate and promote the natural healing response of the body without inducing an immunogenic response, allowing the body to utilise the scaffold to enhance regeneration of “neonative” functional tissues.209 Electrospinning provides the flexibility to 36 coelectrospin polymers with drugs or proteins to form nanofibrous scaffolds for applications in drug delivery.209 Various polymers have been electrospun including synthetic ones such as poly(ε-caprolactone) (PCL), polylactic acid (PLA), polyglycolic acid (PGA), PLGA, polystyrene, polyurethane (PU), polyethylene terephthalate, poly(L-lactic acid)-co-poly(ε-caprolactone) (PLACL) and natural polymers including collagen, gelatin and chitosan to obtain fibers with diameters ranging from few nanometers to several microns24, 224, 228, 231.209 1.16 Skin regeneration using electrospun scaffolds The high surface area to volume ratio, porosity and structural similarity to the ECM architecture of the dermis makes nanopatterned electrospun fiber meshes ideal scaffolds for skin regeneration232-234.3 The limitations associated with electrospun scaffolds include poor mechanical properties, non-uniform thickness distribution and poor integrity.3 For skin regeneration a myriad of both natural polymers,235-237 synthetic polymers,232, 238-240 and combinations of the two have been successfully electrospun and tested.3 Poly(ε-caprolactone)/gelatin nanofibrous scaffold was electrospun on top of a commercial polyurethane wound dressing (TegadermTM; 3M Healthcare, USA) in an attempt to regenerate dermal wounds241.3 It was shown to promote cell proliferation, adhesion and growth in human dermal fibroblasts. Yang et al. employed emulsion electrospinning to fabricate ultrafine core sheath poly(ethylene glycol)-poly(D,L-lactide) fibers loaded with basic fibroblast growth factor (bFGF) and demonstrated its gradual release over 4 weeks232.3 An in vitro study using mouse embryonic fibroblasts showed enhanced cell adhesion, proliferation and secretion of ECM when cultured on a nanofibrous scaffold.232 In vivo studies using a dorsal wound model in diabetic rats treated with bFGF/ poly(ethylene glycol)-poly(D,L-lactide) mats showed elevated rates of healing with complete re-epithelialization and regeneration of skin appendages, while bFGF from the scaffold promoted collagen deposition and its remodelling to normal architecture232.3 In another study, potential of bone-marrow derived mesenchymal stem cells differentiation into the epidermal lineage (keratinocytes) was investigated 37 on poly(L-lactic acid)-co-(ε-caprolactone) nanofibrous scaffold, with and without collagen242.3 It was concluded that better results were obtained for collagen containing poly(L-lactic acid)-co-(ε-caprolactone) nanofibrous scaffolds.242 Electrospun nanofiber scaffolds have also been explored as delivery systems for variety of bioactive molecules including drugs, proteins and even RNA243-247.3 Ignatova et al. reviewed the use of nanofibrous scaffolds loaded with various antibiotics and antibacterial agents including nanoparticles for wound dressing applications248.249 Kenawy et al. electrospun poly(ethylene-co-vinyl acetate), poly(lactic acid) and their blend with tetracycline hydrochloride as a model drug250.249 It was reported that drug release behaviour was dependent on the polymer carrier and drug loading. Scaffolds with a 50/50 blend and relatively low drug loading (5 wt%) showed sustained release for over 5 days. Higher drug loading (25 wt%) resulted in an initial rapid burst release of the surface adsorbed drug that quickly dissolved in tris buffer.249 Sodium alginate was electrospun with poly(vinyl alcohol) containing different concentrations of ZnO nanoparticles as an antibacterial agent in a mouse fibroblasts model251.3 They observed inverse correlation between cell adhesion and spreading, and ZnO concentration in the scaffolds. The antibacterial activity was evaluated in both gram positive and gram negative bacterium cultures through diffusion disc experiments using Staphylococcus aureus and Escherichia coli. The results showed direct correlation between antibacterial activity of the scaffold and increasing concentration of ZnO nanoparticles251.3 Various strategies have been employed to control the drug release mechanism from fibers with the aim to prevent or control the initial burst release of the drug. Dave et al. developed electrospun enzyme-embedded antibiotic-releasing polymer scaffolds using antibiotic gentamicin sulfate (GS) and a polymer degrading enzyme (lipase) in PCL polymer to achieve endogenously triggered controlled drug release252.249The GS release from the scaffold was shown to be dependent on lipase concentration. In an another study, PDLLA was electrospun with Mefoxin to develop a nanofibrous scaffold with the loading efficiency of 90%.249 However, most of the drug was released in the first 3 h which reached its completion within 48 h, indicating that the drug was adsorbed on the nanofiber surface.249 In order to block such initial burst release of Mefoxin from PLGA nanofibers, Kim et al. used an amphiphilic block copolymer (PEG-b-PLA) and reported prolonged drug release for up to 1 week253.249 38 In an another attempt, Yohe et al. used superhydrophobic polymer made up from PCL doped with 0-50 wt% poly(glycerol monostreate-co-ε-caprolactone) (PGCC18) as a hydrophobic polymer and trapped air as a barrier to control the rate of drug release from the scaffold by blocking the fiber pores254, 255 249 . By using the model bioactive agent SN-38 (7-ethyl-10-hydroxy campthothecin), PCL electrospun fibers doped with 10 wt% PGC-C18 was shown to allow to have linear sustained release over 60 days compared to initial burst release (10 days) observed in case of PCL fibers alone which plateaued out at 20 days255.249Alternatively, covalent conjugation of the drug to the fibers can be used as a method to control and regulate the drug release mechanism from the fibrous scaffolds.249 Jiang et al. explored this by covalently functionalising ibuprofen onto poly(ethylene glycol)-g-chitosan (PEG-g-CHN) polymer and coelectrospinning it with poly(lactide-co-glycolide) (PLGA) to yield nanofibrous scaffold with prolonged release of the drug for more than 2 weeks256.249 Zou et al. copolymerised ε-caprolactone with PDLL, to induce functional ketone groups for later functionalization into a PDLL backbone, and used them to develop a fibrous scaffold257.249 Subsequently, the scaffold was surface functionalised with heparin molecules and loaded with bFGF. It was reported that bFGF release was dependent on the amount and molecular weight of heparin used. One of the biggest limitations with this approach is the necessity to select carrier polymers with desired functional groups257.249 In an alternate approach, polymer crosslinking has been explored as a parameter to control the initial burst release of the drug.249 Some of the methods used include UV-irradiation,258-260 dehydrothermal treatment,261 and chemical treatment including glutaraldehyde,262, prepared 263 formaldehyde264 and carbodiimide265, Fenbufen-loaded 266 249 poly(D,L-lactide-co-glycolide) . Meng et al. (PLGA) and PLGA/gelatin nanofibrous scaffolds267.249 Fenbufen release was found to be dependent on the concentration of gelatin used, alignment of the fibers in the scaffold and amount of crosslinking. It was shown that subsequent crosslinking treatment of the fibers effectively curtailed the burst release of the FBF at the initial release stage from a PLGA/gelatin (9/1) nanofibrous scaffold267.249 In an attempt to control the formation of hypertrophic scars (HS), poly(l-lactide) was coelectrospun with 20(R)-ginsenoside Rg3 (GS-Rg3).268 GS-Rg3 was used for its potential in inhibiting the formation of HS in vivo as measured using H&E staining and 39 apoptosis of fibroblasts to control the later stage HS hyperplasia by inhibiting inflammation and down-regulating VEGF expression.268 The in vivo wound rabbit ear HS model suggested sustained release of the drug for 3 months which was reported to be dependent on the drug concentration in the electrospun fibers. GSRg3/ poly(l-lactide) scaffold was shown to significantly inhibit HS formation with decreased expression of collagen fibers and microvessels.268 Bioactive molecules such as proteins, DNA, RNA and growth factors have also been encapsulated in polymer fibers mostly using modified techniques such as blend electrospinning and coaxial electrospinning.249 As the name suggests, blend electrospinning involve premixing of bioactive molecules with polymers prior to electrospinning increasing the surface localization efficiency of bioactive molecules.249 In an attempt to study if the activity of an encapsulated protein can be maintained after electrospinning, human β-nerve growth factor (hNGF) was electrospun along with BSA as carrier protein, into a partially aligned nanofibrous scaffold using poly ε-caprolactone (PCL) and poly(ethyl ethylene phosphate) (PEEP) polymers269.249 While the protein aggregates were heterogeneously distributed within the fibers, its bioactivity was demonstrated in a neurite outgrowth assay in PC12 cells. It was concluded that the prolonged release of an active protein can be achieved over the period of 3 months269.249 The same group also demonstrated the efficacy of this polymer system in the delivery of small interfering RNA (siRNA) and transfection reagent (TKO) complexes270.249 The siRNA/TKO complexes showed better gene knockdown efficiency than siRNA alone both of which were shown to be able to release from the scaffold for almost a month. Coaxial electrospinning was developed to fabricate core-shell fibrous scaffolds where both polymer and biomolecules are coaxially and simultaneously electrospun.249 The strategy involves use of the shell polymer to protect the cargo encapsulated in the core from the physiological degradation whilst allowing slow sustained release of the therapeutic.249 A nanofibrous scaffold encapsulating plasmid DNA (pDNA) within the PEG core and the non-viral gene carrier poly(ethyleimine)hyaluronic acid (PEI-HA) within the PCL sheath was prepared by the coaxial electrospinning technique271.249 They also studied the effects of various processing parameters using fractional factorial design. Extended release of the non-viral gene delivery vector from the sheath was observed over a period of 60 days quantified 40 using EGFP transfection expression and was shown to be dependent on pDNA loading into the fibers. Mickova et al. studied if liposome activity can be maintained post electrospinning process. They studied liposomes blended within nanofibers and polyvinyl alcohol-core/poly-ε-caprolactone-shell nanofibers with embedded liposomes prepared using coelectrospinning technique272.249 They concluded that liposomes encapsulated in the core/shell fibers retained enzymatic activity of encapsulated horseradish peroxidase while blended liposome became inactive. 1.17 Electrospun hybrid materials Despite the evidence of the capabilities of electrospun scaffolds in tissue engineering and drug delivery this technology has not reached its full potential in terms of translation. One of the biggest limitations associated with electrospun scaffolds developed from a single polymer has been the lack of multi-functionality and universality. They require specific surface modification approaches which are limited to specific polymers and can be difficult to translate into other polymer models. In an alternate approach, nanoparticles have shown tremendous potential and versatility not only in drug delivery applications but also in the fabrication of various functional materials which has already been translated into commercial products. Incorporation of these two technologies could open new avenues by utilising the potential of both technologies. In the following sections the potential of this combinatorial approach is demonstrated for applications in material science. This approach can be extrapolated towards tissue engineering using multifunctional scaffolds. Among the various fibers prepared using electrospinning, the nanoparticles (NPs) containing electrospun fibers exhibit a huge variety of potential applications.273 The composite fibrous mats show flexibility, are free standing,273 and incorporate the advantages of both starting materials i.e. polymer and the NPs. There are three main ways specified in the literature to fabricate NPs-electrospun fiber composite materials. 41 1. Electrospun fibers with surface functionalised nanoparticles: This is an efficient method where pre-treated electrospun scaffold is immersed into the colloidal NP solution to facilitate surface adsorption of NPs on fibrous scaffold.273 Various kinds of NPs such as metal, metal oxide, carbon, polymer, fluorescent NPs and even cells have been successfully combined with electrospun fibers to form composite materials15, 274, 275 273 . For example, a water stable poly(vinyl alcohol) (PVA)-AuNPs composite fibrous mat was fabricated by pretreating the surface of the fibers with 3-mercaptopropyltrimethoxysilane, allowing the adsorption of AuNPs on the surface of the fibers via Au-S bonds276.273 In cases where the electrospun fibers are stable but do not readily adsorb NPs, an in situ reduction method is used.273 For example, Au adsorbed TiO2 composite nanofibers were prepared by organic capping agent mediated photocatalytic reduction of HAuCl4277.273 It was reported that shape of the nanomaterial is dependent on type and concentration of the capping reagent.277 It provides the means to fabricate specific nano-structures on the nanofibers for respective applications in chemical and biological sensing277.273 Various NPs have been synthesised on electrospun fibers using this technique including Au, Ag, Pd, Pt, TiO2, WO3 and SnO2278-282.273 2. NPs encapsulating electrospun fibers: In this method NPs are synthesised within the polymer fibers.273 Typically, the polymer solution is electrospun with a metallic or ceramic acetate precursor which is then annealed to reduce the precursor and form NPs using methods such as gas-solid reaction, calcination or laser ablation.273 Yang et al. used AgNO3 as a precursor by electrospinning it with poly(acetonitrile) (PAN) which upon exposure to HCl yielded AgCl NPs both on the surface and interior of the fibers283.273 Similarly, copper nitrate was used as a precursor and electrospun with poly(vinyl butyral) (PVB) to give fibrous scaffold encapsulating copper nitrate. Polymer layer was removed by annealing the scaffold at 450 °C in an air atmosphere for 2 h. Subsequent heating at 300 °C for 1 h under hydrogen atmosphere was then applied to obtain copper fibers284.273This technique can also be used to fabricate more specialised structures.273 For example, Sn NPs were used as a precursor and were electrospun with porous multichannel carbon microtubes. The resulting scaffold underwent calcination to yield Sn-Carbon nanofibers285.273 42 3. NPs-electrospun fibers: This is a one-pot fabrication technique where NPs are homogenised in a polymer solution and electrospun together to obtain composite fibers.273 This technique relies on the homogeneity of the polymer-NP solution to yield scaffold with uniformly distributed NPs.273 Inorganic NPs can easily be encapsulated within the electrospun fibers because of their high electron density.273 Heterogeneous solution of NPs results in NP cluster formation within the electrospun fibers, which can be address by surface treatment of the fibrous scaffold.273 Both hydrophobic and hydrophilic polymers have been explored and electrospun. Some of the examples include (poly(ethylene oxide) (PEO), poly(vinyl alcohol) (PVA), poly(vinylpyrrolidone) (PVP), polyacetonitrile (PAN), PLGA and poly(L-lactide) (PLLA).273 Magnetite (iron oxide (Fe3O4) nanoparticles), another interesting class of nanoparticles have been explored for their magnetic resonance properties. These properties have been shown to be retained in the electrospinning process opening avenues for various applications.273 It can be used as intelligent fabrics in defence clothing, ultrahigh-density data storage, sensors and in health care.273 Problem in their use lies in their aggregation which tends to curb their magnetic response. It is understood that magnetite nanoparticles form clusters to reduce their energy because of their high surface area to volume ratio.273 Stabilisers have been employed to overcome this problem. The other alternatives include the use of electrostatic surfactant and steric polymers286-288.273 Coaxial electrospinning was used to address this problem of aggregation of nanoparticles. Core-shell Fe3O4-poly(ethylene terephthalate) magnetic composite nanofibers were electrospun where Fe3O4NPs form extended aligned structures within the core of the nanofibers and were further shown to have retained their super-paramagnetic behaviour289.273 The core-shell nanofibers demonstrated higher mechanical properties and responded to externally applied magnetic field due to the dipole-dipole interaction between magnetic NPs in magneto-rheological fluid.273 Other composites using similar approach include polyarylene ether nitriles-Fe-phthalocyanine-Fe3O4, Fe3O4-poly(ethylene oxide) and Fe3O4-poly(ethylene terephthalate)273, 289, 290, 291. Many metal and inorganic fluorescing NPs have also been electrospun for their luminescence properties including boron carbonitride oxide, sodium yttrium 43 fluoride, quantum dots and rare earth metals292, 293.273 Some of the examples includes Sm+3/TiO2, NaYF4:Yb+3, Er+3/SiO2 and YVO4: Eu+3/PEO for rare earth metals,294-296 CdTe and CdS quantum dots. #498;, #499}.273 As demonstrated, electrospun polymer/nanoparticle composite materials showcase hyphenated properties of the two constituents. This approach has great potential not only in material science but also in tissue engineering applications. Electrospun scaffolds can be functionalised with different types of nanoparticles to induce multifunctionality, for example magnetite and upconverting nanoparticles can be incorporated within the polymer scaffold to achieve sustained drug delivery and simultaneous imaging using MRI and fluorescence microscopy. 1.18 Summary With surgery continuing to be the mainstay for treating burn injury and reconstructing scars and limited evidence for the efficacy of cell therapies alone, there is a need for the development of alternative treatment modalities. M6P has been explored as an anti-scarring agent and its exogenous delivery has been shown to be antagonistic towards the activation of LTGFβ. TGFβ is a key cytokine that influences scar outcome. Upon activation, TGFβ upregulates the expression and synthesis of ECM proteins and molecules including collagen I, fibronectin, hyaluronic acid and α smooth muscle actin. The extent to which these molecules are expressed closely correlates to the extent of scarring. It has been proposed that clinical trials for commercial M6P (Juvidex®) could not meet the final end point because of the delivery mechanism adopted and the metabolic vulnerability of M6P. The mode of M6P delivery is critical as it gets rapidly metabolised and appropriate vehicles have to be employed to achieve and maintain requisite tissue levels over the post-wounding or healing period. This shortcoming of M6P needs to be addressed for its development as a potent anti-scarring drug and can potentially be solved through the development of more stable analogues. 44 Phosphonate analogues are metabolically more stable than phosphate analogues and they also project higher affinity towards the receptor (M6P/IGFII receptor). Two novel phosphonate analogues of M6P, analogue 1 and analogue 2 were investigated for their toxicity and efficacy in inhibiting the expression of collagen I under TGFβ1 stimulation in a primary skin dermal fibroblast in vitro model. With analogue 2, whilst it is able to penetrate cells due to its lipophilicity, this is also a limitation in its sustained topical delivery and therefore new modalities need to be explored to achieve local, controlled and consistent administration of the drug. Conventional skin substitutes are still prevalent in clinic use. However, they suffer from various disadvantages including low adhesion, creation of a new injury site (in autografts) and immune rejection (in allografts). To address these problems alternatives have been developed. Firstly, cell-based epidermal substitutes, efficient in accelerating the reepithelialisation phase of wound healing but only applicable in superficial deep wounds without additional therapies have been widely used. Dermal substitutes, both natural and synthetic have also been developed and used in deeper injuries. However, these dermal substitutes have problems with inadequate vascularisation, low mechanical strength, uncontrolled degradation profiles and high costs. Finally, the most advanced dermo-epidermal substitutes aimed at simultaneous regeneration of both epidermal and dermal layers by using constructs incorporating respective cells from both layers of the skin have more recently been trialled. The major limitations have been high production costs, poor vascularisation and failure to achieve permanent wound closure due to allogeneic cell rejection.3 Nanotechnology based techniques such as nanoparticles or nanofibrous scaffold based delivery systems have been shown to facilitate the delivery of hydrophobic drugs. Nanofibrous scaffolds are preferred because they mimic ECM, providing congenial environment for cellular ingrowth, and simultaneously deliver drugs to promote regeneration. In addition, the fractal-like geometries of fibrous scaffolds have been of interest recently to design and develop personalised diagnostic devices using fractal models of complex biological processes. For example, fractals have been used to describe the architecture of tumor microenvironments. Most of the polymers used thus far have been designed for specific applications and require specific approaches for surface modifications which are complex and laborious in order to induce any multifunctionality. PGMA, which has an epoxy group in each 45 monomer unit, would be an ideal polymer to address the concerns with multifunctionality. A nanofibrous fractal-like scaffold using PGMA was developed with an ability to incorporate multimodality and multifunctionality. It has been demonstrated that existing clinical treatments for wound healing have various shortcomings and there is a need for improved therapeutic approaches. TE inspired approaches including electrospun scaffolds, although having shown promise in addressing the shortcomings with conventional approaches, are marred with their own limitations restricting their clinical translation. These include poor mechanical strength, inadequate vascularisation and cellular ingrowth, difficult to control drug diffusion and limited ability to be functionalised. The inadequacy of these scaffolds for improving scar formation leaves a current unmet medical need. In this thesis, a multifunctional scaffold has been developed and characterised, with data describing its use in a range of applications presented. In addition, data showing the potential efficacy of a stable analogue of mannose-6-phosphate for wound healing has been demonstrated. This has led to initial work on a combinatorial approach using both scaffold and analogue to overcome many of the limitations previously discussed and potentially enhance wound healing and reduce scar formation in patients into the future. * PXS25 would be referred to as analogue 1 and PXS64 would be referred as analogue 2 except in the paper #3 1.19 Hypotheses and Aims Hypotheses 1. The epoxy functionalised polymers can be used to develop multifunctional fractal-like substrates 2. Stable isosteric analogues of mannose-6-phosphate (M6P) can be used to reduce the overexpression of collagen I gene 46 Aims 1. Optimise the parameters for electrospinning PGMA and study its efficacy as a multifunctional polymeric substrate 2. Investigate the biocompatibility and potential of two M6P analogues as anti- scarring agents in human dermal skin fibroblasts 3. Investigate the potential of a combinatorial scaffold:drug approach to limit fibrosis 47 Chapter 2 Introduction to the series of papers 2.1 Development of a universal multifunctional scaffold The first paper presented as part of this thesis presents the preparation of universal multifunctional fractal-like scaffold. Fractals are structures that maintain their structural characteristics with successive variation of scale.297 In addition to their applications in material science, fractals also find applications in cancer biology. They have been used to understand tumor architecture and morphology with implications in tumor growth and angiogenesis.298 Polyglycidyl methacrylate (PGMA), which contains a reactive epoxy unit per GMA monomer, can be functionalized with dyes, other polymers and drugs using S N2 nucleophilic substitution reactions. Amines, alcohols, carboxylic acids, alkyl halides, thiols and acid anhydrides are examples of highly reactive species that interact with PGMA.299 Different solvent systems such as acetone, chloroform and methylethyl ketone (MEK) were tested in the electrospinning of PGMA. Even though PGMA has high solubility in these solvents, acetone and chloroform were limited in their use because of their high vapor pressure and low conductivity. Therefore PGMA was electrospun in methyl ethyl ketone (MEK) to obtain a nanofibrous scaffold matrix (ES-PGMA). ES-PGMA was heated at 80 °C overnight to allow inter-crosslinking of the polymer. Fiber morphology was determined using scanning electron microscopy (SEM). The 48 PGMA-MEK system yielded uniform nanofibers with an average diameter of 0.69 ± 0.04 µm over a large area. PGMA is a hydrophobic polymer which needs to be functionalized to induce hydrophilicity to the backbone. This is usually carried out by reacting PGMA polymer backbone with hydrophilic reacting groups such as carboxylic acids and amines. ES-PGMA film was end-grafted with carboxylic acid-terminated poly (Nisopropyl acrylamide) (PNIPAM-COOH) to introduce ‘switchability’ in terms of thermally induced hydrophilic-hydrophobic behavior. PNIPAM is a thermoresponsive polymer which undergoes reversible phase transition in water at 32 °C (lower critical solution temperature (LCST)). This is caused by the transition from a swollen hydrated state to a shrunken dehydrated state above the LCST. Surface properties of the scaffolds were examined by contact angle measurements. End grafting of ES-PGMA with PNIPAM-COOH (ES-PGMA-g-PNIPAM-COOH nanofibers) resulted in dramatic temperature dependent changes in the contact angle, confirming the surfacing grafting of the hydrophilic PNIPAM moiety. Polymer composites comprising both polymers and nanomaterials have been regarded as a new class of hybrid materials with many functions. Lack of universality however is the major limitation in the development of a ubiquitous polymeric functional platform. PGMA was explored as a potential ubiquitous polymeric platform. PGMA was coelectrospun with three different types of nanoparticles: (NaGdF4:Yb, Er); palladium (Pd) and magnetite (Fe3O4) to develop polymer nanoparticle composite materials. These fibrous composite materials were characterized using SEM, transmission electron microscopy (TEM) and X-ray microanalysis. Coelectrospun polymer-nanoparticle composites were tested for their respective applications in upconversion fluorescent imaging in the case of (NaGdF4:Yb, Er), hydrogen gas sensing in the Pd composite and magneto-responsive behavior in the case of Fe3O4 composite matrix. Upconversion properties of the ES- PGMA/(NaGdF4:Yb, Er) nanocomposite was evaluated using NIR room temperature emission spectroscopy. This measures the efficiency of the nanocomposites in converting near infra-red excitation wavelength into visible emission. ESPGMA/(NaGdF4:Yb, Er) nanocomposites 49 showed prominent upconversion properties with three major emission peaks observed in the visible region at 521, 541, and 655 nm. Pd has been shown to demonstrate high selectivity and significant adsorption of hydrogen gas. Because of this feature Pd has been used in hydrogen gas sensors. Pd readily absorbs hydrogen gas which diffuses as atomic hydrogen into the lattice to form palladium hydride, PdHx, resulting in to phase transition and a corresponding change in the lattice spacing. Such changes in phase and lattice spacing cause a measurable resistance change in Pd material. Mostly, Pd has been adsorbed onto functionalized electrospun scaffolds. In this case, however, Pd was coelectrospun with PGMA to yield the fibrous composite material. ES-PGMA/Pd composites were evaluated for their hydrogen gas sensing capacity using alternating concentrations of nitrogen and hydrogen gas. Change in current was measured using current-voltage (I-V) sweeps. Response time 90 of ~14 seconds was observed for a hydrogen gas concentration range of 1-10% from 0.16 ng of Pd as ES-PGMA/Pd fibrous composite mounted on 650 x 900 µm IED platform. More recently magnetic materials and matrices were developed for potential applications in monetary currency and defense fabric materials. ES-PGMA/Fe3O4 nanocomposite fibers were evaluated for magnetization properties using SQUID magnetometry (Superconducting Quantum Interference Device). SQUID analysis confirmed superparamagnetic behavior of the composite material at room temperature. The mass specific saturation magnetization (Ms) of the fibers was 4.0 emu g−1. Results are presented in Agarwal, V., Ho, D., Ho, D., Galabura, Y., Yasin, F. M.D., Gong, P., Ye, W., Singh, R., Munshi, A., Saunders, M., Woodward, R. C., St. Pierre, T., Lorenser, D., Wood, F. W., Fear, M., Sampson, D. D., Zdyrko, B., Smith, N.M., Luzinov, I., Iyer, K.S., A Functional Reactive Polymer Nanofiber Matrix, RSC Advances (Submitted) 50 2.2 Evaluation of mannose-6-phosphate analogues as potential antiscarring agents M6P has been shown to control and inhibit the secretion of extracellular matrix proteins including collagen I, one of the key proteins in granulation tissue and scars.144, 145, 300 M6P inhibits TGFβ1 activation, a primary cytokine regulating wound healing and subsequent scar formation, by competitively binding to M6P/IGFII receptor.301 Recombinant M6P (Juvidex®) has been examined in phase II randomised human clinical trials for its efficacy in wound healing and scar reduction at split thickness skin graft donor sites.153 Intradermal delivery of Juvidex showed an accelerated wound healing response. However, no significant reduction was observed in scar formation post healing. It has been postulated that the metabolic vulnerability and delivery mechanism for M6P could have been the limiting factors behind its translation into the clinic. Two new bioisosteric phosphonate analogues of M6P were investigated (Aim 2, analogues were kindly provided by Pharmaxis Pvt Ltd and used without any further modification). Phosphonate drugs were selected because of their inherently higher stability towards phosphatase enzymes.159 Analogue 1 was designed by replacing the P-O bond at C6 with the methylene bridge at C6 in M6P to enhance its stability against phosphatases present in serum. In addition, the anomeric hydroxyl was substituted with m-xylene group as they have been reported to improve their recognition by the receptor.150 Analogue 2 was designed as a prodrug by further derivatization of analogue 1 with bis(pivaloyloxy)methyl (POM) linkers to yield a neutral product. Neutral analogues are interesting because they get rapidly internalized into the cell, where the POM linkers are gradually hydrolysed by microsomal esterases to release an active analogue 1. Hence analogue 2 was developed to achieve intracellular targeting of the M6P/IGFII receptor. In vitro behaviour of the analogues was assessed in primary human dermal skin fibroblasts (HDF) and compared against M6P. Although binding to the M6P/IGFII receptor is one of the key mechanisms behind LTGFβ activation, there are other alternate modes of activation of LTGFβ in vivo. Therefore, it is also important to investigate the efficacy of the new analogues in the presence of recombinant active 51 human TGFβ1. To assess the dose response of the analogues, HDF cells were incubated with varying concentrations of the analogues both in the presence and absence of TGFβ1. The MTS assay was used to assess the cytotoxicity of the analogues. Cells were incubated with the analogues, both in presence and absence of TGFβ1, over a period of 72h. At the specified time, MTS reagent was added to the culture and further incubated to allow metabolically active cells to bioreduce the MTS reagent and form the coloured formazan product which is measured by spectrometry at 490 nm. Colour intensity is correlated to the number of metabolically active cells in culture i.e. proliferating cells. TGFβ1 has been reported to curtail the proliferation of HDF as was observed here. However, no add on effect of the analogues was observed at any of the concentrations studied. Both analogue 1 and 2 behaved similarly to M6P in terms of inherent toxicity. Analogue 2 however, showed toxicity at high concentrations (> 50 µM) due to the formation of formaldehyde, which is one of the degradation products produced as a result of esterase hydrolysis of POM linkers in analogue 2. A complimentary live/dead assay was carried out to study the effects of the analogues on cell viability over a period of 72 h. Cells were incubated with varying concentrations of analogues. At the specified time-point calcein AM/ ethidium bromide I dyes were added. Calcein stains the intracellular esterase in viable cells and fluoresces green whereas ethidium bromide stains non-viable cells and fluoresces red due to the binding of the ethidium homodimer with nucleic acids penetrated by ethidium through the compromised cell membranes. No loss was observed in cell viability. Cell morphology was also assessed. TGFβ1 stimulation significantly inflated the cell body area compared to the non-treated control cells which was similar to what has been reported previously. Interestingly, this inflation in cell area came down to normal levels post analogue treatment. No apparent alteration in cell body area was observed in cells treated with varying concentrations of analogue alone without TGFβ1 stimulation. The potential of the analogues as anti-scarring agents was evaluated in terms of their ability to inhibit the overexpression of key fibrotic indicators such as collagen I gene 52 expression. Collagen I is a key fibrotic marker which has major implications in wound healing and formation of granulation tissue. In order to curb the secretion of collagen matrix it is crucial to control its expression at the transcriptional level. TGFβ1 is known to up-regulate the expression of collagen I in HDF. Real time-quantitative polymerase chain reaction (RT-qPCR) analysis was carried out to quantitate the expression of collagen I post-treatment. TGFβ1 stimulation has been reported to significantly up-regulate the expression of collagen I gene expression as was observed here compared to untreated control cells. Efficiency of the analogues was tested in terms of their potential in inhibiting this increase in collagen I mRNA. 10 µM analogue 2 significantly reduced the TGFβ1 mediated increase in collagen I gene expression. This response was similar to that observed with M6P treatment. The hydrophilic analogue 1 did not show any response. It can be deduced that analogues behave similar to M6P. Further, intracellular targeting of the M6P/IGFII receptor, using analogue 2 as a prodrug of analogue 1, is more efficacious compared to intercellular targeting using analogue 1. Results are presented in Agarwal, V., Toshniwal, P., Smith, N. E., Smith, N. M., Li, B., Clemons, T. D., Byrne, L. T., Hassiotou, F., Wood, F. M., Fear, M., Corry, B., and Iyer, K. S., Enhancing the Efficacy of Cation-Independent Mannose 6Phosphate Receptor Inhibitors by Intracellular Delivery, Angewandte Chemie International Edition (Submitted) 2.3 Delivery of the lipophilic mannose-6-phosphate analogue PXS64 using an electrospun PGMA scaffold One of the biggest limitations in the clinical translation of lipophilic drugs has been their delivery and associated serum stability. Rapid degradation, particularly in a wound environment, is a common issue. In order to achieve prolonged stability, drugs are encapsulated in polymer shells to protect them. One of the niche ways to achieve controlled delivery of sensitive lipophilic drugs is by encapsulating them in nanofibrous scaffolds developed using electrospinning.302 Drugs get released 53 through a diffusion mechanism in a controlled rate over time sufficient to be taken up by the cells before degradation.303 Analogue 2 was coelectrospun with PGMA to fabricate a nanofibrous scaffold (Aim 3). Despite proven toxicity at high concentrations, analogue 2 was selected for its potential in reducing the expression of fibrotic marker, collagen I. In addition, it was rationaled that the release of the drug would be slow and within the tolerated limit. The electrospun scaffold was characterised using scanning electron microscopy (SEM) and drug loading was estimated using HPLC. Release studies were also attempted using HPLC. It is important to conduct release studies in aqueous conditions to mimic the in vivo environment. No drug release was detected from the fibers in the HPLC analysis. It was implicated that either there is no drug in the fibers or it may have been degraded during the electrospinning process. Alternatively, it was not being released. HPLC confirmed the loading of intact drug at ~76 %. Cell-matrix interactions were assessed by fluorescent microscopy and SEM. In vitro studies were carried out on human dermal fibroblasts (HDF) using electrospun PGMA scaffold as a control matrix. It was hypothesised that cells may be more sensitive to the released drug and that despite the lack of release data obtained there would be a detectable physiological effect of the drug even at very low concentrations. Cytotoxicity studies using the MTS assay showed no apparent toxicity of either control or test scaffolds. Interestingly, no reduction was observed in cell proliferation incubated on the scaffolds post TGFβ1 stimulation. Cell viability study using calcein AM/ethidium bromide I dyes confirmed cell viability. Both florescent and electron microscopy revealed positive interactions between cells and the fibrous scaffolds. Cells were seen to take their extended spindle shape morphology as confirmed by SEM images. RT-qPCR studies were carried out to study the effect of the scaffolds on upregulated expression of collagen I post TGFβ1 stimulation. It was observed that cells cultured on scaffolds containing drug (ES-PGMA + analogue 2) showed a tendency for a reduced increase in expression of collagen I after TGFβ1 stimulation. However, this was not statistically significant. Therefore, it is likely that the drug is restricted within the fibers and is not able to be released. It has previously 54 been reported that DMSO can facilitate drug release. It facilitates drug release by providing a congenial environment for drug dissolution in an otherwise hydrophilic physiological environment. When DMSO was incorporated in the culture, a significant reduction in collagen I gene expression was attained as anticipated. No significant effect was observed of DMSO alone or DMSO and control scaffold. Results are presented in Agarwal, V., Wood, F. M., Fear, M. and Iyer, K. S., Inhibiting the activation of transforming growth factor-β using a polymeric nanofiber scaffold, Nanoscale (Submitted) 55 Chapter 3 Series of papers The results of this thesis are presented in the following series of published papers or submitted manuscripts, the citations of which are listed below. Supporting information for these manuscripts can be found in Appendix A, where applicable. 1. Agarwal, V., Ho, D., Ho, D., Galabura, Y., Yasin, F. M.D., Gong, P., Ye, W., Singh, R., Munshi, A., Saunders, M., Woodward, R. C., St. Pierre, T., Wood, F.M., Fear, M., Lorenser, D., Sampson, D. D., Zdyrko, B., Smith, N.M., Luzinov, I., Iyer, K.S., A Functional Reactive Polymer Nanofiber Matrix, RSC Advances (Submitted) 2. Agarwal, V., Toshniwal, P., Smith, N. E., Smith, N. M., Li, B., Clemons, T. D., Byrne, L. T., Hassiotou, F., Wood, F. M., Fear, M., Corry, B., and Iyer, K. S., Enhancing the Efficacy of Cation-Independent Mannose 6-Phosphate Receptor Inhibitors by Intracellular Delivery, Angewandte Chemie International Edition (Submitted) 3. Agarwal, V., Wood, F. M., Fear, M. and Iyer, K. S., Inhibiting the activation of transforming growth factor-β using a polymeric nanofiber scaffold, Nanoscale (Submitted) 56 Journal Name RSCPublishing COMMUNICATION A Functional Reactive Polymer Nanofiber Matrix Cite this: DOI: 10.1039/x0xx00000x Received 00th January 2012, Accepted 00th January 2012 Vipul Agarwal,a Dominic Ho,a Diwei Ho,a Yuriy Galabura,b Faizah M.D. Yasin,a Peijun Gong,c Weike Ye,d Ruhani Singh,a Alaa Munshi,a Martin Saunders,e Robert C. Woodward,f Timothy St. Pierre,f Fiona M. Wood,g Mark Fear,g Dirk Lorenser,c David D. Sampson,c,e Bogdan Zdyrkob, Nicole M. Smith,a Igor Luzinovb,*, and K. Swaminathan Iyera,* DOI: 10.1039/x0xx00000x www.rsc.org/ Synthetic fractal materials have been regarded as a new class of hybrid materials with many potential applications. However, the lack of an efficient, reactive large-area fractal substrate has been one of the major limitations in the development of these materials as advanced functional platforms. Herein, we demonstrate the utility of electrospun polyglycidyl methacrylate (PGMA) fractal-like films as a highly versatile platform for the development of functional nanostructured fractal-like materials anchored to a surface. The utility of this platform as a reactive substrate is demonstrated by grafting poly (N-isopropyl acrylamide) to incorporate stimuli-responsive properties. Additionally, we demonstrate that functional fractal-like nanocomposites can be fabricated using this platform with properties for sensing, fluorescence imaging and magneto-responsiveness. The development of nanostructured polymeric matrices to obtain organic-inorganic nanocomposites has been actively researched to produce hybrid materials for applications in electronics, optics, medical devices, sensors and catalysis. 1-4 Of the various techniques developed to produce large area nanoscale polymeric matrices, one of the most researched, cost effective and facile method is electrospinning. It has been adapted to cover a wide range of polymers and optimized to regulate fiber diameter, alignment and shape. 5-7 There have been numerous reports using this technique to develop matrices with enhanced mechanical strength, 8 matrices with selective filtration/permeability, 9 fire retarding material, optoelectronic devices10 and substrates for catalysis. 11, 12 One of the key steps involved in the development of organic-inorganic nanocomposites is grafting to achieve excellent integration by minimizing interfacial tension of the nanoparticles in the This journal is © The Royal Society of Chemistry 2012 organic nanofiber matrix. Additionally, the ability to modify the surface of the nanofiber to alter the adhesion, lubrication, wettability and biocompatibility is pivotal in its customisation for end-use applications. The achievement of a certain degree of grafting universality requires the establishment of a controlled method of introducing the desired functional groups on a substrate.5 Currently, this is achieved by physisorption. 13, 14 In contrast, chemisorption, which is difficult to achieve on a polymeric nanofibers, would result in permanent irreversible surface modification. Polymers containing epoxy groups are examples of functional polymers that are able to react with a wide range of substrates through ‘‘grafting to’’ interactions mediated by the epoxy groups.15, 16 The versatile chemistry of epoxy groups renders a polymer that is exceptionally suitable as a universal electrospun nanofiber matrix to provide reactive groups for further grafting reactions. To this end, poly(glycidyl methacrylate) (PGMA), which contains an epoxy group in every repeating unit, has been used extensively as a macromolecular anchoring layer for grafting of polymers to the surfaces.17-20 Upon electrospinning, epoxy groups in the polymer will undergo self-crosslinking upon heating, providing mechanical integrity to the matrix. 21 Approximately 40% of the epoxy groups are still available for surface modification following a 12 hour treatment at 120°C. The main advantage of using PGMA as a matrix for electrospinning, as opposed to modifying the surface using monolayers, is the high mobility of the epoxy groups located in the “loops” and “tails” of the polymer. The mobility of the free groups results in the formation of a highly effective interpenetrating anchoring zone. 22 In this article, we report that PGMA can be directly electrospun (ES-PGMA) to form large area nanofibers. We demonstrate that this polymer nanofiber matrix can be used as an effective platform to graft polymers to impart switchability, and can be used to produce J. Name., 2012, 00, 1-3 | 1 COMMUNICATION nanocomposites with upconverting properties, with hydrogen sensing capability or with magneto-responsive properties. The ES-PGMA nanofibers (see Supporting Information for method of synthesis) were uniform over a large area and had an average diameter of 0.69 ± 0.04 µm (average ± standard error mean) (Fig. 1A). The average thickness of the 1 x 1 cm 2 ESPGMA generated was 127 ± 3 µm in 7 hours (Fig. 1B). In order to test the efficacy of the ES-PGMA nanofiber matrix as an anchoring platform, carboxylic acid-terminated poly (Nisopropyl acrylamide) (PNIPAM-COOH) was end grafted to the nanofibers via a ring opening reaction with the epoxy groups to yield ES-PGMA-g-PNIPAM-COOH.23 PNIPAM is a thermo-responsive polymer, which has been utilized in various Fig. 1: (A) SEM secondary electron image of the electrospun PGMA (ESPGMA) fibers, (B) cross-sectional image of ES-PGMA, (C) Water contact angle θ = 60° at 70°C, (D) Water contact angle θ = 15° at room temperature respectively measured on ES-PGMA -g-PNIPAM-COOH. forms, such as thermo-responsive hydrogels, particles, brushes, spheres and micelles. 24-27 Importantly, PNIPAM exhibits a 2 | J. Name., 2012, 00, 1-3 Journal Name temperature-sensitive phase transition in water at what is known as a lower critical solution temperature (LCST), 32 °C. 28 The transition is due to the coil-to-globule transition at the critical temperature resulting in switching from hydrophilic to hydrophobic behavior.29 At temperatures below the LCST, PNIPAM chains arrange into an expanded and hydrated conformation. Conversely, at temperatures above the LCST, PNIPAM chains collapse and arrange into a compact, dehydrated conformation. 29 This thermo-responsive behavior is retained post-grafting and post-end group functionalization. The ES-PGMA-g-PNIPAM-COOH nanofibers demonstrated thermo-responsive behavior as monitored using contact angle measurements. The contact angle changed from 60 ± 2° at 70°C (Fig. 1C) to 15 ± 2° at room temperature (Fig. 1D). The contact angle of the unmodified ES-PGMA remained unchanged at 100 ± 2° both at room temperature and at 70°C. The ability of the ES-PGMA nanofiber matrix to produce nanocomposites was further evaluated using three distinct classes of nanoparticles: upconverting fluorescent particles of NaGdF4:Yb, Er (UCNP), palladium (Pd) and magnetite (Fe3O4). The nanoparticles synthesized (see Supporting Information for methods) had a narrow size distribution of 7.4 ± 1.4 nm (average ± standard error mean) for UCNP, 19.3 ± 0.2 nm for Pd and 6.7 ± 1.4 nm for Fe3O4 respectively (Fig. 2A, C and E). One of the major hurdles in developing functional materials by electrospinning nanocomposites is the lack of control in attaining a homogeneous distribution of nanoparticles throughout the polymer matrix. In the present case, electrospinning PGMA with the aforementioned nanoparticles resulted in relatively uniform distributions of the nanoparticles throughout the fiber matrix (Fig. 2B, D and F) which was observed through various images obtained at similar fields of view. It has been reported that variations in solution properties such as surface tension and solution conductivity in the presence of nanoparticles result in changes in the nanofiber diameter. 30-32 In the present case, the electrospun fiber diameter increased in the presence of nanoparticles to 2.56 ± 0.16 µm (average ± standard error mean) for UCNP, 1.75 ± 0.07 µm for Pd and 4.37 ± 0.44 µm for Fe3O4 (Insets in Fig. 2G, H and I, respectively). Small fibers were chosen for TEM analysis because of the contrast problems related to thicker samples. The ability of the nanocomposites to be used as functional materials was evaluated by testing the upconverting properties, hydrogen sensing properties and magnetic properties of the, UCNP/ES-PGMA, Pd/ES-PGMA and Fe3O4/ES-PGMA fibers, respectively. In the case of UCNP/ES-PGMA fibers, the ability to convert near-infrared excitation into visible emission was evaluated (see supporting information for methods). UCNP have been successfully used as ultrasensitive magnetic/upconversion fluorescent dual-modal molecular probes for MRI and upconversion fluorescence imaging. 33-35 In the present case the UCNP/ES-PGMA fibers demonstrated excellent upconversion properties upon 974 nm laser excitation (Fig. 3A). The three major emissions were located at 521, 541, and 655 nm. Green emission from 500- 600 nm was attributed to 2H11/2 → 4I15/2 and 4S3/2 → 4I15/2 transitions, respectively, and the red emission from 635-670 nm was attributed to the 4F9/2 → 4 I15/2 transition (Fig. 3A). The green to red (G/R) ratio for the fibers was 1.35:1. The UCNP/ES-PGMA composite fibers retained the upconversion signal levels to the pure NaGdF4:Yb,Er nanoparticle samples (G/R ratio = 1.37:1). This journal is © The Royal Society of Chemistry 2012 Journal Name Fig. 2: TEM images of the (A) Upconverting nanoparticles (UCNP), (B) UCNP/ES-PGMA composite fibers, (C) Pd nanoparticles, (D) Pd/ES-PGMA composite fibers, (E) magnetite (Fe3O4) nanoparticles, (F) Fe3O4/ES-PGMA composite fibers. X-ray microanalysis spectrum obtained on: (G) UCNP/ESPGMA composite fibers showing the presence of Gd and Yb among other elements (Inset: SEM micrograph of UCNP/ES-PGMA composite fibers): (H) Pd/ES-PGMA composite fibers showing the presence of Pd (Inset: SEM micrograph of Pd/ES-PGMA composite fibers) and (I) Fe3O4/ES-PGMA composite fibers (Inset SEM micrographs of Fe3O4/ES-PGMA composite fibers). Scale bars for images A, C and E 10 nm, for images B, D and F 1 µm and for inset images G, H and I 20 µm The Pd/ES-PGMA nanofibers were evaluated for sensing hydrogen (see Supporting Information for methods). Palladium has emerged as an important candidate for hydrogen gas sensing because of its ability to absorb high quantities of hydrogen and its highly selective response. 36 Sensing herein is based on the well-established principle that palladium spontaneously absorbs H 2 gas as atomic hydrogen which diffuses into the lattice to form palladium hydride, PdH x, resulting in an to phase transition and a corresponding change in the lattice spacing. 37 The change in phase and lattice spacing leads to a measurable resistance change of the palladium material. However, either replacing precious noble metals with cheaper materials or alternatively development of methods that result in the reduction of material used by several orders of magnitude, especially in applications that require large amounts of material, would be beneficial. Currently, hydrogen sensing platforms are based on all-palladium constructs or hybrids with high Pd loading to stimulate an effective sensing response. Herein, using the electrospun polymer/nanoparticle nanocomposite material we demonstrate a This journal is © The Royal Society of Chemistry 2012 COMMUNICATION Fig. 3: (A) Upconversion fluorescence spectrum of both UCNP’s (blue) and UCNP/ES-PGMA fiber composite (black) showing three main emissions green at 521 and 541 nm and red between 635 and 670 nm upon 974 nm laser excitation, (B) Current response of the Pd/ES-PGMA matrix sensor to 1-10% hydrogen gas, with alternating 4 min hydrogen and 20 min. nitrogen exposure, (C) Zero-field cooled (orange) and field cooled (blue) curves for Fe3O4/ES-PGMA composite (Inset: hysteresis loop at 5 K (pink) and 300 K (green) for Fe3O4/ES-PGMA composite) as measured by SQUID magnetometry. response is obtainable for as low as 0.6 ng of Pd dispersed across a 650 m x 900 m area over interdigitated electrodes (IDE). The ability of Pd/ES-PGMA nanofibers to sense different hydrogen concentrations (between 1 and 10% in N 2 as a carrier gas) was tested 38 (Fig. 3B). An increase in resistance with hydrogen gas-flow and a return to the original state in the absence of a hydrogen gas flow was observed for hydrogen concentrations (1-10%) with a response time 90 of ~14 seconds (Fig. 3B). Finally, the magnetization properties of the Fe 3O4/ESPGMA fibers were measured by SQUID magnetometry J. Name., 2012, 00, 1-3 | 3 COMMUNICATION (Superconducting Quantum Interference Device) (see Supporting Information). The Fe3O4/ES-PGMA fibers are superparamagnetic at room temperature with the zero field cooled/field cooled curves showing a maximum blocking temperature of 30 K (where the two curves merge) and the absence of hysteresis at 300K (Fig. 3C). The mass specific saturation magnetization, M s of the fibers was 4.0 emu g−1. Particle loading was estimated to be ~7% by weight as determined from the M s values of the Fe3O4 nanoparticles and Fe3O4/ES-PGMA fibers. Conclusions In summary, we have developed a robust polymeric platform for the large scale production of electrospun nanofibers based on poly(glycidyl methacrylate) (PGMA). We have demonstrated that the epoxy groups of the polymeric matrix can be effectively used as a grafting platform for surface modifications and the polymer serves as an excellent platform to fabricate functional nanocomposites. We believe our findings presented herein will aid in the design of novel electrospun materials with tailorable surfaces for applications as scaffolds in regenerative medicine, optoelectronics, magnetic filtration and catalysis. Notes and references a School of Chemistry and Biochemistry, The University of Western Australia, WA 6009, Australia b Department of Materials Science and Engineering, Clemson University, Clemson, SC 29634, USA c School of Electrical, Electronic and Computer Engineering, The University of Western Australia, WA 6009 Australia d School of Chemistry and Chemical Engineering, Nanjing University, China e Centre for Microscopy, Characterisation and Analysis, The University of Western Australia, WA 6009 Australia f School of Physics, The University of Western Australia, WA 6009 Australia g Burn Injury Research Unit, School of Surgery, The University of Western Australia, WA 6009, Australia E-mail: swaminatha.iyer@uwa.edu.au, luzinov@clemson.edu The authors would like to acknowledge the Australian Microscopy & Microanalysis Research Facility at the Centre for Microscopy, Characterization & Analysis, The University of Western Australia, funded by the University, State and Commonwealth Governments. Authors would also like to thank Dr C.W.Evans and Dr Peter R. T. Munro for assistance with valuable experimental and result discussions. Peijun Gong is supported by The University of Western Australia and the China Scholarship Council. † Footnotes should appear here. These might include comments relevant to but not central to the matter under discussion, limited experimental and spectral data, and crystallographic data. Electronic Supplementary Information (ESI) available: [details of materials and methods]. See DOI: 10.1039/c000000x/ 1. S. Komarneni, J. Mater. Chem., 1992, 2, 1219-1230. 2. D. R. Paul and L. M. Robeson, Polymer, 2008, 49, 3187-3204. 4 | J. Name., 2012, 00, 1-3 Journal Name 3. T. H. Hwang, Y. M. Lee, B.-S. Kong, J.-S. Seo and J. W. Choi, Nano Lett., 2012, 12, 802-807. 4. P. Wang, L. Zhang, Y. Xia, L. Tong, X. Xu and Y. Ying, Nano Lett., 2012, 12, 3145-3150. 5. Z.-M. Huang, Y. Z. Zhang, M. Kotaki and S. Ramakrishna, Compos. Sci. Technol., 2003, 63, 2223-2253. 6. C. Huang, S. J. Soenen, J. Rejman, B. Lucas, K. Braeckmans, J. Demeester and S. C. De Smedt, Chem. Soc. Rev., 2011, 40, 24172434. 7. A. Baji, Y.-W. Mai, S.-C. Wong, M. Abtahi and P. Chen, Compos. Sci. Technol., 2010, 70, 703-718. 8. I. S. Chronakis, J. Mater. Process. Tech., 2005, 167, 283-293. 9. USA Pat., 1986. 10. X. Wang, C. Drew, S.-H. Lee, K. J. Senecal, J. Kumar and L. A. Samuelson, Nano Lett., 2002, 2, 1273-1275. 11. E. Formo, E. Lee, D. Campbell and Y. Xia, Nano Lett., 2008, 8, 668672. 12. Z. Liu, D. D. Sun, P. Guo and J. O. Leckie, Nano Lett., 2006, 7, 1081-1085. 13. B. Zhao and W. J. Brittain, Prog. Polym. Sci., 2000, 25, 677-710. 14. S. Edmondson, V. L. Osborne and W. T. Huck, Chem. Soc. Rev., 2004, 33, 14-22. 15. K. S. Iyer, B. Zdyrko, H. Malz, J. Pionteck and I. Luzinov, Macromolecules, 2003, 36, 6519-6526. 16. B. Zdyrko and I. Luzinov, Macromol. Rapid Comm., 2011, 32, 859869. 17. K. S. Iyer and I. Luzinov, Macromolecules, 2004, 37, 9538-9545. 18. A. W. Eckert, D. Gröbe and U. Rothe, Biomaterials, 2000, 21, 441447. 19. A. Sidorenko, T. Krupenkin, A. Taylor, P. Fratzl and J. Aizenberg, Science, 2007, 315, 487-490. 20. B. Zdyrko, K. Swaminatha Iyer and I. Luzinov, Polymer, 2006, 47, 272-279. 21. US Pat. US 2004/0185260 A1, 2004. 22. I. Luzinov, V. Klep, S. Minko, K. S. Iyer, J. Draper and B. Zdyrko, PMSE Prep., 2004, 90, 224-225. 23. D. Suzuki and H. Kawaguchi, Langmuir, 2005, 21, 12016-12024. 24. D. Schmaljohann, Adv. Drug Deliver. Rev., 2006, 58, 1655-1670. 25. J. M. Weissman, H. B. Sunkara, A. S. Tse and S. A. Asher, Science, 1996, 274, 959-963. 26. H. Kawaguchi, Prog. Polym. Sci., 2000, 25, 1171-1210. 27. Y. Li, B. S. Lokitz, S. P. Armes and C. L. McCormick, Macromolecules, 2006, 39, 2726-2728. 28. J. Wu, B. Zhou and Z. Hu, Phys. Rev. Lett., 2003, 90, 048304. 29. R. Yoshida, K. Uchida, Y. Kaneko, K. Sakai, A. Kikuchi, Y. Sakurai and T. Okano, Nature, 1995, 374, 240-242. 30. H. Kang, Y. Zhu, X. Yang, Y. Jing, A. Lengalova and C. Li, J. Colloid Interf. Sci., 2010, 341, 303-310. 31. S. Koombhongse, W. Liu and D. H. Reneker, J. Polym. Sci. Pol. Phys., 2001, 39, 2598-2606. 32. Q. P. Pham, U. Sharma and A. G. Mikos., Tissue Eng., 2006, 12, 1197-1211. 33. J. Ryu, H.-Y. Park, K. Kim, H. Kim, J. H. Yoo, M. Kang, K. Im, R. Grailhe and R. Song, The J. Phys. Chem. B, 2010, 114, 21077-21082. 34. J.-C. Boyer, F. Vetrone, L. A. Cuccia and J. A. Capobianco, J. Am. Chem. Soc., 2006, 128, 7444-7445. This journal is © The Royal Society of Chemistry 2012 Journal Name COMMUNICATION 35. J. Zhou, Y. Sun, X. Du, L. Xiong, H. Hu and F. Li, Biomaterials, 2010, 31, 3287-3295. 36. A. Kolmakov, D. Klenov, Y. Lilach, S. Stemmer and M. Moskovits, Nano Lett., 2005, 5, 667-673. 37. F. Favier, E. C. Walter, M. P. Zach, T. Benter and R. M. Penner, Science, 2001, 293, 2227-2231. 38. J. Zou, B. Zdyrko, I. Luzinov, C. L. Raston and K. Swaminathan Iyer, Chem. Comm., 2012, 48, 1033-1035. This journal is © The Royal Society of Chemistry 2012 J. Name., 2012, 00, 1-3 | 5 Intracellular Drug Delivery DOI: 10.1002/anie.201((will be filled in by the editorial staff)) Enhancing the Efficacy of Cation-Independent Mannose 6-Phosphate Receptor Inhibitors by Intracellular Delivery** Vipul Agarwal, Priyanka Toshniwal,# Natalie E Smith,# Nicole M Smith, Binbin Li, Tristan D Clemons, Lindsay T Byrne, Foteini Hassiotou, Fiona M Wood, Mark Fear, Ben Corry* and K Swaminathan Iyer* Abstract: Extracellular targeting of the cation-independent mannose 6-phosphate/insulin-like growth factor II (M6P/IGFII) receptors has been an attractive approach for the development of antifibrotic drugs. Several M6P analogues have been developed to regulate the activity of transforming growth factor-β1 (TGFβ1) by inhibiting its conversion from the latent to the active form. Herein, we adopt a combinatorial approach using an in vitro wound healing model and molecular dynamic simulations, to reveal that the efficacy of M6P/IGFII inhibitors can be significantly enhanced by adopting an intracellular approach. We demonstate this using systematic analysis of a bioisosteric M6P analogue, a lipophilic prodrug which upon cellular internalization undergoes ester hydrolysis to yield an [] V. Agarwal, P. Toshniwal, Dr. N.M. Smith, Dr. T.D. Clemons, Dr. F. Hassiotou, Prof. Dr. K. S. Iyer School of Chemistry and Biochemistry, The University of Western Australia, Crawley, Western Australia, Australia E-mail: swaminatha.iyer@uwa.edu.au B. Li State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan, PR China Dr. L. T. Byrne Centre for Microscopy, Characterization & Analysis, The University of Western Australia, Australia Dr. N. E. Smith, Dr. B. Corry Research School of Biology, Australian National University, Canberra, Australian Capital Territory, Australia Email: ben.corry@anu.edu.au Dr. M. Fear, Prof. Dr. F. M. Wood Burn Injury Research Unit, School of Surgery, The University of Western Australia, Crawley, Western Australia, Australia [] The authors would like to thank the Australian Research Council for funding the work, Pharmaxis Pvt Ltd for kindly providing analogues 1 and 2. The authors would also like to acknowledge the Australian Microscopy & Microanalysis Research Facility at the Centre for Microscopy, Characterization & Analysis, The University of Western Australia, funded by the University, State and Commonwealth Governments. Authors would like to thank Dr Bernard Callus, Dr Cameron Evans and Dr Megan Finch for their assistance with experimental analysis. [#] These authors contributed equally active M6P analogue to effectively downregulate collagen 1 expression in primary human dermal skin fibroblasts. In mammalian cells, the cation-independent mannose 6phosphate/insulin-like growth factor II (M6P/IGFII) and cationdependent mannose 6-phosphate (CD-MPR) receptors, have been identified as pivotal targets that modulate cellular response because of their role in protein trafficking. Both these receptors are functionally complimentary and can partially compensate for the absence of the other.[1] These sorting receptors play an important role of transporting M6P-bearing glycoproteins from the trans-Golgi network (TGN) to lysosomes mediated through their M6P binding sites.[2] Both receptors transport important enzymes to the intracellular acidic pre-lysosomal compartments where low pH leads to the release of the enzymes from the complex. The receptor then gets recycled into the Golgi apparatus.[3] However, only the M6P/IGFII receptor is anchored to the cell surface membrane and has been implicated in the internalization of M6P bearing compounds.[4] Importantly, it modulates the activity of a variety of extracellular M6P bearing glycoproteins including latent transforming growth factor-β (LTGFβ) precursor, urokinase-type plasminogen activator receptor, glycoprotein D of the herpes virus, granzyme B an essential factor for T cell-mediated apoptosis and proliferin.[4] This has resulted in an enormous interest in the design of M6P bearing compounds that target the M6P/IGFII receptor as it offers an efficient means for internalization of high specificity therapeutics.[5] This approach has been used to deliver therapeutic compounds in enzyme replacement therapies in lysosomal diseases like Fabry disease, aid wound healing, as a treatment for breast cancer, and to combat viral infections.[4] However, the approach suffers a major drawback as the phosphomonoester bond of M6P is prone to hydrolysis by various phosphatase enzymes.[6] This dramatically reduces its binding efficiency to the receptor thereby compromising its potency. This problem has been circumvented by Supporting information for this article is available on the WWW under http://dx.doi.org/10.1002/anie.201xxxxxx. Figure 1: a) Chemical structure of mannose-6-phosphate (M6P and the two analogues, b) Schematic representation of the cLogP of three compounds 1 the design of several isosteric M6P analogues with phosphonate, carboxylate or malonate groups, which have higher affinity to the receptor and a stronger stability in human serum than M6P.[7] This approach is successful in overcoming the issues with hydrolysis of the phosphomonoester bond, yet falls short as these analogues can only target the receptors present on the cell surface. In the steady state, ~90 % of the M6P/IGFII receptors are localized in the transmembrane compartments while the remainder stays on the cell surface.[8] The receptor has a relatively long half-life (t1/2 ~ 20 hours) and recycles between the trans-Golgi network, endosomes and the plasma membrane.[9] In this communication, we report a novel approach to improve ligand-receptor protein interaction in cells whilst overcoming stability issues associated with M6P. We demonstrate this by exploring a prodrug (analogue 2) that undergoes intracellular chemical modification by esterases to yield an active M6P analogue (analogue 1) (Figure 1a),[10] resulting in a sustained and focused therapeutic strategy in an in vitro model of wound healing. Table 1. Ligand-Receptor Protein interaction energies obtained for M6P and each of the two analogues in domain 3 and domain 5 as determined from 100 ns of molecular dynamics simulation. Two ligands were placed into the dimer binding pocket, because the receptor is secreted as a dimer. Domain 3 Ligand-Receptor Protein interaction Energy (kcal/mol) M6P Analogue 1 Analogue 2 Ligand 1 -368.4 -309.3 -81.6 Ligand 2 -347.4 -304.3 -79.6 Domain 5 Ligand-Receptor Protein interaction Energy (kcal/mol) M6P Analogue 1 Analogue 2 Ligand 1 -128.2 -44.1 -43.6 Ligand 2 -118.3 -74.0 -47.8 The design of phosphonate analogue 1 is based on established principles of bioisosteric M6P analogues by replacing the P-O bond at C6 by a methylene bridge. Moreover, the replacement of the hydroxyl group at the anomeric position by an aromatic subtituent slightly improves recognition by the M6P/IGFII receptor.[11] This could be due to the hydrophobic interactions between the aromatic moiety of analogue 1 and the binding pocket of the M6P/IGFII receptor. Previous studies have demonstrated that neutral ester prodrugs are relatively benign towards enzymatic degradation, thereby altering their apparent elimination and half-life.[12] Hence analogue 2 was designed by masking analogue 1 via esterification of the phosphate group to yield a non-charged bis(pivaloyloxymethyl) (POM) derivative. Importantly, derivatization of phosphates decreases the polarity of the parent drug thereby promoting its cellular internalization and altering the elimination/distribution mechanism.[12-13] Notably, the clogP (calculated-logP evaluated using Chem Draw) for M6P, analogue 1 and 2 are -3.28, 0.10 and 3.29 respectively (Figure 1b). LogP is an estimate of a compound's overall lipophilicity, a value that influences its physiological properties such as solubility, permeability through biological membranes, hepatic clearance, and non-specific toxicity.[14] Polar compounds with low logP have very low cellular permeability due to their low affinity for the lipid bilayers. Alternatively, lipophilic compounds with high logP have high affinity for the phospholipid phase facilitating their internalization and prohibiting their escape into the aqueous basolateral side.[14] Herein the lipophilic prodrug, Figure 2: Cell viability assays showing percentage of live cells in the culture post incubation with M6P, analogue 1 and 2. First and second column in each condition is representing 24 h and 72 h respectively. Data presented as average ± SEM (n=4). Significance was set at * p < 0.05 using bonferroni post-hoc test in one way ANOVA analysis. analogue 2, will have improved cellular internalization compared to its charged parent analogue, 1. Once internalized the bis(pivaloyloxymethyl) linkers of analogue 2 will be gradually prone to ester hydrolysis by microsomal esterases present within the intracellular compartments,[15] resulting in the conversion to the charged parent analogue, analogue 1. Analogue 1 on the contrary would only target extracellular M6P/IGFII receptors, when administered directly, due to its low cellular permeability deemed to its low logP value. The extracellular region of the M6P/IGFII receptor is comprised of 15 repetitive domains and contains three distinct M6P binding sites located in domains 3, 5, and 9, with only domain 5 exhibiting preference for phosphodiesters.[16] In order to assess our strategy to use the intracellular conversion of the produg analogue 2 to a high receptor binding phosphonate analogue 1, it is pivotal to examine the ligand-receptor interactions to validate the hypothesis that analogue 2 will have minimal interaction with the extracellular receptors. In the current study, we used six independent molecular dynamics simulations to study the ligand-receptor protein interactions of M6P, analogues 1 and 2 with domains 3 and 5 of the extracellular M6P/IGFII receptor (see Supporting Information for experimental details, section S8.1). Domain structures were adopted from previously reported studies and two ligands were placed into the dimer binding pocket, because the receptor is secreted as a dimer.[17] Analogue 1 showed similar ligand-receptor protein interaction energies to M6P in domain 3 (Table 1). Importantly, the m-xylene ring of analogue 1 was positioned in the middle of the binding pocket further stabilizing the binding of this compound in comparison to M6P (see Supporting Information, Figures S1 and S2). This is in accordance with the previous studies of other phosphonate analogues of M6P, which are reported to display higher affinity and stronger stability in human serum than M6P.[6, 7c] The domain 5 binding pocket is larger than in domain 3, hence all the Figure 3: Cell body area showing change in cell area post TGFβ1 stimulation and subsequent analogues treatment. Cell area was measured from the fluorescent images of live cells taken for viability assay (cells from minimum 40 images per group were measured). Significant increase in cell body area was observed for cells treated with TGFβ1 (2 ng/mL), however no such increase was observed in cells treated with analogues +/- TGFβ1 (2 ng/mL). Data presented as average ± SEM (n > 40). Significance was set at * p < 0.05 using bonferroni test in one way ANOVA. 2 compounds displayed weaker interactions with the receptor and occupied more diverse positions in domain 5 due to the increased space (see Supporting Information, Figure S2). Furthermore, in the case of analogue 1 in domain 5, the simulations suggested that one of the two analogue 1 ligands (ligand 1) bound to the protein dimer has weaker interactions with the protein as it primarily interacts with the second molecule of analogue 1 (ligand 2). Overall, the simulations suggested that analogue 1 has high affinity towards domain 3 similar to M6P whilst the prodrug 2 has weak interactions with both domains of the receptor (Table 1 and see Supporting Information, Figure S3). The molecular dynamics simulations further validated our aforementioned hypothesis that the prodrug will be internalized with minimal extracellular receptor-ligand interactions. We next validated our hypothesis in a well-established in vitro model for wound healing using primary human dermal skin fibroblasts (HDF). In mammals, wound healing is not a regenerative process that restores normal tissue architecture, but a reparative process that results in scar formation.[18] This process occurs in all tissues of the body in response to physical, chemical and biological stressors. Scar tissue is functionally and aesthetically inferior to normal tissue. It is a result of the excessive production of extracellular matrix (ECM) that occurs after injury.[19] One of the most important proteins influencing the ECM architecture during wound healing is collagen I. Collagen I is synthesized predominantly by fibroblasts and its synthesis is largely regulated by cytokine transforming growth factor β1 (TGFβ1).[20] TGFβ1 is secreted in an inactive form (LTGFβ1), requiring enzymatic conversion to active TGFβ1 to effect a change in cell function. One of the methods of TGFβ1 activation involves binding of M6P residues within the N-linked oligosaccharides on latent TGFβ1 to the M6P/IGFII receptor.[21] Since the M6P binding sites are involved in various steps of TGFβ1 activation and inactivation, it is believed that small molecule inhibitors that block the binding of M6P residues could present an opportunity to block the activity of TGFβ thereby reducing overproduction of an important profibrotic extracellular matrix protein collagen I. Cytotoxicity and cell viability of the analogues were initially assessed using MTS and live/dead assays (see experimental details in Supporting Information, sections S2.1, S3.1 and S4.1). Previous studies characterizing M6P binding affinity towards the M6P/IGFII receptor reported significant binding affinity at a concentration of 10 µM.[7b, 22] This concentration was therefore selected for our in vitro studies. All compounds showed no effect on cell viability and proliferation both in the presence and absence of TGFβ1 (Figure 2 and see Supporting Information, Figure S4 respectively). Exposure to TGFβ1 in the absence of analogues 1 and 2 resulted in a reduction in HDF proliferation (see Supporting Information, Figure S4). This growth suppressive response has been previously reported in many cell types.[23] The observed change in cell proliferation upon exposure to TGFβ1 influenced HDF cell morphology (and see Supporting Information, Figure S5b). Fibroblasts alter their morphology from dendritic to stellate upon exposure to various external cues caused by changes in actin polarisation and focal adhesion.[24] TGFβ1 has been shown to alter the morphology of many cell types including fibroblasts, potentially by inducing polymerisation of the actin cytoskeleton from globular to filamentous.[24a] Different factors such as cell motility and mechanical strain have also been reported to cause this alteration.[25] In the present case, we observed a reversal of HDF cell morphology back to initial cell morphology without TGFβ1 stimulation when treated with the analogues 1 and 2 (Figure 3 for quantification of cell body area and see Supporting Information, Figure S5c-e for Figure 4: Change in Collagen 1 a) mRNA levels and b) protein levels post TGFβ1 stimulation in the presence and absennce of M6P, analogues 1 and 2 compared to untreated (negative) control. Collagen I protein expression was normalised against β-actin levels. Data are presented as average ± SEM (n = 3). Significance was set at * p < 0.05 using bonferroni post-hoc test in one way ANOVA analysis images). We next assessed if the observed change in morphology is correlated to collagen I gene expression using qRT-PCR, and if changes in collagen I gene expression could be altered by inhibition of TGFβ1 activity by targeting the M6P/IGFII receptor in the presence of the analogues (refer to Supporting Information for method, section S5.1). Indeed as previously reported, exposure of HDF to TGFβ1 (2 ng/mL) resulted in a significant increase in collagen I mRNA expression at 48 hours post-stimulation.[20, 26] It is noteworthy that although collagen I gene expression was upregulated throughout the study period (72 hours), the optimal response was observed after 48 hours exposure to TGFβ1 (see Supporting Information, Figure S6). Therefore, the efficacy of the aformentioned compounds was assessed in the presence of TGFβ1 at 48 hours. TGFβ1 induced collagen I mRNA expression was downregulated significantly (p < 0.05) with the addition of prodrug analogue 2 (10 µM) with levels returning to that of normal untreated cells (Figure 4a). Downregulation was also observed for M6P however, the change did not reach stastistical significance (Figure 4a). This suggests that the variable responses that have been reported in the use of hydrolytically unstable M6P may be due to its realtive instability and that the development of stable analogues may resolve this issue. Importantly, in the present case we observed no significant change in collagen I gene expression in HDF cells treated with analogue 1 (Figure 4a). This was expected given the low cellular permeability which is believed to affect the ligand-receptor protein interactions in cells. Next, we investigated if the observed change at transcription level would have a corresponding influence on protein translation. Changes in collagen I protein expression were quantified using immunoblotting (refer to Supporting Information for method, section S6.1). All protein expression studies were carried out at 72 h post-stimulation. Significant upregulation in collagen I protein expression was observed post TGFβ1 stimulation (Figure 4b; column 2; p < 0.05) which is consistent with previous reports.[27] Analogue 2 (10 µM) was observed to reduce TGFβ1 3 [8] mediated upregulation of collagen I protein to non-stimulated levels (Figure 4b; column 5; p < 0.05). No significant changes were observed in the case of M6P or analogue 1. This further confirms that analogue 2 is a potent repressor of TGFβ1 induced collagen I synthesis and thus can ameliorate the profibrotic effects of TGFβ1 in human skin dermal fibroblasts. In summary, we have developed a novel approach using an intracellular prodrug of M6P, analogue 2, to target M6P receptors. This approach overcomes the physiological problems associated with the hydrolysis of M6P whilst successfully targeting the receptors using an intracellular coversion of the analogue. We believe that this approach of intracellular drug coversion for receptor targeting will have far reaching implications in the design of highly potent drug candidates for enzyme replacement therapies of lysosomal storage diseases, to aid wound healing and in cancer therapy. [9] [10] [11] [12] [13] [14] [15] [16] Received: ((will be filled in by the editorial staff)) Published online on ((will be filled in by the editorial staff)) Keywords: drug development • intracellular drug delivery • Mannose6-Phosphate analogue [1] [2] [3] [4] [5] [6] [7] M. Qian, D. E. Sleat, H. Zheng, D. Moore, P. Lobel, Mol. Cell. Proteomics 2008, 7, 58-70. P. Ghosh, N. M. Dahms, S. Kornfeld, Nat. Rev. Mol. Cell Biol. 2003, 4, 202-212. a) J. R. Duncan, S. Kornfeld, J. Cell Biol. 1988, 106, 617-628; b) M. Jin, G. G. Sahagian, M. D. Snider, J. Biol. Chem. 1989, 264, 7675-7680. M. Gary-Bobo, P. Nirde, A. Jeanjean, A. Morere, M. Garcia, Curr. Med. Chem. 2007, 14, 2945-2953. a) S. Hoogendoorn, G. H. van Puijvelde, J. Kuiper, G. A. van der Marel, H. S. Overkleeft, Angew. Chem. Int. Ed. 2014, 126, 11155-11158; b) O. Vaillant, K. E. Cheikh, D. Warther, D. Brevet, M. Maynadier, E. Bouffard, F. Salgues, A. Jeanjean, P. Puche, C. Mazerolles, P. Maillard, O. Mongin, M. Blanchard-Desce, L. Raehm, X. Rébillard, J.-O. Durand, M. Gary-Bobo, A. Morère, M. Garcia, Angew. Chem. Int. Ed. 2015, DOI: 10.1002/ange.201500286. S. Vidal, M. Garcia, J.-L. Montero, A. Morère, Bioorgan. Med. Chem. 2002, 10, 4051-4056. a) D. B. Berkowitz, G. Maiti, B. D. Charette, C. D. Dreis, R. G. MacDonald, Org. Lett. 2004, 6, 4921-4924; b) A. Jeanjean, M. Garcia, A. Leydet, J.-L. Montero, A. Morère, Bioorgan. Med. Chem. 2006, 14, 3575-3582; c) S. Vidal, C. Vidil, A. Morère, M. Garcia, J.-L. Montero, Eur. J. Org. Chem. 2000, 2000, 3433-3437. [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] a) E. C. Dell'Angelica, G. S. Payne, Cell 2001, 106, 395-398; b) H. M. El‐Shewy, L. M. Luttrell, Vitam. Horm. 2009, 80, 667-697. N. M. Dahms, L. J. Olson, J.-J. P. Kim, Glycobiology 2008, 18, 664-678. J. Zhang, M. G. Wong, M. Wong, S. Gross, J. Chen, C. Pollock, S. Saad, PloS one 2014, 10, e0116888-e0116888. J. J. Distler, J. F. Guo, G. W. Jourdian, O. P. Srivastava, O. Hindsgaul, J. Biol. Chem. 1991, 266, 21687-21692. J. P. Krise, V. J. Stella, Adv. Drug Delivery Rev. 1996, 19, 287310. S. J. Hecker, M. D. Erion, J. Med. Chem. 2008, 51, 2328-2345. M. J. Waring, Expert Opin. Drug Discov. 2010, 5, 235-248. a) U. Pradere, E. C. Garnier-Amblard, S. J. Coats, F. Amblard, R. F. Schinazi, Chem. Rev. 2014; b) W. N. Aldridge, Chem. Biol. Interact. 1993, 87, 5-13. a) L. J. Olson, N. M. Dahms, J.-J. P. Kim, J. Biol. Chem. 2004, 279, 34000-34009; b) L. J. Olson, R. D. Yammani, N. M. Dahms, J.-J. P. Kim, EMBO J. 2004, 23, 2019-2028; c) L. J. Olson, F. C. Peterson, A. Castonguay, R. N. Bohnsack, M. Kudo, R. R. Gotschall, W. M. Canfield, B. F. Volkman, N. M. Dahms, Proc. Natl. Acad. Sci. 2010, 107, 12493-12498. J. C. Byrd, R. G. MacDonald, J. Biol. Chem. 2000, 275, 1863818646. G. C. Gurtner, S. Werner, Y. Barrandon, M. T. Longaker, Nature 2008, 453, 314-321. M. Ferguson, Ulster Med. J. 1998, 67, 37. W. A. Border, E. Ruoslahti, J. Clin. Invest. 1992, 90, 1-7. P. A. Dennis, D. B. Rifkin, Proc. Natl. Acad. Sci. 1991, 88, 580584. a) A. Jeanjean, M. Gary-Bobo, P. Nirdé, S. Leiris, M. Garcia, A. Morère, Bioorg. Med. Chem. Lett. 2008, 18, 6240-6243; b) M. K. Christensen, M. Meldal, K. Bock, H. Cordes, S. Mouritsen, H. Elsner, J. Chem. Soc., Perkin Trans. 1 1994, 1299-1310. M. Laiho, J. A. DeCaprio, J. W. Ludlow, D. M. Livingston, J. Massague, Cell 1990, 62, 175-185. a) A. Moustakas, C. Stournaras, J. Cell Sci. 1999, 112, 11691179; b) E. Tamariz, F. Grinnell, Mol. Biol. Cell 2002, 13, 39153929. M. Chiquet, A. S. Renedo, F. Huber, M. Flück, Matrix Biol. 2003, 22, 73-80. B. Wang, R. Komers, R. Carew, C. E. Winbanks, B. Xu, M. Herman-Edelstein, P. Koh, M. Thomas, K. Jandeleit-Dahm, P. Gregorevic, M. E. Cooper, P. Kantharidis, J. Am. Soc. Nephrol. 2012, 23, 252-265. B. Wang, M. Herman-Edelstein, P. Koh, W. Burns, K. JandeleitDahm, A. Watson, M. Saleem, G. J. Goodall, S. M. Twigg, M. E. Cooper, P. Kantharidis, Diabetes 2010, 59, 1794-1802. 4 Entry for the Table of Contents (Please choose one layout) Layout 2: Intracellular Drug Delivery # Vipul Agarwal, Priyanka Toshniwal, # Natalie E Smith, Nicole M Smith, Binbin Li, Tristan D Clemons, Lindsay T Byrne, Foteini Hassiotou, Fiona M Wood, Mark Fear, Ben Corry* and K Swaminathan Iyer* __________ Page – Page Enhancing the Efficacy of CationIndependent Mannose 6-Phosphate Receptor Inhibitors by Intracellular Delivery Intracellular delivery of M6P/IGFII receptor inhibitors exhibits better efficacy than extracellular inhibitors to regulate TGFβ1 mediated upregulation of profibrotic marker, collagen I. 5 Journal Name RSCPublishing COMMUNICATION Cite this: DOI: 10.1039/x0xx00000x Inhibiting the activation of transforming growth factor-β using a polymeric nanofiber scaffold Vipul Agarwal,a Fiona M. Wood,b,c Mark Fearb and K. Swaminathan Iyer*a Received 00th January 2012, Accepted 00th January 2012 DOI: 10.1039/x0xx00000x www.rsc.org/ Electrospun poly(glycidyl-methacrylate) (PGMA) nanofibers were fabricated in the presence of a hydrophobic analogue of mannose-6-phosphate, (PXS64). The nanofibers were tested for biocompatibility as a tissue engineering scaffold and for their efficacy to inhibit transforming growth factor β1 (TGFβ1) activation in human dermal skin fibroblasts. The principle of tissue engineering involves the development of advanced functional biomaterials that incorporate biochemical cues to aid in the regeneration of tissues to restore and maintain normal organ function.1 This approach has resulted in the fabrication of several advanced biomaterial platforms that have been used to repair damaged or diseased tissues and to create therapeutic approaches for entire tissue replacement.2-4 While tissue regeneration has successfully been demonstrated in the presence of biocompatible scaffolds, scarring remains one of the unresolved issues. In mammals postnatal wound healing results in scar formation, characterised by excessive collagen deposition and dysfunctional extracellular matrix formation. This is also a hallmark of fibrotic disease which occurs in many tissues. Central to wound repair is transforming growth factor β1 (TGFβ1), a cytokine secreted by several different cell types involved in wound healing.5 In the case of skin, scarring during the wound healing process is a result of TGFβ1 mediated imbalance in the fibroblast activity resulting in an architecturally disorganised extracellular matrix.6 One of the most important proteins influencing the ECM architecture during wound healing is collagen I. Collagen I is synthesized predominantly by fibroblasts and its synthesis is largely regulated by TGFβ1 signalling.7 Following wound healing in skin, a scar is not only aesthetically and psychologically detrimental but can also cause functional disability and pain.8 Importantly, inhibition of TGFβ1 activity has been documented to reduce scar formation9 whereas subcutaneous delivery of exogenous TGFβ1 in newborn mice was shown to promote fibrosis and angiogenesis at the site of injection.10 During wound healing TGFβ1 activation from its latent state involves binding of the mannose 6-phosphate (M6P) residues within This journal is © The Royal Society of Chemistry 2012 the N-linked oligosaccharides on the latent TGFβ to the M6P/IGFII receptor.11 Since the M6P binding sites are involved in various steps of the TGFβ1 activation and inactivation route, it is believed that the presence of small molecule analogues of M6P that competitively bind to the receptor could present an opportunity to block the activation of TGFβ1 thereby reducing overproduction of extracellular matrix protein collagen and potentially reducing scarring.12 The major drawback with using M6P as a drug in the reduction of scarring is that the phosphomonoester bond of M6P is prone to hydrolysis by various phosphatase enzymes.13 Isosteric M6P analogues with phosphonate, carboxylate or malonate groups have been shown to circumvent the aforementioned issues and have been reported to have greater stability in human serum than M6P.14-17 The analogue PXS64 reported in the present study is a lipophilic bioisosteric phosphonate analogue developed by [(bis(pivaloyloxymethyl)) (POM)] ester derivatization of M6P. Importantly, the high lipophilicity of PXS64 limits its solubility in aqueous solutions thereby reducing its bioavailability.18 In this communication we report that PXS64 can be incorporated in an electrospun poly(glycidyl methacrylate) (PGMA) nanofibrous scaffold. Furthermore, we demonstrate the utility of this scaffold as a biomaterial platform for wound healing using human dermal skin fibroblasts (HDF). We demonstrate its effectiveness as a drug delivery platform to mitigate TGFβ1 mediated upregulation of collagen I. Electrospinning is a widely used technique to fabricate large area nanofibrous scaffolds19 mimicking the architecture of extracellular matrix (ECM).20 They have been used as tissue engineering scaffolds to promote cell growth and migration and to achieve controlled delivery of drugs and growth factors.21 They have been widely used in skin, bone, cartilage, vascular and neural tissue engineering. 22 For example, Yang et al. employed emulsion electrospinning to fabricate ultrafine core sheath poly(ethylene glycol)-poly(D,L-lactide) fibers loaded with basic fibroblast growth factor (bFGF) and demonstrated its gradual release over 4 weeks.23 In vitro studies on mouse embryonic fibroblasts showed enhanced cell adhesion, proliferation and secretion of ECM when cultured on the nanofibrous scaffold.24 J. Name., 2012, 00, 1-3 | 1 COMMUNICATION Journal Name Figure 2: a) Cell viability assay showing percentage of live cells in culture post incubation on ES-PGMA and ES-PGMA + PXS64, in the presence and absence of TGFβ1 (2 ng/mL), b) and c) showing HDF cell morphology on ES-PGMA + PXS64 + TGFβ1 (2 ng/mL) using fluorescent microscopy and scanning electron microscopy respectively. Red arrows highlighting the cell attachment and adhere points on the fibers. Scale bars: b) and c) 2 µm. Figure 1: SEM secondary electron image of electrospun fibers of a) PGMA, b) PGMA + PXS64, C) showing the chemical structure of PXS64. Scale bars: a) and b) 10 µm. ‘‘grafting to’’ interactions mediated by the epoxy groups.27, 28 To this end, poly(glycidyl methacrylate) (PGMA) used in the current study, has an epoxy group in every repeating unit and has been used extensively as a macromolecular anchoring layer.29, 30 In the present study, PGMA was electrospun in the presence of PXS64 and in the absence of PXS64 as a control (see supporting information for experimental conditions). PXS64 loading in the electrospun PGMA fibers was measured using high pressure liquid chromatography (HPLC) to be ~76 % w/w. Electrospun fibrous scaffolds were characterized using scanning electron microscopy (SEM). Fiber diameter was measured from the SEM images to be 0.69 ± 0.31 µm (average ± standard deviation) for ES-PGMA and 2.24 ± 1.06 µm in the case of ES-PGMA + PXS64 (Figure 1). Biocompatibility of the scaffold was investigated using the colorimetric MTS assay and live/dead cell viability assay (see Supporting Information for method) and were imaged using fluorescence and scanning electron microscopy.31 Human dermal fibroblasts (HDF) cells were cultured on the scaffolds and incubated In the in vivo studies on dorsal wound model in diabetic rats, bFGF/ poly(ethylene glycol)-poly(D,L-lactide) mats showed elevated healing with complete re-epithelialization and regeneration of skin appendages such as hair and sebum, while bFGF promoted collagen deposition and its remodelling similar to normal architecture.23 In the case of electrospun scaffolds the ability to modify the surface properties of the nanofibers such as adhesion, wettability and biocompatibility is pivotal in its integration as a tissue engineering platform. Currently one of the biggest challenges using this technique to produce scaffolds is that the polymers used thus far to develop nanofibers need specific approaches for surface modification that are complex and laborious.25 The achievement of a certain degree of grafting universality requires the establishment of a controlled method of introducing the desired functional groups. Currently, this is achieved by physisorption.26 In contrast, chemisorption which is difficult to achieve on polymeric nanofibers would result in covalent attachment. Polymers containing epoxy groups are examples of functional polymers that are able to react with a wide range of biologically relevant molecules through Figure 3: Collagen I gene expression analysis showing the percentage change in gene expression as compared to non-treated negative control (column 1). HDF cells incubated on both scaffolds and plastic tissue culture plate both in presence and absence of TGFβ1 and release media. Data presented as average ± SEM (n = 3). Significance was set at * p < 0.05 using bonferroni post-hoc test in one way ANOVA analysis. 2 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 2012 Journal Name both with and without TGFβ1 over a period of 72 hours. No significant changes were observed on cell proliferation (see Supporting Information; Figure S1) or cell viability (Figure 2a) on both scaffolds with or without TGFβ1. To study the cell-matrix interactions, scaffolds were incubated with HDF cells and imaged using fluorescence and electron microscopy (see Supporting Information for method). HDF cells were observed to adopt their characteristic spindle shape morphology on both ES-PGMA and ESPGMA + PXS64 scaffolds in the presence and absence of TGFβ1 (Figure 2b, c and see Supporting Information; Figure S2 and S3). Finally, the ability of ES-PGMA + PXS64 scaffold in regulating the over expression of collagen I gene in HDF cells post TGFβ1 stimulation was evaluated using RT-qPCR (see Supporting Information). HDF cells cultured on the biocompatible ES-PGMA control scaffold showed significant increase in collagen I gene expression in the presence of TGFβ1 (Figure 3). Cells incubated on ES-PGMA + PXS64 scaffold under similar conditions showed reduced expression of collagen I gene compared to ES-PGMA scaffold, which reached significance in the presence of supplemented release media. The use of DMSO in release media was adopted from previously reported methodology32 to accelerate the release of hydrophobic drugs in vitro and was shown to have no effect on collagen expression alone (Figure 3). Here we developed a PGMA scaffold and encapsulated an antiscarring drug, PXS64. The biocompatibility of the scaffold was demonstrated in human dermal skin fibroblasts. Finally, the efficacy of scaffold and drug in mitigating the increased expression of collagen I in response to TGFβ1 stimulation was also demonstrated. We believe that this potential proof of principle approach can easily be adapted in the design of scaffolds for tissue regeneration in the presence of other antifibrotic agents. The authors would like to thank the Australian Research Council for funding the work. The authors would also like to acknowledge the Australian Microscopy & Microanalysis Research Facility at the Centre for Microscopy, Characterization & Analysis, The University of Western Australia, funded by the University, State and Commonwealth Governments. The authors would like to thank Dr Foteini Hassiotou to kindly provide the RT-qPCR instrument and Pharmaxis Ltd, Sydney for providing PXS64. MF is supported by Chevron Australia. This work was partly funded by the Fiona Wood Foundation. Notes and references a School of Chemistry and Biochemistry, The University of Western Australia, Crawley, Western Australia, Australia. E-mail: swaminatha.iyer@uwa.edu.au b Burn Injury Research Unit, School of Surgery, The University of Western Australia, Crawley, Western Australia, Australia. c Burns Service Western Australia, Department of Health, Perth, Western Australia, Australia † Footnotes should appear here. These might include comments relevant to but not central to the matter under discussion, limited experimental and spectral data, and crystallographic data. Electronic Supplementary Information (ESI) available: [details of any supplementary information available should be included here]. See DOI: 10.1039/c000000x/ 1. T. Dvir, B. P. Timko, D. S. Kohane and R. Langer, Nat. Nanotechnol., 2011, 6, 13-22. This journal is © The Royal Society of Chemistry 2012 COMMUNICATION 2. S. Van Vlierberghe, P. Dubruel and E. Schacht, Biomacromolecules, 2011, 12, 1387-1408. 3. D. I. Braghirolli, D. Steffens and P. Pranke, Drug Discovery Today, 2014, 19, 743-753. 4. C. Huang, S. J. Soenen, J. Rejman, B. Lucas, K. Braeckmans, J. Demeester and S. C. De Smedt, Chem. Soc. Rev., 2011, 40, 2417-2434. 5. A. Biernacka, M. Dobaczewski and N. G. Frangogiannis, Growth Factors, 2011, 29, 196-202. 6. M. Walraven, M. Gouverneur, E. Middelkoop, R. H. J. Beelen and M. M. W. Ulrich, Wound Repair Regen., 2014, 22, 3-13. 7. C. C. Yates, P. Hebda and A. Wells, Birth Defects Res. C Embryo Today, 2012, 96, 325-333. 8. A. Bayat, D. A. McGrouther and M. W. J. Ferguson, The BMJ, 2003, 326, 88-92. 9. G. G. Gauglitz, H. C. Korting, T. Pavicic, T. Ruzicka and M. G. Jeschke, Mol. Med., 2011, 17, 113. 10. A. B. Roberts, M. B. Sporn, R. K. Assoian, J. M. Smith, N. S. Roche, L. M. Wakefield, U. I. Heine, L. A. Liotta, V. Falanga and J. H. Kehrl, Proc. Natl. Acad. Sci., 1986, 83, 4167-4171. 11. P. A. Dennis and D. B. Rifkin, Proc. Natl. Acad. Sci., 1991, 88, 580584. 12. J. Chamberlain, Cardiovascular drug reviews, 2001, 19, 329-344. 13. A. Jeanjean, M. Garcia, A. Leydet, J.-L. Montero and A. Morère, Bioorgan. Med. Chem., 2006, 14, 3575-3582. 14. S. Vidal, C. Vidil, A. Morère, M. Garcia and J.-L. Montero, Eur. J. Org. Chem., 2000, 2000, 3433-3437. 15. S. Vidal, M. Garcia, J.-L. Montero and A. Morère, Bioorgan. Med. Chem., 2002, 10, 4051-4056. 16. C. Vidil, A. Morère, M. Garcia, V. Barragan, B. Hamdaoui, H. Rochefort and J.-L. Montero, Eur. J. Org. Chem., 1999, 1999, 447-450. 17. D. B. Berkowitz, G. Maiti, B. D. Charette, C. D. Dreis and R. G. MacDonald, Org. Lett., 2004, 6, 4921-4924. 18. J. Zhang, M. G. Wong, M. Wong, S. Gross, J. Chen, C. Pollock and S. Saad, PloS one, 2014, 10, e0116888-e0116888. 19. M. M. Hohman, M. Shin, G. Rutledge and M. P. Brenner, Phys. Fluids, 2001, 13, 2201-2220. 20. A. Hasan, A. Memic, N. Annabi, M. Hossain, A. Paul, M. R. Dokmeci, F. Dehghani and A. Khademhosseini, Acta Biomater., 2014, 10, 11-25. 21. H. Liu, X. Ding, G. Zhou, P. Li, X. Wei and Y. Fan, J Nanomater., 2013, 2013, 3. 22. V. Jayarama Reddy, S. Radhakrishnan, R. Ravichandran, S. Mukherjee, R. Balamurugan, S. Sundarrajan and S. Ramakrishna, Wound Repair Regen., 2013, 21, 1-16. 23. Y. Yang, T. Xia, W. Zhi, L. Wei, J. Weng, C. Zhang and X. Li, Biomaterials, 2011, 32, 4243-4254. 24. N. Abbasi, S. Soudi, N. Hayati-Roodbari, M. Dodel and M. Soleimani, Cell J., 2014, 16, 245. 25. H. S. Yoo, T. G. Kim and T. G. Park, Adv. Drug Delivery Rev., 2009, 61, 1033-1042. 26. S. Edmondson, V. L. Osborne and W. T. S. Huck, Chem. Soc. Rev., 2004, 33, 14-22. 27. K. S. Iyer, B. Zdyrko, H. Malz, J. Pionteck and I. Luzinov, Macromolecules, 2003, 36, 6519-6526. J. Name., 2012, 00, 1-3 | 3 COMMUNICATION Journal Name 28. B. Zdyrko and I. Luzinov, Macromol. Rapid Commun., 2011, 32, 859-869. 29. B. Zdyrko, K. Swaminatha Iyer and I. Luzinov, Polymer, 2006, 47, 272-279. 30. A. Sidorenko, T. Krupenkin, A. Taylor, P. Fratzl and J. Aizenberg, Science, 2007, 315, 487-490. 31. T. Mosmann, J. Immunol. Methods, 1983, 65, 55-63. 32. T. Ur-Rehman, S. Tavelin and G. Gröbner, Int. J. Pharm., 2010, 394, 92-98. 4 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 2012 Chapter 4 Conclusion and Future Works In this thesis two complementary approaches were investigated to address the problem of scarring after injury. The first part of the work focused on a tissue engineering approach towards promotion of skin regeneration and subsequent drug delivery application using a nanoscaffold while the latter part focused on the use of novel anti-scarring therapeutics. The tissue engineering approach involved the development of a multifunctional scaffold which could have a number of different applications in addition to a potential drug-delivery scaffold including upconversion probes, magneto-responsive material and biosensor. Subsequently, two analogues of the natural anti-scarring molecule mannose-6-phosphate (analogue 1 and 2) were investigated as potential scar modulators. Finally, the two concepts were blended to fabricate a nanofibrous scaffold for the delivery of analogue 2 in scar therapy. In this section, a summary of the work detailed in the publications will be presented. 4.1 Tissue engineering of a nanoscaffold Tissue engineering has evolved as an innovative approach towards tissue regeneration which usually relies on a platform to drive the cellular response. Electrospinning was employed to fabricate a stable nanofibrous universal grafting substrate using a polyglycidyl methacrylate (ES-PGMA) polymer. PGMA is an interesting polymer because of the presence of an ester linkage and highly reactive epoxy ring in each monomer unit. Presence of the epoxy ring provides a thermally induced self-crosslinking capability to the scaffold which eliminates the need for external chemical or physical crosslinking methods. 71 Multi-functionality of ES-PGMA was demonstrated first by end grafting it with carboxy-terminated poly N-isopropyl acrylamide (PNIPAM-COOH) to design a thermo-responsive scaffold. Contact angle measurements confirmed the surface grafting of ES-PGMA with PNIPAM-COOH which was demonstrated to undergo phase transition at the glass transition temperature of PNIPAM. Next, ESPGMA/nanoparticle composite scaffolds using three different types of nanoparticles ((NaGdF4:Yb, Er); Pd and Fe3O4) were generated using the electrospinning protocol established above. Nanoparticle loading in the case of hybrid scaffolds was confirmed by x-ray microanalysis, TEM, and ICP and AAS analysis. SEM analysis however revealed an increase in the fiber diameters in the hybrid materials. Changes in the solvent conductivity and surface tension have been claimed to have major influence on fiber diameter. One of the major problems encountered during the electrospinning process caused by the low vapour pressure of solvent used was the coagulation and drying of Taylor’s cone at the tip of the needle which required periodic cleaning and resulted in loss of material. Solvents with relatively higher vapour pressure such as THF can be used to improve the electrospinning process and address the above mentioned problem. Applicability of these hybrid materials were examined in gas sensing for Pd, upconversion imaging in the case of upconverting nanoparticles and magnetoresponsiveness for magnetite nanoparticles. Optimisation of the hydrogen gas sensing measurement was difficult as the Pd concentration on the surface of the matrix was unknown. Fiber porosity also limited the sensitivity to high concentrations of hydrogen. Potentially, surface functionalization of the scaffold with Pd nanoparticles, in addition to encapsulation, could be attempted to elevate the sensitivity of the system towards lower concentrations of hydrogen gas. SQUID analysis on the magnetite/polymer composite showed superparamagnetic behaviour of the matrix making it ideal for applications in protective fabric material and currency notes.304 72 4.2 Scar therapy and mannose-6-phosphate analogues The M6P/IGFII receptor is important in trafficking of cargoes from the cell surface and trans-golgi network to lysosomes.305-307 It plays an important role in various diseases including fibrosis and lysosomal storage disorders.307 The trafficking function of the receptor is exploited mainly for enzyme replacement therapy but has not been a focus of anti-fibrotic therapeutic approaches 308 M6P is a natural occurring ligand of the M6P/IGFII receptor found in a number of glycoproteins. The M6P/IGFII receptor consists of 15 repeating domains out of which domains 3, 5 and 9 are specific towards M6P binding. With the emergence of the anti-scarring properties of M6P309 and the structural determination of the M6P/IGFII receptor,305 a great deal of research has focused on the development of more efficient analogues of M6P.155, 156 This is because the therapeutic potential of M6P is limited because of its low metabolic stability against phosphatases and difficulty in retaining high concentrations at the injury site. Research has mainly focussed on manipulating the structure of M6P in order to enhance its binding efficiency towards the receptor while improving metabolic stability.157, 310, 311 Bioisosteric phosphonate analogues have been shown to have greater serum stability than M6P. Taking advantage of this, two bioisosteric phosphonate analogues, 1 and 2, were explored for their potential as anti-scarring agents in this thesis. Analogue 1 was rationalised to have higher metabolic stability than M6P, thereby enhancing its bioavailability, without lowering its affinity for the receptor. Molecular dynamics simulation studies confirmed that analogue 1 has a similar interaction towards the receptor as M6P (carried out in collaboration with Corry group). However, intercellular delivery of analogue 1 showed no significant response in impeding the expression of collagen I gene, the model used to study anti-scarring potential of the analogues. Neutral lipophilic phosphate compounds have been reported to undergo rapid internalisation increasing their local concentration within the cellular microenvironment.312, 313 Therefore, analogue 2 was designed as a neutral molecule with high logP to achieve intracellular drug delivery. Analogue 1 was masked by esterification of the phosphate group to yield a non-charged bis(pivaloyloxymethyl) 73 derivative, analogue 2. Upon internalisation, phosphoester linkers undergo ester hydrolysis by esterases present within the cellular microenvironment and release analogue 1, allowing intracellular targeting of the M6P/IGFII receptor. Gene expression studies showed significant reduction in collagen I expression, similar to treatment with M6P, upon analogue 2 treatment. Therefore, it was concluded from this part of the study that intracellular targeting of the receptor is effective using analogue 2 as a prodrug of analogue 1. One of the major problems encountered was in the analysis of protein expression. Analysis of collagen I and fibronectin expression was attempted to supplement the data on RNA expression levels. Immunoblotting was unreliable and had low reproducibility, most likely in part due to the large molecular weights of the target proteins and ineffective protein transfer. Alternative techniques such as HPLC and a recently developed ‘scar-in-jar’ protocol may be more effective methods for the quantitation of protein expression, especially for collagen I.314, 315 Despite the positive response in reducing the effects of TGFβ1, analogue 2 has limitations. It is vulnerable to metabolic degradation mediated by esterases present in growth serum. In addition, one of the hydrolysis products is formaldehyde, which is toxic to cells at high concentrations limiting the dose that can be administered. 4.3 Scaffold based delivery of analogue 2 To achieve sustained delivery of the effective analogue 2, the electrospun PGMA nanofibrous platform was manipulated for a drug delivery application. It was anticipated that enzymatic vulnerability of analogue 2 could be subdued by encapsulating it within the polymer scaffold by limiting its direct exposure to serum. Using the parameters established above, analogue 2 was electrospun with PGMA to yield a fibrous drug delivery platform with loading efficiency of ~76 % as measured by HPLC. Despite the high drug loading its subsequent release from the fibers could not be quantified. It was postulated that either the rate of drug release was below the limit of detection of the instrument or there is no drug release from the fibers. In an 74 attempt to see if any cellular response could be detected, loaded scaffold was incubated with HDF cells. Cyto-compatibility of the scaffold was confirmed by MTS and a fluorescent staining live/dead assay. Interestingly, efficacy of the released drug from the fibers was evident, in the presence of DMSO (facilitating the release of drug from the fibers), from the collagen I gene expression studies which showed a similar significant response as free drug in curtailing the upregulated expression of collagen I post TGFβ1 stimulation. Therefore, it can be concluded that slow sustained release of analogue 2 can be accomplished using electrospun PGMA scaffold. 4.4 Future recommendations There are still several questions that need to be addressed to translate this preliminary work into the clinic. For example, the lack of biodegradability of the PGMA scaffold, the serum stability of the drug and its release kinetics all need further development. The general strategy adopted to induce biodegradability is by copolymerising nondegradable polymers with another biodegradable polymer. Polymers have also been copolymerised with peptides and proteins to not only improve biodegradability but also to provide biologically active functional cues for their integration into the matrix. For example, the RGD (arg-gly-asp) sequence peptide motif can be copolymerised on PGMA using atom transfer living radical polymerisation (ATRP). The RGD motif is strongly recognised by a subset of integrins which are protein receptors necessary for cell adhesion and migration.316-318 Wound healing is a combination of cellular proliferation and migration working in tandem. Copolymers with RGD motifs are shown to promote cell migration. In addition, PGMA scaffolds can be covalently functionalised with growth factors such as insulin like growth factor I (IGF I) and epidermal growth factor (EGF) via an SN2 substitution ring opening reaction. The ester groups present in the PGMA backbone can undergo esterase mediated ester hydrolysis to sustainably deliver growth factors. Both IGF I and EGF are growth inducing proteins which play pivotal roles in wound healing.319 Chemical conjugation of the growth factor to the scaffold allows their passive 75 diffusion which is continued over the life-time of the scaffold.320 Incorporation of growth factors and the RGD peptide sequence in the electrospun PGMA matrix could elevate the rate of healing by simultaneously promoting cell growth and migration. Electrospun scaffolds have found clinical applications because of their propensity to mimic extracellular matrix thereby providing a suitable environment to promote regeneration. Matrix based delivery of growth factors are preferred over their exogenous delivery because it protects the growth factor from degradation by limiting their exposure to the proteolytic wound environment.321 Despite the efficacy of analogue 2 ([(bis(pivaloyloxymethyl) (POM)] derivative of analogue 1) in reducing the expression of collagen I gene after TGFβ1 treatment, its structural integrity in serum limits it’s in vivo translational potential. [Bis(S-acyl-2thioethyl) (SATE)] derivatives of phosphonate analogues have been shown to possess higher enzymatic stability against serum and gastric juices as compared to POM derivatives.313, 322 Benzaria et al. synthesised SATE derivatives of 9-[2- (Phosphonomethoxy)ethyl]adenine (PMEA) and studied their stability in vitro in comparison to bis[(pivaloyloxymethyl) (POM)]- and bis[dithiodiethyl (DTE)]PMEA analogues.322 They concluded that although bis(POM)- and bis(tBu-SATE)PMEA had similar anti-viral activity, bis(tBu-SATE)PMEA had greater stability than bis(POM)PMEA in human gastric juice and human serum. It would potentially be beneficial if the replacement of the POM functionality by a SATE derivative in analogue 2 could improve its serum stability whilst maintaining its functional efficacy. This could lead to oral or intravenous delivery of the drug. However, SATE as a linker has its own limitations, with one of the degradation products, ethylene sulphide, inducing toxicity at high concentrations in cells. Analogue 2 can be modified at the anomeric centre by replacing the lipophilic group with a fluorophore which would be beneficial in visualising its internalisation within the cell microenvironment which can be tracked using confocal microscopy. This will also allow another platform to study drug release profile from the scaffold. Further, in a more combinatorial approach, analogue 2 could be combined with a lysyl oxidase inhibitor. While analogue 2 works predominantly on controlling the expression of collagen, lysyl oxidase inhibitors reduce collagen crosslinking, making collagen more prone to degradation. A combined approach could be more effective 76 in reducing fibrosis than the use of a single compound targeting one part of the fibrotic pathway. Incorporation of such approaches within the functionalised scaffold would address the shortfalls with both the scaffold and analogue 2 and could lead to pre-clinical assessments. The efficacy of the system can be studied on an in vivo porcine wound healing model. Porcine is a well-established wound healing model which has similar characteristics to humans i.e. porcine heals with scars unlike rodents which heal more via contraction instead of granulation tissue formation. The developed scaffold could potentially be used as a dressing, incorporating growth factors and/or other anti-scarring agents to be applied directly at the wound site. Scaffold mediated delivery of the anti-scarring agents would need to be compared against free drug to confirm the efficacy of this combinatorial approach. 4.5 Final remarks The three overarching aims of the work compiled in this thesis are set out below: 1. Optimise the parameters for electrospinning PGMA and study its efficacy as a biocompatible universal polymeric scaffold 2. Investigate the biocompatibility and potential of M6P analogues as anti- scarring agents in an in vitro model using human dermal skin fibroblasts 3. Investigate the potential of a combinatorial scaffold:drug approach to limit fibrosis The data presented within this thesis provides evidence of successful incorporation of tissue engineering and drug delivery approaches to tackle the problem of skin scarring. The scaffold developed by electrospinning PGMA can be used in various other applications, some of which has been demonstrated in this thesis. Two bioisosteric phosphonate analogues were investigated as potential anti-fibrotic treatments. Analogue 2, a POM derivative of analogue 1, was shown to function as a prodrug. Its mode of action was substantiated to be intracellular where its POM 77 linkers get hydrolysed to produce the active analogue 1. Therefore, it was demonstrated that intracellular targeting of M6P/IGFII receptor using this analogue is a potentially effective approach to inhibit TGFβ1 signalling and therefore reduce fibrosis. In a tissue engineering approach, analogue 2 was co-electrospun with PGMA to fabricate a fibrous scaffold. The biocompatibility of the scaffold was confirmed in a human dermal skin fibroblast in vitro model. Scaffold mediated delivery of analogue 2 was shown to regulate collagen I expression. This is the first example of using this tissue engineering approach in this context and has the potential with further work to address some of the major limitations with current treatments aimed at reducing fibrosis. 78 Appendix A Supporting information for papers 1. Agarwal, V., Ho, D., Ho, D., Galabura, Y., Yasin, F. M.D., Gong, P., Ye, W., Singh, R., Munshi, A., Saunders, M., Woodward, R. C., St. Pierre, T., Wood, F.M., Fear, M., Lorenser, D., Sampson, D. D., Zdyrko, B., Smith, N.M., Luzinov, I., Iyer, K.S., A Functional Reactive Polymer Nanofiber Matrix, RSC Advances (Submitted) 2. Agarwal, V., Toshniwal, P., Smith N. E., Smith, N. M., Li, B., Clemons, T. D., Byrne, L. T., Hassiotou, F., Wood, F. M., Fear, M., Corry, B., Iyer, K.S., Intracellular Enhancing the Efficacy of Cation-Independent Mannose 6-Phosphate Receptor Inhibitors by Intracellular Delivery, Angewandte Chemie International Edition (submitted) 3. Agarwal, V., Wood, F. M., Fear, M. and Iyer, K. S., Inhibiting the activation of transforming growth factor-β using a polymeric nanofiber scaffold, Nanoscale (Submitted) 79 Supporting Information A Functional Reactive Polymer Nanofiber Matrix Vipul Agarwal,a Dominic Ho,a Diwei Ho,a Yuriy Galabura,b Faizah M.D. Yasin,a Peijun Gong,c Weike Ye,d Ruhani Singh,a Alaa Munshi,a Martin Saunders,e Robert C. Woodward,f Timothy St. Pierre,f Fiona M. Wood,g Mark Fear,g Dirk Lorenser,c David D. Sampson,c, e Bogdan Zdyrkob, Nicole M. Smith,a Igor Luzinovb, *, and K. Swaminathan Iyera,* Materials Polyglycidyl methacrylate (PGMA) with Mn = 220515 and Mw = 433730 was synthesized by radical polymerisation as reported previously,1 Carboxy terminated N-isopropyl acrylamide (PNIPAM-COOH) (Polymer source, Cat. # P5589, Mn = 42000). Methyl ethyl ketone (MEK, Merck), iron (II) acetylacetonate, tetradecanediol, oleic acid, oleylamine, dibenzyl ether, 1- octadecene and palladium (II) acetylacetonate purchased from Sigma Aldrich. Gadolinium chloride, ytterbium chloride and erbium chloride purchased from GFS chemicals. All other chemicals were purchased from Sigma-Aldrich. All chemicals used were of analytical grade purity. Methods Electrospinning Procedure: PGMA (15 w%) was dissolved in MEK (10 mL) overnight at room temperature with constant stirring. In the case of composites, metal nanoparticles resuspended in MEK were added (magnetite – 35.7 mg, palladium - 40 mg, Upconverting particles - 23.6 mg) to maintain PGMA concentration at 15 w% the next day and further stirred for 1h to homogenise the polymer solution. 80 PGMA polymer and polymer composite fibers were obtained via electrospinning (ESPGMA) (Nanofiber Electrospinning Unit, Cat. # NEU-010, Kato Tech, Japan). The electrospinning parameters after optimisation were a voltage of 9.1 kV, working distance of 9 cm, syringe pump speed of 0.04 mm/min (1 mL/h). Fibers were annealed at 80 °C for 5 hours post electrospinning. Nanoparticle loading in the ES-PGMA was determined by ICP-MS and AAS analysis (in the case of the magnetite/ES-PGMA composite). Magnetite Synthesis: Magnetite nanoparticles were synthesised using the thermal decomposition method as described previously.2 Briefly, iron (II) acetylacetonate {Fe(acac)3} (1 mmol), tetradecanediol (5 mmol), oleic acid (3 mmol) and oleylamine (3 mmol) were dissolved in dibenzyl ether. The reaction mixture was heated under a N2 atmosphere at 100 °C for 30 min, 200 °C for 2h and further refluxed at 300 °C for another 1h. The black coloured product was collected via precipitation and washed with 3 x ethanol by centrifugation at 4000 rpm and dispersed in hexane. Nanoparticles were stored under an inert atmosphere. Palladium Nanoparticles Synthesis: Palladium nanoparticles were synthesised as per the method described.3 Briefly, palladium(II) acetylacetonate {Pd(acac)3} (150 mg) was mixed with oleylamine (246 µL) in toluene (61.5 mL) and stirred vigorously for 10 min at room temperature yielding a yellow coloured reaction mixture. To this formaldehyde (300 µL) was added and further reacted for 10 min. The reaction was carried out for a further 8h at 100 °C and the colour changed to black, indicating the completion of the reaction. The product was brought to 81 room temperature, washed with 3 x ethanol and resuspended in chloroform. Nanoparticles were stored under an inert atmosphere. NaGdF4: Yb, Er Upconverting Nanoparticle Synthesis: NaGdF4:Yb, Er upconverting nanoparticles were synthesised as per the method described.4 Briefly, GdCl3.6H2O (0.80 mmol), YbCl3.6H2O (0.18 mmol) and ErCl3.6H2O (0.02 mmol) were added to the solution mixture of oleic acid (14 mL) and 1- octadecene (16 mL) and homogenized under N2 while heated to 150 °C. The solution was then cooled to 50 °C, methanol (10 mL) containing NaOH (2.5 mmol) and NH4F (4 mmol) was then added slowly and reacted for another 30 min. Next, methanol was removed by heating the reaction mixture at 100 °C under vacuum for 10 min. Under atmospheric pressure, the reaction temperature was raised to 300 °C and the reaction was carried out for 1h under a N2 atmosphere. The reaction was terminated by cooling to room temperature. The product was precipitated in ethanol, collected by centrifugation and washed with ethanol. Finally, the product was resuspended in THF. PNIPAM- COOH attachment on ES-PGMA: The PGMA polymer was electrospun onto 12 mm diameter glass coverslips (Cat. # G40112, ProSciTech) and annealed at 80 °C for 5h. Dried polymer was then exposed to carboxy terminated poly N-isopropyl acrylamide (PNIPAM- COOH) (0.6 w% in water) and reacted in the oven for 2h at 120 °C. The polymer mat on the coverslip was washed twice with THF to remove any excess PNIPAM. Finally, the coverslip was dried in the oven above 40 °C to remove the THF. The contact angle was measured on this coverslip at room temperature and again at 40 °C.5 82 Scanning Electron Microscopy: Dried electrospun PGMA fibers both with and without nanoparticles were coated with 4 nm of platinum and imaged using scanning electron microscope (Zeiss 1555, VP-FESEM) at an accelerating voltage of 4-5 kV. The fiber diameters and width of the matrix were calculated using the image analysis software ImageJ (NIH). A minimum of 50 random fibers were measured. The data is reported as an average ± standard error mean. X-ray Microanalysis: Elemental analysis was carried out using the Oxford x-ray microanalysis system (Oxford instrument X-MAX (80mm2)) set up on the SEM (Zeiss 1555, VP-FESEM). Elemental data was collected at higher accelerating voltage of 15 kV at 10 mm working distance required for x-ray microanalysis. Data was analysed using the Aztec version 2.1a analysis instrument software. Gas Sensing: The hydrogen sensor setup consists of a test chamber with inlet gas and outlet gas, a potentiostat and an electronic recorder. See reference for complete detail of the experimental setup.6 Prior to entering the test chamber, hydrogen and nitrogen gas, which are tested and carrier gas respectively, were mixed in a cyclonic mixer. Two silver epoxy electrodes were painted to the Pd/ES-PGMA composite fibers mounted on an IDE and the whole integration was mounted into the gas chamber subject to current-voltage (I-V) sweeps. The procedure involved alternating nitrogen gas (20 min) and varying concentrations of hydrogen gas (4 min). The change in the current was monitored simultaneously. The total gas flow rate was 1000 mL/min. The voltage applied between 83 the electrodes was 100 mV dc. An IDE consists of 15 fingers, each 15 μm in width and 550 μm in length, with a finger gap of 10 μm. Each individual electrode is connected to a boding pad (200 μm×250 μm) to provide a sufficient area for subsequent wire bonding. Transmission Electron Microscopy (TEM): Nanoparticles were air dried onto carbon-coated copper grids and imaged using a JEOL 3000F TEM operating at 300 kV. The nanoparticle size was determined using Image J software. A minimum of 200 particles were measured and the data reported as the average ± standard error mean. Superconducting Quantum Interference Device (SQUID) Magnetometry: The magnetic properties of the dried magnetite/ES-PGMA composite (88.2 mg) were measured using a SQUID magnetometer (Quantum Design 7 Tesla MPMS) operating between 5K and 300K. Magnetite nanoparticles (22.5 mg) were lyophilised prior to their measurement and their saturation magnetisation was recorded at 70 kOe at 5K. The zero field cooled and field cooled measurements were conducted in a field of 0.1 kOe. Contact Angle: Static contact angles of MilliQ water on the surface of the electrospun polymer matrix were measured using a home-built goniometer with Rame- Hart scope attachment.7 The polymer was electrospun on the 12 mm glass coverslips (Cat. # G401-12, ProSciTech) both with and without nanoparticles (refer PNIPAM attachment on ES-PGMA). 5 µL of water was pipetted onto the membrane and images were taken after the drop edges came to rest (~2 min) using a Canon EOS450D and the angle was measured using the calibrated 84 Rame-Hart scope. Measurements were carried out at room temperature and again at 40 °C. Images were processed using ImageJ software. Measurements were done in duplicates and reported as average ± standard deviation. NIR Room Temperature Emission Spectroscopy (λ exc = 974 nm): Upconversion spectra measurements were carried out as previously described.8 Briefly, upconverting particles were suspended in chloroform and were air dried on glass slides. The upconversion spectra were obtained with an optical set up incorporating a 980 nm laser diode. The peak wavelength of the laser diode is 974.5 nm. The optical excitation intensity for obtaining the spectrum shown in Fig. 3A was 7665 W/mm2 References 1. V. Tsyalkovsky, R. Burtovyy, V. Klep, R. Lupitskyy, M. Motornov, S. Minko and I. Luzinov, Langmuir, 2010, 26, 10684-10692. 2. S. Sun, H. Zeng, D. B. Robinson, S. Raoux, P. M. Rice, S. X. Wang and G. Li, J. Am. Chem. Soc., 2003, 126, 273-279. 3. Z. Niu, Q. Peng, M. Gong, H. Rong and Y. Li, Angew. Chem. Int. Ed., 2011, 50, 6315-6319. 4. C. Liu, Z. Gao, J. Zeng, Y. Hou, F. Fang, Y. Li, R. Qiao, L. Shen, H. Lei, W. Yang and M. Gao, ACS Nano, 2013, 7, 7227-7240. 5. G. D. Fu, L. Q. Xu, F. Yao, K. Zhang, X. F. Wang, M. F. Zhu and S. Z. Nie, ACS Appl. Mater. Inter., 2009, 1, 239-243. 85 6. J. Zou, L. J. Hubble, K. S. Iyer and C. L. Raston, Sensor Actuat.B-Chem., 2010, 150, 291-295. 7. M. V. Baker and J. D. Watling, Langmuir, 1997, 13, 2027-2032. 8. M. Challenor, P. Gong, D. Lorenser, M. Fitzgerald, S. Dunlop, D. D. Sampson and K. Swaminathan Iyer, ACS Appl. Mater. Inter., 2013, 5, 7875-7880. 86 Enhancing the Efficacy of Cation-Independent Mannose 6-Phosphate Receptor Inhibitors by Intracellular Delivery** Vipul Agarwal, Priyanka Toshniwal,# Natalie E Smith,# Nicole M Smith, Binbin Li, Tristan D Clemons, Lindsay T Byrne, Foteini Hassiotou, Fiona M Wood, Mark Fear, Ben Corry* and K Swaminathan Iyer* Supporting Information S1.1 Cell culture Human primary dermal fibroblast cell cultures from normal skin were cultured in Dulbecco’s Modified Eagle’s Medium (DMEM/F12 - GlutaMAX; Invitrogen Gibco) supplemented with 10 % fetal bovine serum (FBS; Invitrogen Gibco) and 1% penicillin/streptomycin (Invitrogen Gibco). Analogues 1 and 2 were provided by Pharmaxis Ltd after stringent QC analysis. M6P and analogue 1 were dissolved in PBS filter sterilized, aliquoted and stored at -20 °C. Each aliquot was used for maximum 2 freeze thaw cycles. Analogue 2 was dissolved in DMSO and treated the similar way as other analogues. The cells were incubated at 37 °C in a humidified atmosphere of 5 % CO2. All experiments were carried out with cells between passages 3-6. S2.1 Cell Viability Cell viability was determined using a LIVE/DEAD viability/cytotoxicity Kit (Invitrogen, UK) which measures the membrane integrity of cells, as per manufacturer’s protocol. In brief, 20000 cells were seeded in each well in a 24 well plate and incubated with analogues at 10 µM concentration in cell culture media (DMEM F-12 containing 10% FBS and 1% Penicillin/ Streptomycin) for 30 min (media was added in case of controls). Following which human recombinant TGFβ1 87 (2 ng/mL) (Cat.# 14-8348-62, eBioscience) was added to the wells (fresh media was added in controls) and plates were incubated for 24 h and 72 h in the humidified incubator at 37 °C with 5 % CO2. At the stipulated time (24 h and 72 h), cells were washed with PBS (3 times) and then stained with calcein (100 µL, 1 µM)/ ethidium bromide (100 µL, 2 µM) in PBS and incubated in the humidified incubator at 37 °C and 5 % CO2 for 30 min. Images were captured using an Olympus IX71 inverted microscope with a 20 x objective with fixed exposure time. Both live and dead cells were counted using Image J software (NIH) with cell counter plug in. Data presented as mean ± standard error mean (n = 4). S3.1 Cell Body Area Cell size was measured using Image J software (NIH).[1] Minimum of 40 cells were randomly selected from the fluorescence images and their area was measured. Values reported as mean ± standard error mean. S4.1 Cell Proliferation Cell proliferation was measured using the MTS assay (Cell Titer 96 ®Aqueous, Promega, Madison, USA) as per the manufacturer’s protocol. Briefly, 1500 cells were seeded in each well of a 96 well plate and treated with analogues at 10 µM concentration in cell culture media (DMEM F-12 containing 10 % FBS and 1 % Penicillin/ Streptomycin) for 30 min (media was added in case of controls) prior to the addition of human recombinant TGFβ1 (2 ng/mL) to each well (fresh media was added in controls) and the plates (individual for each time point) were incubated for 72 h in the humidified incubator at 37 °C with 5 % CO2. MTS solution (40 µL) was added in each well the next day, and was considered as 0 h time point, and incubated for 3 h in the humidified incubator at 37 °C with 5 % CO2. Following which 80 µL 88 from each well was transferred into a new 96 well plates and read under a plate reader at 490 nm excitation wavelength. Same protocol was followed at every time point for next 72 h. Data presented as mean ± standard error mean (n = 5). S5.1 Gene Expression Gene expression was measured using real time quantitative polymerase chain reaction. Cells (50000) were seeded in 24 well plates and incubated for 24 h in the humidified incubator at 37 °C and 5% CO2. Next day culture media (10 % FBS) was replaced with starve media (0.1 % FBS) and incubated for further 24 h in the incubator to bring all the cells under same physiological cycle. Next day, cells were treated with required concentration of M6P analogue 30 min prior to TGFβ1 (2 ng/mL in starve media) stimulation and further incubated for 48 h in the humidified incubator at 37 °C and 5% CO2. mRNA was extracted using RNeasy mini kit® according to manufacturers’ protocol (Qiagen GmbH). For reverse transcription 1.5 µg of total mRNA was converted to cDNA using Superscript VILO (Cat.# 11754, Applied Biosystems) according to manufacturers’ protocol. 150 ng of cDNA was analysed by ABI 7500 fast analysis real-time PCR system using TaqMan® master mix and col1a1 probes (Hs01076777_m1, Life Technologies). GAPDH was used as a reference gene (Cat.# 4326317E, Life Technologies). Analysis was carried out using the instrument software. Data presented as mean ± standard error mean (n = 3). S6.1 Protein Expression 89 Protein expression was measured using western blotting. Cells (1 x 105) were seeded in each well of a 6 well plate and incubated for 24 h in the humidified incubator at 37 °C and 5% CO2. Next day culture media (10 % FBS) was replaced with starve media (0.1 % FBS) and incubated for further 24 h in the incubator to bring all the cells under same physiological cycle. Next day, cells were treated with required concentration of M6P analogue 30 min prior to TGFβ1 (2 ng/mL in starve media) stimulation and further incubated for 72 h in the humidified incubator at 37 °C and 5% CO2. Whole cell lysates were prepared from treated cells. 35 µg of protein was denatured and subjected to SDS-PAGE, and transferred to nitrocellulose membrane (Cat.# 10600007, Amersham, General Healthcare Lifesciences) by standard transfer. Post transfer, membrane was blocked with 5% skim milk/0.1% TBST for 30 min at room temperature, then incubated overnight with rabbit anti-human collagen I antibody (1:2000 in 5% skim milk/0.1% TBST, Cat. # NB600-408, Novus) at 4 °C. Membranes were washed with 0.1% TBST and incubated with peroxide conjugated mouse anti-rabbit (1:5000 in 5% skim milk/0.1% TBST, Cat.# NA934VS, GE Healthcare Lifesciences) for 1 h in 5% skim milk/0.1% TBST at room temperature. Immunoreactivity was detected using the chemiluminescent HRP substrate (Cat.# WBKLS0100, Millipore IMMobilon) and the signal was captured with the Chemidoc (BioRad, Model #731BR02144) and analysed using ImageJ software (NIH). To confirm equality of protein loading, all membranes were stripped and reanalysed for β-actin expression using 1° antibody (1:50000 in 5% skim milk/0.1% TBST, Cat.# A1978, Sigma) and 2° (1:5000 in 5% skim milk/0.1% TBST, Cat.# NA9310V, GE Healthcare Lifesciences). Data presented as mean ± standard error mean (n = 3). 90 S7.1 Statistics The results for cell viability, cell body area, cell proliferation, gene and protein expression are expressed as mean ± standard error mean (SEM) and analysed for analysis of variance (ANOVA). Significance was evaluated using Bonferroni and Turkey’s post-hoc analysis and set at 95% confidence (p < 0.05). S8.1 Simulation Systems: Experimental Methods: 6 simulation systems of the M6P/IGFII receptor were investigated; these included the domain 3 and domain 5 dimers with a ligand in each binding site (2 binding sites per dimer). These ligands were M6P, analogues 1 and 2. The coordinates for the domain 3 dimer with M6P bound (pdb accession code 1SYO)[2] and the domain 5 monomer (pdb accession code 2KVB)[3] were obtained from the protein database. In order to obtain the dimer for domain 5 and position M6P in the binding pocket, the domain 5 beta sheet regions were aligned with the corresponding beta sheet regions of domain 3. Analogue 1 and 2 were positioned in each binding pocket by aligning the mannose ring and phosphate group to the M6P coordinates obtained for domain 3.[2] Each system was then solvated in a TIP3P water box of dimensions 65 x 60 x 114 Å and ionised with 150 mM KCl. All simulations were run with periodic boundary conditions, constant temperature (310 K) maintained using Langevin dynamics and constant pressure (1 atm) maintained with a Langevin piston, and the particle mesh Ewald method was used to compute full system electrostatics.[4] The CHARMM 36 force field was used for protein, water and M6P parameters.[5] The ion parameters were obtained from Joung and Cheatham.[6] Missing parameters for analogues 1 and 2 were obtained using ab-initio techniques 91 [7] with the program Gaussian 09.[8] All molecular dynamics simulations were run with the program NAMD[9] using rigid bonds to hydrogen and 2 fs time steps. Molecular graphics were generated using VMD.[10] Water and ions were energy minimized for 5000 steps and equilibrated for 20 ps with the protein and substrate held fixed. A harmonic restraint with a force constant of 20 kcal/mol was then applied to the backbone atoms of the protein and on each ligand and the system was minimized for a further 10,000 steps prior to 500 ps of equilibration. This step was repeated with gradual reductions in the force constant with values of 10, 5, 2.5 and 1 kcal/mol. Finally, to replicate the influence of the surrounding protein domains on the individual domain being simulated, a harmonic restraint with a force constant of 0.1 kcal/mol was placed on all of the protein Cα atoms which were located more than 10 Å from the binding pocket in order to ensure no loss of secondary structure throughout the simulation. The system was then minimized for a further 10,000 steps prior to 10 ns of equilibration. Subsequently 100 ns of equilibrium simulation were obtained for each of the 6 systems. Data Analysis: Cluster analysis was performed on the final 100 ns of equilibrium simulation for each ligand in order to determine the most occupied binding positions. Each ligand was clustered according to the RMSD of its coordinates with a cut-off of 3 Å. Subsequently the NAMDEnergy plugin of VMD was utilized to determine the interaction energy of each ligand with the protein for the entire time and for the most populated clusters. 92 Figure S1: Most populated binding positions (clusters) of each drug in the domain 3 and domain 5 binding domains as determined from 100 ns of molecular dynamics simulation. In each case the protein is shown in grey, and the most to the least populated clusters are for ligand 1 (initially bound to dimer 1): cyan, pink, mauve, white and green and for ligand 2 (initially bound to dimer 2): dark blue, red, orange, yellow and green. a) M6P in domain 3: Ligands 1 and 2 both remain in the vicinity of the binding site determined in the crystal structure. However as there are multiple hydrogen bond acceptors and donors, both ligands sample multiple positions as the mannose ring hydroxyl groups form hydrogen bonds with various residues and these change over the 100 ns simulation. b) Analogue 1 in domain 3: Ligand 1 has only one binding position (cyan) and ligand 2 has only two clusters indicating that in both cases it binds stably to the protein and remains in the binding pocket throughout the simulation. This binding appears to be stabilized by the interaction between the m-xylene rings of the 2 ligands. c) Analogue 2 Domain 3: Both ligand 1 and 2 are oriented such that the benzyl groups associate in the center of the dimerization domain, hence while the ligands do have more than one binding position they remain in the vicinity of the predicted binding site. d) Domain 5 M6P: In this case the ligands occupy positions which are not asymmetrical, while ligand 1 favors a horizontal orientation with one major binding position (cyan) ligand 2 favors a vertical orientation where it has more flexibility to sample new positions. e) Analogue 1 in domain 5: Ligand 1 moves away from its initial position in the binding site and occupies a position where it is in close contact with ligand 2. f) Analogue 2 in domain 5: In this case both ligands remain in the vicinity of the original binding position. However, while ligand 1 has one major cluster implying its motion is restricted, ligand 2 has more flexibility to move occupying multiple clusters. 93 Figure S2: Snapshots from 100 ns molecular dynamics simulations representative of the most heavily occupied binding positions for each ligand in the domain 3 and domain 5 dimers. The two protein subunits are shown in pink and grey respectively and residues which form common hydrogen bonds are labelled. a) M6P and b) Analogue 1 in domain 3: Ligand 1 and 2 form the strongest energetic interactions with Lys350, Lys358, Ser386 and Arg391. c) Analogue 2 in domain 3: Ligand 1 and 2 form the strongest interactions with Gln348, Arg391 and Glu416. d) M6P in domain 5: Ligand 1 and 2 are oriented differently to each other in the binding pocket. M6P forms its strongest interactions with Arg687. e) Analogue 1 in domain 5: interacts most favorably with Asn680 andArg687. f) Analogue 2 in domain 5: Ligand 1 and 2 form the strongest interactions with Gln644, Trp653, Arg687 and Tyr714. Figure S3: Average interaction energies obtained for the most occupied positions of each ligand in domains 3 and 5. R391 in domain 3 is equivalent to R687 in domain 5. M6P (a) and analogue 1 (b) in domain 3 both show similar interactions for each ligand and the protein subunit it is most closely associated with (for example Protein 1 with Ligand 1 as compared to Ligand 2 with Protein 2). This is because the binding mode is similar for each ligand in its respective binding site (Figure S2). Similarly, in each case the ligands also interact closely with Lys350 which is actually located on the alternate protein subunit (for example Ligand 1 to Protein 2). When compared to (a) and (b) analogue 2 (c) in domain 3 shows significant decreases in the interaction energies which is reflected by its lower overall interaction energy. M6P (d) and analogue 1 (e) in domain 5 have two major interactions the first one to Arg687 and the second to Glu709. Overall the interaction energies for domain 5 are much lower than those observed for domain 3. Analogue 2 (f) in domain 5 shows a similar pattern with a further reduction in the observed interaction energies. 94 Figure S4: Cell Proliferation assay showing cell growth over the period of 72 h post incubation with analogues both in the presence and absence of TGFβ1. First and second column in each condition is representing 24 h and 72 h respectively. Significant proliferation was observed in all groups despite the reduction in proliferation in TGFβ1 treated groups. Data presented as Mean ± SEM. Significance was set at * p < 0.05 using one way ANOVA with Bonferroni post-hoc analysis. Figure S5: HDF cell morphology post calcein AM staining imaged using fluorescent microscopy. Cells were treated for 72h, stained and imaged: a) untreated (control), b) TGFβ1 (2 ng/mL) treatment, c) M6P (10 µM) + TGFβ1 (2 ng/mL), d) Analogue 1 (10 µM) + TGFβ1 (2 ng/mL) and e) Analogue 2 (10 µM) + TGFβ1 (2 ng/mL). Scale bar 2 µm. Figure S6: Collagen I gene time response curve post TGFβ1 (2 ng/mL) stimulation. Collagen I gene expression was significantly upregulated at 24 and 48 h post TGFβ1 stimulation compared to non-treated control. The 48 h stimulation was selected for all further experiments as it yielded significantly higher response. Data presented as average ± SEM (n = 3). Significance was set at * p < 0.05 using bonferroni post-hoc test in one way ANOVA analysis. 95 References: [1] T. J. Collins, BioTechniques 2007, 43, S25-S30. [2] L. J. Olson, N. M. Dahms, J.-J. P. Kim, J. Biol. Chem. 2004, 279, 34000- 34009. [3] L. J. Olson, F. C. Peterson, A. Castonguay, R. N. Bohnsack, M. Kudo, R. R. Gotschall, W. M. Canfield, B. F. Volkman, N. M. Dahms, Proc. Natl. Acad. Sci. 2010, 107, 12493-12498. [4] U. Essmann, L. Perera, M. L. Berkowitz, T. Darden, H. Lee, L. G. Pedersen, J. Chem. Phys. 1995, 103, 8577-8593. [5] A. D. MacKerell, D. Bashford, M. Bellott, R. L. Dunbrack, J. D. Evanseck, M. J. Field, S. Fischer, J. Gao, H. Guo, S. Ha, D. Joseph-McCarthy, L. Kuchnir, K. Kuczera, F. T. K. Lau, C. Mattos, S. Michnick, T. Ngo, D. T. Nguyen, B. Prodhom, W. E. Reiher, B. Roux, M. Schlenkrich, J. C. Smith, R. Stote, J. Straub, M. Watanabe, J. Wiorkiewicz-Kuczera, D. Yin, M. Karplus, J Phys Chem B 1998, 102, 3586-3616. [6] I. S. Joung, T. E. Cheatham, 3rd, J. Phys. Chem. B 2008, 112, 9020-9041. [7] N. Foloppe, A. D. MacKerell, J Comput Chem 2000, 21, 86-104. [8] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. M. Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. 96 Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, D. J. Fox, 2009, Revision D.01. [9] J. C. Phillips, R. Braun, W. Wang, J. Gumbart, E. Tajkhorshid, E. Villa, C. Chipot, R. D. Skeel, L. Kale, K. Schulten, J Comput Chem 2005, 26, 1781-1802. [10] W. Humphrey, A. Dalke, K. Schulten, J Mol Graph Model 1996, 14, 33-38. 97 Inhibiting the activation of transforming growth factorβ using a polymeric nanofiber scaffold Vipul Agarwal,a Fiona M. Wood,b,c Mark Fearb and K. Swaminathan Iyer*a Supporting Information S1. Materials Polyglycidyl methacrylate (PGMA) with Mn = 220515 and Mw = 433730 was synthesized by radical polymerisation as reported previously.1 MEK was purchased from Merck. PXS64 was kindly provided by Pharmaxis Ltd after intensive QC testing. S2. Electrospinning Procedure: PGMA (15 wt%) was dissolved in MEK overnight at room temperature with constant stirring. In case of PGMA + PXS64, PXS64 (2.95 mg) was mixed with PGMA and dissolved in MEK overnight at room temperature. PGMA polymer and PGMA + PXS64 fibrous scaffold were obtained via electrospinning (Nanofiber Electrospinning Unit, Cat. # NEU-010, Kato Tech, Japan) onto 12 mm and 6 mm glass coverslips (Cat. # G40112, and G401-06 respectively, ProSciTech). The electrospinning parameters after optimisation were a voltage of 9.1 kV, working distance of 9 cm, syringe pump speed of 0.04 mm/min (1 mL/h), and the traverse and collection drum speeds were set at 0 cm/min and 0 m/min respectively. Fibers were dried and crosslinked by annealing at 80 °C for 5 h and stored at room temperature for maximum of a month. Scaffolds were UV sterilised in the tissue culture hood for 15 min and further washed with 2 x PBS prior to cell experiments. S3. Scanning Electron Microscopy: Dried electrospun PGMA fibres both with and without PXS64 were coated with 4 nm of platinum and imaged using scanning electron microscope (Zeiss 1555, VP-FESEM) at an accelerating voltage of 4-5 kV. The fiber diameter was calculated using the image analysis software ImageJ (NIH). A minimum of 50 random fibers were measured. The data is reported as an average ± standard deviation. In the cell experiments, cells were incubated as per cell viability protocol. At the stipulated time point, culture media was removed and coverslips loaded with scaffold and cells were washed with 2 x PBS and fixed for 10 min using glutaraldehyde (2.5 % in PBS), 98 followed by 2 x PBS washes. Next serial dehydration steps were performed with increasing concentration of ethanol to replace all the water in the samples with dry ethanol before critical point drying step. Post critical point drying samples were coated with 4 nm of platinum and imaged using aforementioned parameters. S4. High Pressure Liquid Chromatography Drug loading analysis was carried out using HPLC (solvent A: water with 0.1 % TFA, solvent B: acetonitrile with 0.1 % TFA, Phenomenex-C18 (2) 100A column (4.6 × 150 mm, 5 microns) at room temperature, 1 mL/ min, λ = 280 nm, gradient: 100 % solvent B in 17 min) with Water 2695 pumping system and 2489 UV/Vis detector. Drug loading was calculated by dissolving 35 mg of fibers (PGMA + PXS64) in acetonitrile (0.5 mL) and calculated from the standard curve developed by dissolving free PXS64 in acetonitrile. S5. Cell culture Human primary dermal fibroblast cell cultures from normal skin were cultured in Dulbecco’s Modified Eagle’s Medium (DMEM/F12 - GlutaMAX; Invitrogen Gibco) supplemented with 10% fetal bovine serum (FBS; Invitrogen Gibco) and 1 % penicillin/streptomycin (Invitrogen Gibco). The cells were incubated at 37 °C in a humidified atmosphere with 5 % CO2. All experiments were carried out with cells between passages 3-6. S6. Cell Viability Cell viability was determined using a LIVE/DEAD Viability/Cytotoxicity Kit (Invitrogen, UK) which measures the membrane integrity of cells, as per manufacturer’s protocol. In brief, 20000 cells were seeded on the sterilised electrospun scaffold on glass coverslips in a 24 well plate and supplemented with cell culture media (DMEM F-12 containing 10 % FBS and 1 % Penicillin/ Streptomycin) for 30 min. Next, human recombinant TGFβ1 (2 ng/mL, Cat. # 14-8348-62, eBioscience) was added to the wells (fresh media was added in controls) and plates were incubated for 24 h and 72 h in the humidified incubator at 37 °C with 5 % CO2. At the stipulated time (24 h and 72 h), cells were washed with PBS (3 times) and then stained with calcein AM (100 µL, 1 µM)/ethidium bromide I (100 µL, 2 µM) in PBS and further incubated for 30 min in the incubator. Images were captured using an Olympus IX71 inverted microscope with a 20 x objective with fixed exposure time. Both live and dead cells were counted using Image J software (NIH) with cell counter plug in. Data presented as mean ± standard error mean (n = 4). 99 S7. Cell Proliferation Cell proliferation was measured using the MTS assay (Cell Titer 96 ®Aqueous, Promega, Madison, USA) as per the manufacturer’s protocol. Briefly, 1500 cells were seeded on the sterilised electrospun scaffold on 6 mm glass coverslips in each well of a 96 well plate and supplemented with cell culture media (DMEM F-12 containing 10 % FBS and 1 % Penicillin/ Streptomycin) for 30 min followed by the addition of human recombinant TGFβ1 (2 ng/mL) to the wells (fresh media was added in controls) and the plates (individual for each time point) were incubated for 72 h in the humidified incubator at 37 °C with 5 % CO2. MTS solution (40 µL) was added in each well the next day, and was considered as 0 h time point, and incubated for 3 h in the humidified incubator at 37 °C with 5 % CO2. Following which 80 µL from each well was transferred into a new 96 well plates and read under a plate reader at 490 nm excitation wavelength. Same protocol was followed at every time point for next 72 h. Data presented as mean ± standard error mean (n = 5). S8. Gene Expression Gene expression was measured using real time quantitative polymerase chain reaction. Cells (50000, in DMEM/F12 media containing 10% FBS) were seeded on sterilised scaffolds (electrospun on coverslips) in 24 well plates and incubated for 24 h in the humidified incubator at 37 °C and 5% CO2. Next day culture media (containing 10 % FBS) was replaced with starve media (containing 0.1 % FBS) and incubated for further 24 h in the incubator to bring all the cells under same physiological cycle. Next day, TGFβ1 (2 ng/mL in starve media) was added in the culture and further incubated for 48 h in the incubator. mRNA was extracted using RNeasy mini kit® according to manufacturers’ protocol (Qiagen GmbH). For reverse transcription 1.5 µg of total mRNA was converted to cDNA using Superscript VILO (Cat. # 11754, Applied Biosystems) according to manufacturers’ protocol. 150 ng of cDNA was analysed by ABI 7500 fast analysis real-time PCR system using TaqMan® master mix and col1a1 probes (Hs01076777_m1, Life Technologies). GAPDH was used as a reference gene (Cat. # 4326317E, Life Technologies). Analysis was carried out using the instrument software. Release media was prepared using DMEM/F-12 media (containing 10 % FBS and 1 % Penicillin/ Streptomycin) supplemented with 0.2 % DMSO. Data presented as mean ± standard error mean (n = 3). S9. Statistics 100 The results for cell viability, cell proliferation and gene expression are expressed as mean ± standard error mean (SEM) and analysed for analysis of variance (ANOVA). Significance was evaluated using Bonferroni and Turkey’s post-hoc analysis and set at 95% confidence (p < 0.05). Figure S1: Cell proliferation assay showing cell growth over the period of 72 h post incubation with the two scaffolds both with and without TGFβ1 (2 ng/mL). Data presented as Mean ± SEM. Significance was set at * p < 0.05 using one way ANOVA and Bonferroni post-hoc analysis. Figure S2: A representative florescent images showing cell morphology and the number of live and dead cells in culture. HDF cells were incubated on a) ES-PGMA scaffold, b) ES-PGMA + TGFβ1 (2 ng/mL) and c) ES-PGMA + PXS64 scaffold. Scale bar 2 µm. Cells were stained using Calcein AM/Ethidium bromide I staining where live cells fluoresce green while dead cells fluoresce red. 101 Figure S3: A representative image showing cell adherence on the two scaffolds. HDF cells were incubated a) ES-PGMA scaffold, b) ES-PGMA + TGFβ1 (2 ng/mL) and c) ES-PGMA + PXS64 scaffold. Scale bar: a) 2 µm, b) and c) 1 µm. Red arrows highlighting the cells. References 1. K. S. Iyer, B. Zdyrko, H. Malz, J. Pionteck and I. Luzinov, Macromolecules, 2003, 36, 6519-6526. 102 Appendix B Published work not directly included in the thesis Peer reviewed publications contributed by the candidate are listed below. However, only published articles are attached in the thesis. 1. Agarwal, V., Tjandra, E.S., Iyer, K. S., Humfrey, B., Fear, M., Wood, F. W., Dunlop, S. and Raston, C. L., Evaluating the effects of nacre on human skin and scar cells in culture, Toxicology Research, 3, 223-227 (2014) 2. Eroglu, E., Chen, X., Bradshaw, M., Agarwal, V. , Zou, J., Stewart, S.G., Duan, X., Lamb, R.N., Smith, S.M., Raston, C. and Iyer, K. S., Biogenic production of palladium nanocrystals using microalgae and their immobilization on chitosan nanofibers for catalytic applications, RSC Advances, 3, 1009-1012 (2013) 3. Eroglu, E., Agarwal, V., Bradshaw, M., Chen, X., Smith, S. M., Raston, C. and Iyer, K. S., Nitrate removal from liquid effluents using microalgae immobilized on chitosan nanofiber mats, Green chemistry, 14 (10), 2682-2685 (2012) 4. Ho, D., Zou, J., Chen, X., Munshi, A., Smith, N. M., Agarwal, V., Hodgetts, S.I., Plant, G. W., Bakker, A., Harvey, A. R., Luzinov, I., Iyer, K. S., Hierarchical patterning of multifunctional conducting polymer nanoparticles as a bionic platform for topographic contact guidance, ACS nano, 9 (2), 1767-1774 (2015) 103 Volume 3 Number 4 July 2014 Pages 217–292 Toxicology Research www.rsc.org/toxicology ISSN 2045-452X Chinese Society Of Toxicology COMMUNICATION Colin L. Raston et al. Evaluating the effects of nacre on human skin and scar cells in culture Toxicology Research Published on 02 May 2014. Downloaded by University of Western Australia on 09/10/2014 10:23:29. COMMUNICATION Cite this: Toxicol. Res., 2014, 3, 223 Received 6th January 2014, Accepted 2nd May 2014 DOI: 10.1039/c4tx00004h View Article Online View Journal | View Issue Evaluating the effects of nacre on human skin and scar cells in culture† Vipul Agarwal,a Edwin S. Tjandra,a K. Swaminathan Iyer,a Barry Humfrey,b Mark Fear,c Fiona M. Wood,c Sarah Dunlopd and Colin L. Raston*e www.rsc.org/toxicology Pearl nacre, a biomineralisation product of molluscs, has growing applications in cosmetics, as well as dental and bone restoration, yet a systematic evaluation of its biosafety is lacking. Here, we assessed the biocompatibility of nacre with two human primary dermal fibroblast cell cultures and an immortalised epidermal cell line and found no adverse effects. There are three main types of pearl oysters of the genus Pinctada: the “Akoya” pearl oyster called Pinctada fucata, the “Golden lipped” oyster Pinctada maxima and the “Black lipped” oyster named Pinctada margaritifera. Mollusc shells are mainly made up of two layers of calcium carbonate, comprising an outer layer of calcite and an inner layer of aragonite. Nacre (mother of pearl) in all oyster shells is a calcified structure that forms the lustrous inner layer. It is mainly composed of aragonite (∼95–97%) tablets oriented in multiple layers, each surrounded by organic matrix.1,2 This organic matrix makes up ∼5% of the nacre composition and is mainly comprised of polysaccharides and proteins.3 According to a European Commission report published in 2007 the cosmetic and toiletries industry in the EU, Japan, China and the US had a total market value of €136.2 billion.4 The cosmetics industry maintains its edge by constantly developing novel topical skin treatments. A popular example is the use of all-natural or organic ingredients, such as fruit and plant extracts to offer wrinkle relief that mimics the painful and potentially dangerous side effects associated with invasive chemical remedies.5 Clinically, topical treatments containing, for example, aloe vera, vitamin C, corticosteroids and tacrolimus are used with a School of Chemistry and Biochemistry, The University of Western Australia, Crawley, Australia b Pearl Technology Pty Ltd, Geraldton, Australia c Burn Injury Research Unit, School of Surgery, The University of Western Australia, Crawley, Australia d School of Animal Biology, The University of Western Australia, Crawley, Australia e School of Chemical and Physical Sciences, Flinders University, Bedford Park, Australia. E-mail: colin.raston@flinders.edu.au; Tel: +61 82017958 † Electronic supplementary information (ESI) available. See DOI: 10.1039/c4tx00004h This journal is © The Royal Society of Chemistry 2014 the aim of minimizing scarring.6 Recently, there has been interest in the cosmetics industry in the use of nacre as a key ingredient.7 Most of the formulations are reported to either use powdered pearl shell or powdered nacreous layer shell. Powdered shell and powdered nacre comprises of both organic and inorganic components. It is reported that nacre stores in its mineral-based organic structure a variety of bioactive molecules. Efficacy of this water soluble matrix (WSM) has been tested in a porcine burn injury model.8 WSM was obtained by suspending powdered nacre in ultra-pure water and collecting the supernatant via precipitation of insoluble components by centrifugation. It was concluded that the active mineral based organic component has beneficial effects on the skin with enhanced wound healing.8,9 Nacre has also attracted attention for its potential in supporting bone grafting and bone regeneration. In culture under physiological conditions, nacre can transform to hydroxyapatite, the phosphorous rich main constituent of the mammalian bone framework.10,11 Nacre and its WSM can also aid in osteogenic regeneration.9,12–17 High phosphorous rich domains have been described at the interface between bone and implants made from Margaritifera shells which are biocompatible, biodegradable and osteoconductive and thus are thought to promote bone formation.18 Furthermore, nacre powder has been used as an implantable material for reconstruction and regeneration of maxillary alveolar ridge bone in humans.19 In this example, the implanted nacre dissolves gradually and is eventually replaced by the mature lamellar bone suggesting that the nacre acts as a biocompatible substrate for bone replacement.19 The water soluble components of the crushed nacre have also been investigated for their potential in bone regeneration in a similar vein.20,21 Lee et al., demonstrated the wound healing potential of WSM component in a deep burn porcine skin model and showed enhanced collagen secretion and deposition at the injury site resulting in enhanced healing.8 In another in vivo study using a rat skin incisional injury model, powdered nacre was implanted between the epidermis and dermis at the incisional site, with an aim of studying the effect of nacre on the synthesis of certain constituents of the dermal Toxicol. Res., 2014, 3, 223–227 | 223 View Article Online Published on 02 May 2014. Downloaded by University of Western Australia on 09/10/2014 10:23:29. Communication extracellular matrix. It was concluded that implanted nacre increased collagen synthesis by dermal fibroblasts.22 While extensive investigations have been carried out in bone, the evaluation of the biocompatibility of nacre with human skin cells is lacking. Thus the growing number of cosmetic formulations in the market with nacre as a key ingredient7,23,24 clearly warrants a thorough assessment with human skin cells. Since scars are also common, and contain cells with a phenotype distinct from normal skin,25 it is also important to test potential cosmetic ingredients with both cell types. In the present study, we use nacre from the inner calcified layer of the shell of Pinctada margaritifera and report the in vitro toxicity assessment of the material on three cell types representing both epidermal and dermal layers of human skin. These were HaCaT cells, a human derived immortalised keratinocytes cell line, primary human dermal skin fibroblasts (HDF) and primary human scar fibroblasts (HSF). Nacre used in the study was gently scraped26,27 from the inner layer of the shell to avoid the post processing required in the case of powdered shell. SEM images (Fig. 1) confirmed that the nacre was composed of pseudo-hexagonal shaped aragonite tablets which have basal plane dimensions of 2–6 µm, and a thickness of 300–400 nm.28 This structure is characteristic of previously reported nacre, which is a composite material consisting of alternating layers of mineral tablets separated by thin layers of biomacromolecular “glue”.29,30 Fig. 1 Top view of the scrapped nacre, imaged using scanning electron microscopy (SEM). Scale bar (a) 1 µM and (b) 2 µM respectively. Sample was coated with 4 nm platinum prior to imaging. 224 | Toxicol. Res., 2014, 3, 223–227 Toxicology Research To test the cytotoxicity of nacre, a live/dead assay was carried out (see ESI S1.4†). Cells were incubated with nacre in culture media at physiological conditions for 24 h to 72 h and were then stained for viability using calcein AM/ethidium bromide I solutions. Viable cells fluoresce green through the reaction of calcein AM with intracellular esterase, whereas non-viable cells fluoresce red due to the diffusion of ethidium homodimer across damaged cell membranes and binding with nucleic acids. Fig. 2 shows live cells as the percentage of the total cells in human primary dermal skin fibroblast (HDF), human primary scar fibroblast (HSF) and human derived immortalised HaCaT cell cultures when exposed to various concentrations of nacre for 24 h and 72 h. Cytotoxicity of nacre was not observed for any of the concentrations examined in HDF cells (Fig. 2a). However, interestingly at a concentration of 2.5 mg ml−1 of nacre (highest concentration tested) there was a significant reduction in viability at both 24 and 72 hours in the HSF cells (Fig. 2b). This underlines the importance of testing both scar and normal skin cell types for cosmetic application. Toxicity was also observed at a concentration of 0.5 mg ml−1 in HaCaT cells (Fig. 2c), although this was only observed at 24 hours and Fig. 2 Cell viability assays showing percentage of live cells in the culture post incubation with nacre. (a) Human dermal skin fibroblasts cells, (b) human scar fibroblasts cells and (c) human derived immortalised HaCaT cells were incubated with various concentrations of scrapped nacre and treated with calcein AM/ethidium bromide I to stain for live and dead cells. Both live and dead cells were counted using fluorescence microscopy. ‘None’ is the untreated control. Data presented as average ± SEM (n = 4). Significance was set at *p < 0.05 using bonferroni test in one way ANNOVA. This journal is © The Royal Society of Chemistry 2014 View Article Online Published on 02 May 2014. Downloaded by University of Western Australia on 09/10/2014 10:23:29. Toxicology Research was no longer present at the 72 hour time point. These results are in line with the results obtained previously where constituents of nacre were shown to promote wound healing in a rat model22 and deep burn porcine skin.8 In both these studies, nacre has been shown to promote the recruitment of fibroblasts for restoration and coverage of the injury site while showing no apparent signs of cytotoxicity. It has also been shown to promote bone formation when implanted in the femur of sheep with midshaft hemidiaphysis resection of their femur in vivo reiterating the non-cytotoxic advantage of nacre.31 Fibroblasts have been reported to undergo morphological changes from dendritic to stellate shapes upon exposure to external cues caused by changes in actin polarisation and adhesion.32,33 Cell morphology in fibroblasts is known to be influenced by cytokines such as transforming growth factor β which can potentially induce polymerisation of globular to filamentous actin.34 Fibroblast morphology can also be modulated by extracellular matrix architecture during wound healing via cell–matrix interaction.32 Such morphological change has been observed in cells undergoing oxidative stress.35,36 In our study, we found similar changes in fibroblast morphology for both HDF and HSF cells at the highest concentration of nacre of 2.5 mg mL−1 (Fig. 3). Similar altered mor- Fig. 3 Cell morphology post calcein AM staining and imaged using fluorescent microscopy. Cells were treated with various concentrations of nacre for 24 h, stained and imaged. (i) untreated (control) primary human dermal skin fibroblasts (HDF), (ii) HDF treated with 0.05 mg mL−1 nacre, (iii) HDF treated with 2.5 mg mL−1 nacre, and (iv), untreated (control) primary human dermal scar fibroblasts (HSF), (v) HSF treated with 0.05 mg mL−1 nacre and (vi) HSF treated with 2.5 mg mL−1 nacre. Scale bar 1 µm. This journal is © The Royal Society of Chemistry 2014 Communication phology was also observed for HaCaT cells (see ESI Fig. S1†). It could be postulated that the high concentration of nacre induces cellular stress, resulting in changes in the actin cytoskeleton and a more stellate morphology (Fig. 3i, iii, ii and iv). Cell area was calculated from the fluorescent images shown in Fig. 3 using Image J software.37 It was found that both HDF and HSF had significantly larger cell areas ( p < 0.05) when treated with 2.5 mg mL−1 of nacre (HDF: 2.79 ± 0.13 µm and HSF: 3.0 ± 0.19 µm respectively) as compared to the nontreated controls (HDF: 1.56 ± 0.08 µm and HSF: 1.54 ± 0.10 µm respectively) (see ESI Fig. S2†). Altered fibroblast morphology has been thought to occur in response to various factors including aging,38 strength of the extracellular matrix39 or other etiologies that induce mechanical stress on the cell. Changes in morphology also commonly indicate oxidative as well as mechanical stress.39 Therefore, we explored whether the morphological changes and increase in Fig. 4 Reactive oxygen species (ROS) assay showing ROS levels in cells stressed with various concentrations of nacre for 24 h. No significant stress was observed as a result of calcium (from nacre) induced oxidative stress at the concentrations studied. (a) Human dermal skin fibroblasts cells, (b) human scar fibroblasts cells and (c) human derived immortalised HaCaT cells were incubated with various concentrations of scrapped nacre for the specified period of time. Cells were then incubated with 2’,7’-dichlorodihydrofluorescein diacetate (DCFH-DA) solution which fluoresce in the presence of reactive oxygen species. ‘None’ is the untreated control. Data presented as average ± SEM (n = 3). Significance was set at *p < 0.05 using bonferroni test in one way ANNOVA. Toxicol. Res., 2014, 3, 223–227 | 225 View Article Online Published on 02 May 2014. Downloaded by University of Western Australia on 09/10/2014 10:23:29. Communication cell area at the highest concentration of nacre in culture was a result of, or induced oxidative stress in, the cells. Oxidative stress was tested using the cell permeable fluorogenic probe 2′,7′-dichlorodihydrofluorescein diacetate (DCFH-DA). DCFH-DA is taken up by cells and is deacetylated by cellular esterases to non-fluorescent 2′,7′-dichlorodihydrofluorescein (DCFH) which is rapidly oxidised by reactive oxygen species (ROS) to highly fluorescent 2′,7′-dichlorodihydrofluorescein (DCF). The fluorescent intensity is proportional to the ROS levels within the cytosol (see ESI S1.5†). Cell responsiveness to the assay was carried out by stressing the cells with the H2O2 solution provided in the kit which was also used to generate the calibration curve (see ESI Fig. S3†). No changes in levels of reactive oxygen species were observed in any cell type at any concentration of nacre (Fig. 4). This is important as oxidative stress is known to be a significant contributor to skin damage and excessive scarring in previous studies.40 It has been known that cells alter their morphology depending on their environment.41,42 It is therefore hypothesized that the altered fibroblast morphology in the present case is mainly due to the regulation of cell motility through geometrical constraint in the presence of nacre. Indeed, it has been previously reported that when cells probe their physical surroundings, they acquire mechanical information or signals that help to determine the direction of migration, with a consequential change in cell morphology.43 Conclusions We have established the biocompatibility of nacre using three human cell types representing the two primary layers of human skin, using immortalised keratinocytes from the epidermal layer and two primary human dermal cell cultures. The nacre used in the present study showed limited cytotoxicity at high concentrations in scar derived cells, with the morphology of the cells significantly changed by exposure at such concentrations of nacre. No apparent oxidative stress was evident in any of the cell types. Overall, the data support the use of low concentrations of nacre in aesthetic formulations, with the potential for high concentrations to cause changes in skin and/or scar cells which may have impact on efficacy. The authors would like to thank the Australian Research Council and Pearl Technologies for funding the work under the grant number LP100100812. The authors would also like to acknowledge the Australian Microscopy & Microanalysis Research Facility at the Centre for Microscopy, Characterization & Analysis, The University of Western Australia, funded by the University, State and Commonwealth Governments. Notes and references 1 J. H. E. Cartwright and A. G. Checa, J. R. Soc. Interface, 2007, 4, 491–504. 226 | Toxicol. Res., 2014, 3, 223–227 Toxicology Research 2 F. Nudelman, E. Shimoni, E. Klein, M. Rousseau, X. Bourrat, E. Lopez, L. Addadi and S. Weiner, J. Struct. Biol., 2008, 162, 290–300. 3 Y. Oaki and H. Imai, Angew. Chem., Int. Ed., 2005, 44, 6571–6575. 4 D. G. f. E. a. I. European Commission, A Study of the European Cosmetics Industry, Report, European Commission, 2007. 5 P. M. Reddy, M. Gobinath, K. M. Rao, P. Venugopalaiah and N. Reena, Int. J. Adv. Pharm. Nanotechnol., 2011, 1, 121–139. 6 J. M. Zurada, D. Kriegel and I. C. Davis, J. Am. Acad. Dermatol., 2006, 55, 1024–1031. 7 Pearlcium, Genesis Framework – WordPress, vol. 2013, p. US. 8 K. Lee, H. Kim, J. Kim, Y. Chung, T. Lee, H.-S. Lim, J.-H. Lim, T. Kim, J. Bae, C.-H. Woo, K.-J. Kim and D. Jeong, Mol. Biol. Rep., 2012, 39, 3211–3218. 9 M. J. Almeida, C. Milet, J. Peduzzi, L. Pereira, J. Haigle, M. Barthélemy and E. Lopez, J. Exp. Zool., 2000, 288, 327–334. 10 Y. Guo and Y. Zhou, J. Biomed. Mater. Res., Part A, 2008, 86, 510–521. 11 P. Westbroek and F. Marin, Nature, 1998, 392, 861–862. 12 M. Lamghari, M. J. Almeida, S. Berland, H. Huet, A. Laurent, C. Milet and E. Lopez, Bone, 1999, 25, 91S–94S. 13 M. Lamghari, H. Huet, A. Laurent, S. Berland and E. Lopez, Biomaterials, 1999, 20, 2107–2114. 14 M. Lamghari, S. Berland, A. Laurent, H. Huet and E. Lopez, Biomaterials, 2001, 22, 555–562. 15 F. Moutahir-belqasmi, N. Balmain, M. Lieberrher, S. Borzeix, S. Berland, M. Barthelemy, J. Peduzzi, C. Milet and E. Lopez, J. Mater. Sci.: Mater. Med., 2001, 12, 1–6. 16 D. Duplat, A. Chabadel, M. Gallet, S. Berland, L. Bédouet, M. Rousseau, S. Kamel, C. Milet, P. Jurdic, M. Brazier and E. Lopez, Biomaterials, 2007, 28, 2155–2162. 17 C. Silve, E. Lopez, B. Vidal, D. Smith, S. Camprasse, G. Camprasse and G. Couly, Calcif. Tissue Int., 1992, 51, 363–369. 18 H. Liao, H. Mutvei, M. Sjöström, L. Hammarström and J. Li, Biomaterials, 2000, 21, 457–468. 19 G. Atlan, N. Balmain, S. Berland, B. Vidal and É. Lopez, C. R. l’Academie. Sci., Ser. III, 1997, 320, 253–258. 20 L. Pereira Mouriès, M.-J. Almeida, C. Milet, S. Berland and E. Lopez, Comp. Biochem. Physiol., Part B: Biochem. Mol. Biol., 2002, 132, 217–229. 21 M. Rousseau, H. Boulzaguet, J. Biagianti, D. Duplat, C. Milet, E. Lopez and L. Bédouet, J. Biomed. Mater. Res., Part A, 2008, 85, 487–497. 22 E. Lopez, A. L. Faou, S. Borzeix and S. Berland, Tissue Cell, 2000, 32, 95–101. 23 E. Lopez and A. E. Chemouni, US Pat, US10/089,982, 2005. 24 D. M. DeLaRosa, US Pat, US2008/0199533 A1, 2008. 25 C. Chipev and M. Simon, BMC Dermatol., 2002, 2, 13. 26 M. Norizuki and T. Samata, Mar. Biotechnol., 2008, 10, 234–241. 27 C. Ma, C. Zhang, Y. Nie, L. Xie and R. Zhang, Tsinghua Sci. Technol., 2005, 10, 499–503. This journal is © The Royal Society of Chemistry 2014 View Article Online Published on 02 May 2014. Downloaded by University of Western Australia on 09/10/2014 10:23:29. Toxicology Research 28 K. S. Katti, B. Mohanty and D. R. Katti, J. Mater. Res., 2006, 21, 1237–1242. 29 R. A. Boulos, C. Harnagea, X. Duan, R. N. Lamb, F. Rosei and C. L. Raston, RSC Adv., 2013, 3, 3284– 3290. 30 E. S. Tjandra, R. A. Boulos, P. C. Duncan and C. L. Raston, CrystEngComm, 2013, 15, 6896–6900. 31 S. Berland, O. Delattre, S. Borzeix, Y. Catonné and E. Lopez, Biomaterials, 2005, 26, 2767–2773. 32 E. Tamariz and F. Grinnell, Mol. Biol. Cell, 2002, 13, 3915–3929. 33 Q. M. Chen, V. C. Tu, J. Catania, M. Burton, O. Toussaint and T. Dilley, J. Cell Sci., 2000, 113, 4087–4097. 34 A. Moustakas and C. Stournaras, J. Cell Sci., 1999, 112, 1169–1179. 35 W. Malorni, F. Iosi, F. Mirabelli and G. Bellomo, Chem. Biol. Interact., 1991, 80, 217–236. This journal is © The Royal Society of Chemistry 2014 Communication 36 G. Bellomo, F. Mirabelli, M. Vairetti, F. Iosi and W. Malorni, J. Cell. Physiol., 1990, 143, 118–128. 37 T. J. Collins, BioTechniques, 2007, 43, S25–S30. 38 J. Varani, M. K. Dame, L. Rittie, S. E. G. Fligiel, S. Kang, G. J. Fisher and J. J. Voorhees, Am. J. Pathol., 2006, 168, 1861–1868. 39 G. J. Fisher, T. Quan, T. Purohit, Y. Shao, M. K. Cho, T. He, J. Varani, S. Kang and J. J. Voorhees, Am. J. Pathol., 2009, 174, 101–114. 40 R. E. Barrow and M. R. K. Dasu, J. Surg. Res., 2005, 126, 59–65. 41 M. Dalby, M. Riehle, H. Johnstone, S. Affrossman and A. Curtis, Tissue Eng., 2002, 8, 1099–1108. 42 K. Burridge and M. Chrzanowska-Wodnicka, Annu. Rev. Cell Dev. Biol., 1996, 12, 463–519. 43 R. J. Petrie, A. D. Doyle and K. M. Yamada, Nat. Rev. Mol. Cell Biol., 2009, 10, 538–549. Toxicol. Res., 2014, 3, 223–227 | 227 Electronic Supplementary Material (ESI) for Toxicology Research. This journal is © The Royal Society of Chemistry 2014 Evaluating the effects of nacre on human skin and scar cells in culture Vipul Agarwal,a Edwin S. Tjandra,a K. Swaminathan Iyer,a Barry Humfrey,b Mark Fear,c Fiona M. Wood,c Sarah Dunlopd and Colin L. Rastone, * Supporting Information S1. Experimental S1.1 Materials Pinctada margaritifera shells were provided by Pearl Technologies Pty Ltd, which were grown in the waters of the Abrolhos Islands, Western Australia. The decalcified organic conchiolin layer was removed by wet sand blasting of the shell followed by gentle brushing to remove any dust particles that might otherwise contaminate the samples. Inner nacreous layer was then scraped using a surgical scalpel and stored at room temperature for a maximum of 2 weeks. S1.2 Scanning Electron Microscopy Scraped nacre above from the inner layer of the shell was mounted on SEM stubs (ProSciTech, Cat.# G040). Samples were coated with 4 nm of platinum. Images were taken using scanning electron microscope (Zeiss 1555, VP-FESEM) at 4-5 kV at 30µm aperture. Images were analyzed with the image analysis software ImageJ (NIH).1 S1.3 Cell culture A human derived immortalized keratinocyte cell line, HaCaT 2 and two human primary dermal (fibroblast) cell cultures from normal skin and normal scar were used. All three cell types were cultured in Dulbecco’s Modified Eagle’s Medium (DMEM/F12 - GlutaMAX; Invitrogen Gibco) supplemented with 10% fetal bovine serum (FBS; Invitrogen Gibco) and 1% penicillin/streptomycin (Invitrogen Gibco). The cells were incubated at 37°C in an atmosphere of 5% CO2. Primary cells used were between the passages 7-10. Nacre was sterilized using UV sterilization technique in the tissue culture hood for 15 min prior its dissolution in media. Fresh nacre solution was prepared before every experiment. S1.4 Cell Viability Cell viability was determined using a LIVE/ DEAD Viability/Cytotoxicity Kit (Invitrogen, UK) which measures the membrane integrity of cells,3, 4 as per manufacturer’s protocol. In brief, 20000 cells were seeded in each well in a 24 well plate and treated with scraped nacre at various concentrations in cell culture media (DMEM F-12 containing 10% FBS and 1% Penicillin/ Streptomycin) and incubated for 24 h or 72 h in the humidified incubator at 37°C with 5% CO2. At the stipulated time (24h and 72h), cells were washed with PBS (3 times) and then stained with calcein (100 µL, 1µM)/ ethidium bromide (100 µL, 2 µM) in PBS and incubated in the humidified incubator at 37°C and 5% CO2 for 30 min. Images were captured using an Olympus IX71 inverted microscope with a 20 x objective with fixed exposure time. Both live and dead cells were counted using Image J with cell counter plug in. Experiments were performed in triplicate. Minimum of fifty images were captured per condition. S1.5 Reactive Oxygen Species (ROS) ROS was measured using the ROS assay kit (Oxiselect ROS assay kit, Cat.# STA 342, Cell Biolabs) following manufacturer’s protocol. In brief, 6000 cells were seeded in a 96 well plate and incubated in the humidified incubator at 37°C with 5% CO2 for 24h. Next day, wells were washed with PBS (3 times) and incubated with 2’, 7’dichlorodihydrofluorescin diacetate (DCFH-DA) solution (0.1 x, 100 µL/ per well) for 1 h in the humidified incubator at 37°C with 5% CO2. DCFH-DA solution was removed and the wells washed with 3 x PBS. Cells were then treated with scraped nacre solution in culture media at a specified concentration for 24 h, wells were washed with 3 x PBS and cells were lysed using the lysis buffer provided (1 x, 100µL/ per well, incubated for 20 min. at room temperature) before reading the plate at 480 nm excitation/ 530 nm emissions using the plate reader. Experiments were performed in triplicate. S1.6 Cell Body Area Cell size was measured using Image J software (NIH).1 A minimum of 25 cells were randomly selected from the fluorescence images and their area was measured. Values reported as mean ± standard error mean. S1.7 Statistics The results for cell viability, ROS experiments and cell area are expressed as mean ± standard error mean (SEM) and analysed by analysis of variance (ANOVA). Significance was evaluated using Bonferroni and Turkey’s post-hoc analysis and set at 95% confidence (p < 0.05). Supporting Figures Figure S1: Cell morphology post calcein AM/ ethidium bromide staining and imaged using fluorescent microscopy. HaCaT cells were treated with various concentrations of nacre for 24h, stained and imaged. a) Untreated (control), b) HaCaT cells treated with 2.5 mg/mL nacre. Scale bar 1µm. Figure S2: Cell area size showing increase in the cell area post incubation with nacre. Cell area was measured from the fluorescent images of live cells taken for viability assay. ‘None’ is the untreated control. Data presented as average ± SEM (n>25). Significance was set at * p < 0.05 using bonferroni post hoc test in one way ANNOVA Figure S3: Reactive oxygen species (ROS) assay showing increase in ROS levels in cells stressed with various concentrations of H2O2 in dose dependent manner. Human dermal skin fibroblasts cells were incubated with various concentrations of H2O2 for the specified period of time to generate the standard curve. Cells were then incubated with 2’, 7’- dichlorodihydrofluorescin diacetate (DCFH-DA) solution which fluoresce in the presence of reactive oxygen species. ‘None’ is the untreated control. Data presented as average ± SEM (n=3). References 1. 2. 3. 4. T. J. Collins, BioTechniques, 2007, 43, S25-S30. P. Boukamp, R. T. Petrussevska, D. Breitkreutz, J. Hornung, A. Markham and N. E. Fusenig, J. Cell Biol., 1988, 106, 761-771. L. W. Zhang and N. A. Monteiro-Riviere, Toxicol. Sci., 2009, 110, 138-155. I. Gotman and S. Fuchs, in Active Implants and Scaffolds for Tissue Regeneration, ed. M. Zilberman, Springer Berlin Heidelberg, 2011, vol. 8, pp. 225-258. RSC Advances Published on 13 November 2012. Downloaded by University of Western Australia on 09/10/2014 10:17:39. COMMUNICATION Cite this: RSC Advances, 2013, 3, 1009 Received 4th October 2012, Accepted 12th November 2012 DOI: 10.1039/c2ra22402j www.rsc.org/advances View Article Online View Journal | View Issue Biogenic production of palladium nanocrystals using microalgae and their immobilization on chitosan nanofibers for catalytic applications3 Ela Eroglu,ab Xianjue Chen,a Michael Bradshaw,a Vipul Agarwal,a Jianli Zou,a Scott G. Stewart,c Xiaofei Duan,d Robert N. Lamb,d Steven M. Smith,*b Colin L. Raston*a and K. Swaminathan Iyera Spherical palladium nanocrystals were generated from aqueous Na2[PdCl4] via photosynthetic reactions within green microalgae (Chlorella vulgaris). Electrospun chitosan mats were effective for immobilizing these biogenic nanocrystals, as a material for recycling as a catalyst for the Mizoroki–Heck cross-coupling reaction. This photosynthetically-driven metal transformation system can serve as a good candidate for an environmentallyfriendly method for the synthesis of metal nanocatalysts. Biogenic systems can control the phase, structure, and topography of inorganic nanocrystals with a level of precision similar to synthetic approaches.1 Various organisms such as bacteria,2 yeast,3 fungi4 and algae5 are capable of processing a wide range of metals. Such bioprocessing has been effectively used in the reduction of environmental pollution and also for the recovery of metals from waste.6 The formation of metal nanoparticles in the presence of microorganisms primarily involves the reduction of metal ions in solution by enzymes generated by microbial cell activities, which can be intracellular or extracellular. This depends on the nanoparticle crystallization site, as being either inside the cell or on the cell surface.7,8 When considering palladium, the biosynthesis of nanoparticles of the metal have been reported in bacteria (Desulfovibrio desulfuricans, Shewanella oneidensis, Bacillus sphaericus),9 cyanobacteria (Plectonema boryanum UTEX 485),10,11 plants,12–14 and viruses (e.g., tobacco mosaic virus).15,16 Herein we report a biogenic synthesis of palladium nanocrystals in the presence of Chlorella vulgaris with the photoautotrophic microa Centre for Strategic Nano-Fabrication, School of Chemistry and Biochemistry, The University of Western Australia, M313, 35 Stirling Highway, Crawley, WA 6009, Australia. E-mail: colin.raston@uwa.edu.au; Fax: +61 8 6488 8683; Tel: +61 8 6488 3045 b ARC Centre of Excellence in Plant Energy Biology, The University of Western Australia, M313, 35 Stirling Highway, Crawley, WA 6009, Australia. E-mail: steven.smith@uwa.edu.au; Fax: +61 8 6488 4401; Tel: +61 8 6488 4403 c School of Chemistry and Biochemistry, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia d Surface Science & Technology Group, School of Chemistry, The University of Melbourne, VIC 3010, Australia 3 Electronic supplementary information (ESI) available: Experimental details, characterization, and an additional TEM image. See DOI: 10.1039/c2ra22402j This journal is ß The Royal Society of Chemistry 2013 algal metabolism most likely providing the necessary reducing agents. Studies on the biogenic reduction of palladium by various organisms are usually limited to the generation of Pd(0). We have used a non-toxic and environmentally-available microorganism to achieve this, and then collected the particles from the liquid culture via immobilizing on electrospun chitosan nanofibers. In addition, we have established the utility of this composite material as a recyclable catalyst which is functional even at very low loading rates. The overall integrated approach is without precedent, as are the individual components. Algae are a large group of photosynthetic organisms that are ubiquitous in various aquatic habitats including sea, freshwater, wastewater and also in moist solids.17 Microalgal remediation has been reported for several metal ions,18,19 and involves both intracellular (polyphosphates) and extracellular (polysaccharides on algal cell wall) metal binding groups.20–22 Based on these observations, we initially evaluated the ability of a common green microalga, Chlorella vulgaris, to reduce palladium(II) to Pd(0) starting with a solution of Na2[PdCl4]. Various concentrations of Pd salt were investigated (100, 50, 25, 12.5, 0 mg L21) in order to detect their effects on cell viability. The cells were grown in solutions of Na2[PdCl4] added to algal freshwater media (MLA media)23 at 25 uC and a rotation speed of 120 rpm, under diurnal illumination of 16 h light/8 h dark cycles, with the total chlorophyll content used to validate the viability of the cell cultures.24 Accumulation of total chlorophyll pigment (Chl a + Chl b) as a function of growth time was monitored (Fig. 1a), and the threshold concentration for toxicity on the cells established amongst 25–50 mg L21. Within the first five days, the culture flasks containing less than this concentration clearly showed an increase in cellular growth with solution retaining the green color which is associated with the chlorophyll content of the cells. In contrast, higher concentrations of Na2[PdCl4] resulted in the color fading and loss of cell viability (Fig. 1b). The solution containing 25 mg L21 of Na2[PdCl4] had the highest amount of palladium precursor while showing little reduction in growth rates compared to the control. Thus, it was chosen as the prototype for further experiments. RSC Adv., 2013, 3, 1009–1012 | 1009 View Article Online Published on 13 November 2012. Downloaded by University of Western Australia on 09/10/2014 10:17:39. Communication Fig. 1 (a) Total amount of accumulated chlorophyll (Chl a + Chl b) measured as a function of growth time for Chlorella vulgaris cultures in the presence of various concentrations of Na2[PdCl4] (100, 50, 25, 12.5 mg L21), and control cultures in the absence of Na2[PdCl4] (0 mg L21, hollow circles), with the total volume of solution in all flasks at 40 mL during the growth experiments. (b) Chlorella vulgaris culture flasks containing various concentrations of Na2[PdCl4]. Transmission electron microscopy (TEM) of the liquid culture media revealed the presence of crystalline spherical palladium nanoparticles with an average particle size of around 7 nm, with a range of 2 to 15 nm in diameter (Fig. 2a). The corresponding electron diffraction pattern further confirmed the presence of elemental Pd nanocrystals with a characteristic face-centered cubic (fcc) structure and average d-spacing values of 0.22, 0.19, 0.14, and 0.12 nm for the (111), (200), (220) and (311) planes, respectively (Fig. 2b). High resolution TEM analysis further confirmed the presence of single palladium nanocrystals (Fig. 2c and d). Reducing agents produced by photosynthesis are the key components for the reduction of Pd(II) into Pd(0) nanoparticles (Scheme 1). Photosynthetic processes in green microalgae can take place either under oxygenic or anoxic environments.17,25 Green algae have chlorophyll-a and chlorophyll-b pigments, and can Fig. 2 (a) TEM image of palladium nanoparticles precipitated in the four week old microalgae solution (scale bar: 20 nm). (b) Selective area electron diffraction pattern of the palladium nanoparticles in 2(a), giving d-spacings of 0.22, 0.19, 0.14, and 0.12 nm which correspond to 111, 200, 220, and 311 reflections, respectively, for Pd(0). (c) High resolution TEM image of palladium nanoparticles. (d) Pd(0) nanocrystals with (111) lattice spacings of around 0.22 nm. Inset shows the Fast Fourier Transform (FFT) pattern corresponding to the area shown within the yellow rectangle in 2(c), indicating the crystal structure of palladium nanoparticles. 1010 | RSC Adv., 2013, 3, 1009–1012 RSC Advances Scheme 1 Palladium nanoparticle synthesis by photosynthetic green microalgae, and their uptake on an electrospun chitosan mat for use as a catalyst in Mizoroki– Heck reactions. The left stage shows a combination of mechanisms taking place within the photosynthetic organisms, resulting in the production of reducing agents.17,25,26 (ADP: adenosine diphosphate, ATP: adenosine triphosphate, Fd: ferredoxin, NADP+: oxidized form of nicotinamide adenine dinucleotide phosphate, NADPH: reduced form of nicotinamide adenine dinucleotide phosphate, PGA: phosphoglycolic acid, RuBisCO: rubilose biphosphate carboxylase). NADPH is likely one of the main reducing agents for the reduction of Na2[PdCl4], which is partially oxidised as a result of aerobic culture conditions. accomplish oxygenic (oxygen-evolving) photoautotrophic reactions while using H2O as an electron donor.17 Oxygenic photoautotrophic processes include two sets of photochemical reactions: (i) light reactions conserving chemical energy in the form of adenosine triphosphate (ATP), and the reduced form of nicotinamide adenine dinucleotide phosphate (NADPH), and (ii) ‘dark’ reactions in which CO2 is reduced to organic compounds using ATP and NADPH. Light reactions have two separate sets of photosystems (PS), namely PS I and PS II, in which the electron transfer follows a Z-scheme between these two photosystems. PS II is mainly responsible for the splitting of water (H2O A KO2 + 2e2 + 2H+) as a first stage of cyclic electron flow. While electron flow follows the Z-scheme between PS II and PS I, a proton motive force is created and used for the generation of ATP. CO2 is fixed by the enzyme ribulose bisphosphate carboxylase (RuBisCO) and reduced in the Calvin cycle using NADPH.17 Reducing equivalents can be exported from the chloroplast in the form of carbon metabolites, particularly triose phosphates (triose-P).26 Their oxidation by dehydrogenases in the cytosol can also generate NADH and NADPH for other reduction reactions. We postulate that such reducing agents promote the reduction of Pd(II) into crystalline Pd(0) nanoparticles, which is further partially oxidized due to the aerobic culture conditions. An hypothesis for the reduction of palladium(II) by photoheterotrophic bacteria (Rhodobacter sphaeroides) cultures involves the reduced electron carriers (such as ferredoxin, NADH, and FADH) as the electron donors.27 Consistent with this hypothesis, we found that addition of NADPH to the MLA growth medium23 resulted in reduction of Na2[PdCl4] to produce Pd nanocrystals in solution which were effectively recovered on exposure to cross-linked electrospun chitosan nanofiber mats (see ESI3, Fig. S1). This journal is ß The Royal Society of Chemistry 2013 View Article Online Published on 13 November 2012. Downloaded by University of Western Australia on 09/10/2014 10:17:39. RSC Advances Chitosan is derived from chitin present in the exoskeletons of crustaceans,28 and is an attractive, renewable feedstock which is effective in binding a number of metal ions,29 and effective as a support for Pd, Ni, Cu, Cr and Zn based catalysts.30 Cross-linked chitosan mats (3 6 2 cm rectangular shape) generated using electrospinning31,32 were placed into a four week old microalgal culture containing 25 mg L21 Na2[PdCl4]. A control experiment involved exposing chitosan mats to 25 mg L21 Na2[PdCl4] solution in the absence of microalgae. Chitosan mats were kept in the solution for about two weeks as to achieve sufficient particle adsorption from the liquid media. Scanning electron microscopy (SEM) revealed the nanofiber structure of the chitosan mats with an average diameter of 100 nm, which was maintained after exposure to the growth media (Fig. 3a). It is noteworthy that there was a detectable difference in color of the recovered chitosan mats from the microalgae and the control solutions. The initial light yellow chitosan mat turned dark brown after being incubated with algae (Fig. 3b), whereas no color change was observed for the control experiment without algae (Fig. 3c). The darker color is considered to be consistent with the immobilization of the reduced Pd nanocrystals from the growth media, and corresponds to approximately 1.03 wt% Pd loading on each rectangular piece of electrospun chitosan mat (3 6 2 cm), which was established using inductively coupled plasma-optical emission spectroscopy (ICP-OES). X-ray photoelectron spectrometric (XPS) analysis of the Pd loaded chitosan mats showed a dominating peak in the PdO 3d5/2 region (binding energy of 337.9 eV) and also the presence of the Pd(0) peak in the Pd 3d5/2 region Fig. 3 (a) SEM image of the as-prepared electrospun chitosan nanofibers before being exposed to palladium solution (scale bar: 1 mm). Mats of this material placed into 4 week old 25 mg L21 Na2[PdCl4] solutions with and without microalgae which were collected after two weeks are shown in (b) and (c) respectively. Curve-fitted Pd 3d XPS spectra for (d) as collected and (e) etched internal surface of a chitosan mat under argon ions for 60 s, both showing Pd(0) (blue curve) and PdO (red curve). This journal is ß The Royal Society of Chemistry 2013 Communication (binding energy of 336.0 eV) (Fig. 3d). Analysis of the sample following argon ion etching primarily showed a dominant Pd(0) peak in the Pd 3d5/2 region (binding energy of 335.9 eV) with an additional lower peak of PdO in the 3d5/2 region (binding energy of 337.7 eV) (Fig. 3e). Thus the XPS data establish that the surface of the palladium nanoparticles is partially oxidized under the aerobic culture conditions. Slight oxidation was also reported when palladium containing samples were exposed to an oxygen containing atmosphere before XPS analysis.33 Chlorine was not detected during these XPS analyses, indicating the absence of palladium(II) chloride species. The binding energies of both Pd 3d5/2 and PdO 3d5/2 were found to be slightly higher than the values for pure Pd metal,34 and PdO samples.35 This slight shift is reported to be common when the Pd/PdO nanoparticles are embedded in an insulating substrate,36 which is the chitosan mat in the present case. In their study, Schildenberger and his colleagues also reported similar peaks for Pd 3d5/2 (336.0 eV) and PdO 3d5/2 (337.9 eV) due to the isolated arrangement of metal nanoparticles on the layers of oxidized silicon wafers.36 The utility of these palladium loaded chitosan mats were tested as catalyst supports for the standard Mizoroki–Heck reaction.37,38 Six dried mats (a total of 0.23% mol Pd per mol of iodobenzene) were introduced into a solution containing iodobenzene, butyl acrylate, triethylamine and dimethylformamide (DMF). The reaction temperature was kept constant at 80 uC for 16 h, and after each reaction, the supported catalyst was recovered for recycling studies after washing with DMF under nitrogen gas to avoid any oxygen induced regrowth of Pd(0) nanoparticles.39 The catalyst can be recycled at least four times with quantitative conversion for each cycle, with the conversion yields of: 68% (1st cycle), 62% (2nd cycle), 45% (3rd cycle) and 36% (4th cycle) by weight. For the control comparison, a mat exposed to a Na2[PdCl4] solution without algae resulted in a conversion yield of only 5%. We have previously reported that a chitosan mat containing palladium nanoparticles generated by reduction of pre-absorbed Na2[PdCl4] is also an effective catalytic support with no appreciable leaching of the metal.31 Conventional palladium catalysts usually require a 1–5 mol% loading rate for effective Mizoroki–Heck crosscoupling reactions.40 Our biogenic palladium nanocatalysts have high catalytic activity (68%) with respect to commercial material (5%), even at low palladium loadings (0.23 mol%), which is significant for applications in the fine chemical industries. In conclusion we have established a biogenic synthesis of palladium nanocrystals in the presence of Chlorella vulgaris and their subsequent immobilization on an electrospun chitosan mat as a novel support for application in catalysis. In addition, we have demonstrated that NADPH involved in the photoautotrophic microalgal metabolism is likely to play a role in the biogenic synthesis. Utilization of easy-to-grow, nontoxic and environmentally-available microalgae for the synthesis of palladium has potential for the development of green chemistry processes for other metals. RSC Adv., 2013, 3, 1009–1012 | 1011 View Article Online Communication Published on 13 November 2012. Downloaded by University of Western Australia on 09/10/2014 10:17:39. Acknowledgements We kindly acknowledge support of this work by the Australian Research Council (ARC) and internal grants of The University of Western Australia. The microscopy analysis was carried out using the facilities in the Centre for Microscopy, Characterization and Analysis (The University of Western Australia). The authors also would like to thank Dr L. Byrne for his kind help during the NMR analysis. Notes and references 1 D. Cui and H. Gao, Biotechnol. Prog., 2003, 19, 683–692. 2 T. Y. Beveridge and R. G. E. Murray, J. Bacteriol., 1980, 141, 876–887. 3 C. P. Huang, C. P. Juang, K. Morehart and L. Allen, Water Res., 1990, 24, 433–439. 4 B. Volesky and Z. R. Holan, Biotechnol. Prog., 1995, 11, 235–250. 5 D. W. Darnall, B. Greene, M. J. Henzel, M. Hosea, R. A. McPherson, J. Sneddon and M. D. Alexander, Environ. Sci. Technol., 1986, 20, 206–208. 6 P. Yong, I. P. Mikheenko, K. Deplanche, M. D. Redwood and L. E. Macaskie, Biotechnol. Lett., 2010, 32, 1821–1828. 7 K. Simkiss and K. Wilbur, Biomineralization. Cell Biology and Mineral Deposition, Academic Press, Inc., San Diego, 1989. 8 S. Mann, Biomimetic Materials Chemistry, VCH, Weinheim, New York, 1996. 9 P. Yong, N. A. Rowson, J. P. G. Farr, I. R. Harris and L. E. Macaskie, Biotechnol. Bioeng., 2002, 80, 369–379. 10 J. R. Lloyd, P. Yong and L. E. Macaskie, Appl. Environ. Microbiol., 1998, 64, 4607–4609. 11 M. F. Lengke, M. E. Fleet and G. Southam, Langmuir, 2007, 23, 8982–8987. 12 X. Yang, Q. Li, H. Wang, J. Huang, L. Lin, W. Wang, D. Sun, Y. Su, J. B. Opiyo, L. Hong, Y. Wang, N. He and L. Jia, J. Nanopart. Res., 2010, 12, 1589–1598. 13 X. Yang, Z. Fei, D. Zhao, W. H. Ang, Y. Li and P. J. Dyson, Inorg. Chem., 2008, 47, 3292–3297. 14 M. N. Nadagouda and R. S. Varma, Green Chem., 2008, 10, 859–862. 15 A. K. Manocchi, N. E. Horelik, B. Lee and H. Yi, Langmuir, 2010, 26, 3670–3677. 16 C. X. Yang, A. K. Manocchi, B. Lee and H. M. Yi, J. Mater. Chem., 2011, 21, 187–194. 17 M. T. Madigan, J. M. Martinko and J. Parker, Brock: Biology of Microorganisms, 9th edn, Prentice Hall Inc., New Jersey, 2000. 1012 | RSC Adv., 2013, 3, 1009–1012 RSC Advances 18 A. Malik, Environ. Int., 2004, 30, 261–278. 19 B. Suresh and G. A. Ravishankar, Crit. Rev. Biotechnol., 2004, 24, 97–124. 20 J. L. Gardea-Torresdey, M. K. Becker-Hapak, J. M. Hosea and D. W. Darnall, Environ. Sci. Technol., 1990, 24, 1372–1378. 21 D. Kaplan, D. Christiaen and S. M. Arad, Appl. Environ. Microbiol., 1987, 53, 2953–2956. 22 W. Zhang and V. Majidi, Environ. Sci. Technol., 1994, 28, 1577–1581. 23 R. A. Andersen, Algal Culturing Techniques, Elsevier Academic Press, 2005. 24 H. K. Lichtenthaler and C. Buschmann, in Current protocols in food analytical chemistry, ed. R. E. Wrolstad, John Wiley & Sons Inc., New York, 2001, pp. F4.3.1–8. 25 E. Eroglu and A. Melis, Bioresour. Technol., 2011, 102, 8403–8413. 26 M. Taniguchi and H. Miyake, Curr. Opin. Plant Biol., 2012, 15, 252–260. 27 M. D. Redwood, Bio-hydrogen and biomass supported palladium catalyst for energy production and waste minimisation, PhD Thesis, The University of Birmingham, 2007. 28 M. Rinaudo, Prog. Polym. Sci., 2006, 31, 603–632. 29 E. Guibal, Sep. Purif. Technol., 2004, 38, 43–74. 30 A. B. Sorokin, F. Quignard, R. Valentin and S. Mangematin, Appl. Catal., A, 2006, 309, 162–168. 31 M. Bradshaw, J. Zou, L. Byrne, K. S. Iyer, S. G. Stewart and C. L. Raston, Chem. Commun., 2011, 47, 12292–12294. 32 K. Ohkawa, D. I. Cha, H. Kim, A. Nishida and H. Yamamoto, Macromol. Rapid Commun., 2004, 25, 1600–1605. 33 J. Zou, K. S. Iyer, S. G. Stewart and C. L. Raston, New J. Chem., 2011, 35, 854–860. 34 A. Tressaud, S. Khairoun, H. Touhara and N. Watanabe, Z. Anorg. Allg. Chem., 1986, 540, 291–299. 35 T. H. Fleisch, G. W. Zajac, J. O. Schreiner and G. J. Mains, Appl. Surf. Sci., 1986, 26, 488–497. 36 M. Schildenberger, R. Prins and Y. C. Bonetti, J. Phys. Chem. B, 2000, 104, 3250–3260. 37 T. Mizoroki, K. Mori and A. Ozaki, Bull. Chem. Soc. Jpn., 1971, 44, 581. 38 R. F. Heck and J. P. Nolley, Jr., J. Org. Chem., 1972, 37, 2320–2322. 39 J. Zou, S. G. Stewart, C. L. Raston and K. S. Iyer, Chem. Commun., 2011, 47, 1803–1805. 40 B. M. Bhanage and M. Arai, Catal. Rev. Sci. Eng., 2001, 43, 315–344. This journal is ß The Royal Society of Chemistry 2013 Electronic Supplementary Material (ESI) for RSC Advances This journal is © The Royal Society of Chemistry 2012 Supplementary Information Biogenic production of palladium nanocrystals using microalgae and their immobilization on chitosan nanofibers for catalytic applications Ela Eroglua,b, Xianjue Chena, Michael Bradshawa, Vipul Agarwala , Jianli Zoua, Scott G. Stewartc, Xiaofei Duand, Robert N. Lambd, Steven M. Smithb*, Colin L. Rastona*, and K. Swaminathan Iyera a Centre for Strategic Nano-Fabrication, School of Chemistry and Biochemistry, The University of Western Australia, Crawley, WA 6009, Australia, E-mail: colin.raston@uwa.edu.au b ARC Centre of Excellence in Plant Energy Biology, The University of Western Australia M313, 35 Stirling Highway, Crawley, WA 6009, Australia, E-mail: steven.smith@uwa.edu.au c School of Chemistry and Biochemistry, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia d Surface Science & Technology Group, School of Chemistry, The University of Melbourne, VIC 3010, Australia S1. Biosynthesis of Palladium Nanoparticles Wild type Chlorella vulgaris cultures (from the Australian National Algae Culture Collection at CSIRO, Tasmania) were used as the green microalgae source for the biosynthesis of palladium. Sterile algal freshwater media (MLA media)[1] containing standard micronutrients, nitrate, phosphate, carbonate buffer and vitamins was used as the substrate source. The microalgae cultures were mixed with Na 2 [PdCl 4 ] solution at various concentrations (100, 50, 25, 12.5, 0 mg/L), towards reaching an initial total-chlorophyll content of around 1.8 mg/L (Figure 1a). Na 2 [PdCl 4 ] solution was prepared by dissolving Na 2 [PdCl 4 ] powder (Sigma-Aldrich) in distilled and sterilized water overnight, and subsequently filtered through a sterile Pall® Acrodisc® 32 mm syringe filter (0.2 µm membrane) for further sterilization. Concentrations of Na 2 [PdCl 4 ] in microalgae solutions are reported here as the initial concentrations measured by the gravimetric analysis before the filtration process. The experiments were conducted under batch conditions and cyclic diurnal conditions (16 h light/8 h dark) at a constant temperature (25 °C). Algae cultures (total liquid volume of 40 mL) were grown in 250 mL Erlenmeyer flasks, under continuous cool-white fluorescent illumination at an incident intensity of around 200 µmol photons m-2s-1(PAR) upon orbital shaking (Thermoline Electronic Supplementary Material (ESI) for RSC Advances This journal is © The Royal Society of Chemistry 2012 Scientific) at 120 rpm. Algal growth was investigated by measuring the total chlorophyll content (Chl a + Chl b) with respect to time, following the spectrophotometric method involving methanol extraction.[2] Chlorophyll content was also used to validate the viability of the cell cultures. All experiments were conducted in triplicates, with the standard deviation of each value given in the form of error bars within the related figure. S2. Characterization A JEOL 2100 TEM instrument operating at 80 kV was used for determining the size and morphology of chitosan nanofibers and palladium nanoparticles. Samples were prepared by inserting the solutions on top of carbon-coated 200 mesh copper grids and allowing them to dry. High resolution images were obtained using a JEOL 3000F instrument operating at 300 kV. A Zeiss 1555 VP-FESEM with a 3 kV accelerating voltage was used to image samples coated with platinum (~3nm). XPS data was acquired using a VG ESCALAB220i-XL X-ray Photoelectron Spectrometer equipped with a hemispherical analyzer. The incident radiation was monochromatic Al Kα X-rays (1486.6 eV) at 220 W (22 mA and 10kV). Survey (wide) and high resolution (narrow) scans were taken at analyzer pass energies of 100 eV and 50 eV, respectively. Survey scans were carried out over 1200-0 eV binding energy range with 1.0 eV step size and 100 ms dwell time. Narrow high resolution scans were run over a 20 eV binding energy range with 0.05 ev step size and 250 ms dwell time. Base pressure in the analysis chamber was 4.0x10-9 mbar and during sample depth profile analysis 1.5x10-7 mbar. A low energy flood gun (~6 eV) was used to compensate the surface charging effect. Argon ions at 3 keV beam energy were used to sputter off approximately 18 nm surface layers at a rate of ~3 Angstrom/second. The ion source gave a crater of approximately 3x3 mm. The energy calibration was referenced to the C 1s peak at 284.7 eV. S3. Cross Coupling Reactions A rectangular piece of electrospun chitosan mat (3 x 2 cm per each piece) was placed into four week old 25 mg/L Na 2 [PdCl 4 ] containing microalgae solutions (total volume: 40 mL per flask) for two weeks. The amount of palladium was established using inductively coupled plasma - optical emission spectroscopy (ICP-OES), at ~1.03 % w/w per each rectangular mat (3x2 cm). ICP-OES analysis was acquired with ARL-3520B sequential scanning ICP-OES, Electronic Supplementary Material (ESI) for RSC Advances This journal is © The Royal Society of Chemistry 2012 with a 20mm torch, 1200W incident power, and a MDSN nebuliser. For the Mizoroki-Heck reactions iodobenzene (55 µL; 0.49 mmol), butyl acrylate (85 µL; 0.59 mmol), triethylamine (171 µL; 1.23 mmol) in dimethylformamide (DMF, 0.5 mL) were added to 6 hybrid palladium nanoparticle-chitosan mats with a total of 0.23% mol Pd per mol iodobenzene. The reaction mixture was heated to 80°C for 16 hours, whereupon the catalyst mat was filtered and recovered for its recycling after washing with DMF five times under nitrogen gas. Quantitative conversion yields were assessed gravimetrically upon the confirmation of the final product (butyl cinnamate) using 1H and 13 C NMR spectroscopy (Varian® 400 NMR). S4. Electrospinning of Chitosan Pd was collected by electrospun chitosan mat following the electrospinning procedure optimized by Bradshaw et al. (2011)[3], which was slightly modified from the original protocol given by Ohkawa et al. (2004)[4]. The variables of electrospinning processes were as follows: (i) syringe pump speed: 0.1 mm/min, (ii) voltage: 18 kV, (iii) distance between the target and the tip of the syringe: 11 cm, (iv) target speed: 1 m/ min, (v) traverse speed: 0.5 cm/min.[3] Chitosan (2-amino-2-deoxy-(1-4)-β-D-glucopyranose), 75-85% deacetylated (Sigma-Aldrich), powder (6% wt) was mixed with TFA (trifluoroacetic acid): DCM (dichloromethane) solution (70:30 v/v), and stirred overnight for its complete dissolution. These ratios were taken to be optimum for attaining ultrafine chitosan nanofibers (Bradshaw et al. 2011).[3] TFA (trifluoroacetic acid) and DCM (dichloromethane) were obtained from Chem Supply. Before the electrospinning process, the mixture was sonicated for 15 min and 5.4% v/v glutaraldehyde (25% in H 2 O, Sigma-Aldrich) was immediately added to the chitosan/TFA/DCM solution for an effective crosslinking. Electrospinning of this crosslinked solution created insoluble nanofibers. Two rectangular pieces of electrospun chitosan mat (3 x 2 cm per each piece) were placed into one flask of (four weeks old) 25 mg/L Na 2 [PdCl 4 ] containing microalgae solutions (total volume of 40 mL per flask) and left inside for two weeks. The amount of palladium was established by using ICP-OES, yielding ∼1.03 % w/w per each rectangular mat (3x2 cm). Electronic Supplementary Material (ESI) for RSC Advances This journal is © The Royal Society of Chemistry 2012 S5. Reduction of Palladium with NADPH Figure S1. High resolution TEM images of Pd(0) nanoparticles after the inoculation of Na 2 [PdCl 4 ] precursor with an excess amount of NADPH, dissolved inside MLA algal growth-media,[1] indicating high crystallinity with (111) lattice constant of around 0.22 nm (inset: FFT pattern corresponding to the area shown with red rectangle). References 1. R.A. Andersen, in Algal Culturing Techniques, Elsevier Academic Press, 2005. 2. H.K. Lichtenthaler and C. Buschmann, in Current protocols in food analytical chemistry, eds: R.E. Wrolstad RE, John Wiley & Sons Inc., New York, 2001, pp. F4.3.1- 8. 3. M. Bradshaw, J. Zou, L. Byrne, K.S. Iyer, S.G. Stewart, C.L. Raston, Chem. Commun. 2011, 47, 12292-12294. 4. K. Ohkawa, D. I. Cha, H. Kim, A. Nishida, H. Yamamoto, Macromol. Rapid Commun. 2004, 25, 1600–1605. View Article Online / Journal Homepage / Table of Contents for this issue Green Chemistry Cite this: Green Chem., 2012, 14, 2682 Published on 08 August 2012. Downloaded by University of Western Australia on 09/10/2014 10:19:02. www.rsc.org/greenchem Dynamic Article Links COMMUNICATION Nitrate removal from liquid effluents using microalgae immobilized on chitosan nanofiber mats Ela Eroglu,a,b Vipul Agarwal,a Michael Bradshaw,a Xianjue Chen,a Steven M. Smith,*b Colin L. Raston*a and K. Swaminathan Iyera Received 25th June 2012, Accepted 8th August 2012 DOI: 10.1039/c2gc35970g Mats of electrospun chitosan nanofibers were found effective in immobilizing microalgal cells. These immobilized microalgal cells were also used as a durable model system for wastewater treatment which has been demonstrated by the removal of around 87% nitrate from liquid effluents ([NO3−N]initial 30 mg L−1). Virtuous nitrate removal rates were derived by superimposing physicochemical adsorption and biological nutrient consumption phenomena for chitosan and microalgae, respectively. Wastewater treatment focuses on eliminating unwanted chemicals and/or biological impurities from contaminated water. In general the treatment methods are mainly based on the separation of pollutants from the wastewater with a requirement for a further processing stage to eliminate these pollutants.1 Integrated wastewater treatment processes are important in eliminating undesired species, ideally converting them into valuable products. As relatively recent bioprocesses, algal cultivation in wastewaters has a combination of several advantages such as wastewater treatment and simultaneous algal biomass production, which can be further exploited for biofuel production, food additives, fertilizers, cosmetics, pharmaceuticals, and other valuable chemicals.2 However, there are inherent difficulties associated with algae-based bioprocessing in the harvesting, dewatering and processing of the algal biomass. Immobilization of the cells on solid surfaces confer advantages over free cells in suspension, namely the immobilized cellular matter occupy less space, require smaller volume of growth medium, are easier to handle, and can be used repeatedly for product generation.3,4 Moreover, immobilization can also increase the resistance of cell cultures to harsh environmental conditions such as salinity, metal toxicity and variations in pH.3,4 Entrapment is one of the most common immobilization methods which involves capturing the cells in a three dimensional gel matrix, made of polymeric materials or inorganic spheres.4 Both synthetic polymers (e.g., acrylamide, a Centre for Strategic Nano-Fabrication, School of Chemistry and Biochemistry, The University of Western Australia, M313, 35 Stirling Highway, Crawley, WA 6009, Australia. E-mail: colin.raston@uwa.edu.au; Tel: +61 8 6488 3045; Fax: +61 8 6488 8683 b ARC Centre of Excellence in Plant Energy Biology, The University of Western Australia M313, 35 Stirling Highway, Crawley, WA 6009, Australia. E-mail: steven.smith@uwa.edu.au; Tel: +61 8 6488 4403; Fax: +61 8 6488 4401 2682 | Green Chem., 2012, 14, 2682–2685 polyurethane, polyvinyl) and natural polymers (e.g., collagen, agar, cellulose, alginate, carrageenan) have been used for this purpose.5 Several studies have been reported on wastewater treatment involving the entrapment of microalgae cultures inside alginate beads, porous glass, and several synthetic polymers.6–8 However, most attempts to immobilize viable algae cells inside such insoluble materials have limitations, with the encapsulating materials having volume/surface ratios usually orders of magnitude larger than thin films. As a consequence, algal viability is mostly reported to decrease which relates to the need for the nutrients or reactants to diffuse far into the material to reach the algal cells.5 We have developed a new technique to overcome these problems using electrospun nanofiber mats as the matrix for immobilizing the algal cells, with an overall strategy to combine wastewater treatment processing with algal harvesting in a single process. Electrospun nanofibers of chitosan were employed as a polymer/matrix support for green microalgae in the current study. Chitosan is composed of D-glucosamine and N-acetyl-Dglucosamine, and is formed by the deacetylation of chitin (β-Nacetyl-D-glucosamine polymer).9 Chitin is usually extracted from the exoskeletons of crustaceans (e.g., crab, lobster and shrimp) and even from the cell walls of fungi.9,10 Chitosan is non-toxic and biodegradable, and can be used as an animal feed.9 Furthermore, it can be used as a coagulant for wastewater treatment and for the recovery of waste sludge.11 The removal of nitrate ions is regulated by law which relates to its hazardous effects on human health and the environment. Several methods have been reported for the removal of nitrate from water bodies, including biological denitrification,12 chemical reduction,13 electrodialysis,14 and a combined bioelectrochemical/adsorption process.15 In this study, we aimed to superimpose the treatment efficiencies of microalgae in reducing nitrate and electrostatic binding of the nitrate ion by chitosan, i.e. combining biological and chemical processing. Chlorella vulgaris cultures were used as the green microalgae which were originally obtained from the Australian National Algae Culture Collection at CSIRO, Tasmania. Cell growth was carried out under artificial diurnal-illumination (16 h light/8 h dark cycle) at around 22 °C. Electrospun chitosan mats were fabricated following the procedure optimized in previous studies.16,17 This involved dissolving chitosan powder (6 wt%) in a mixture of trifluoroacetic acid (TFA) and dichloromethane (DCM) (70 : 30 v/v), with 5.4% (v/v) addition of glutaraldehyde This journal is © The Royal Society of Chemistry 2012 Published on 08 August 2012. Downloaded by University of Western Australia on 09/10/2014 10:19:02. View Article Online solution (25% in H2O) immediately prior to electrospinning.17 This was deemed necessary to cross-link the chitosan to avoid polymer breakup and dissolution. The experimental settings of electrospinning processes were as follows: (i) syringe pump speed: 0.1 mm min−1, (ii) voltage: 18 kV, (iii) distance between the target and the tip of the syringe: 11 cm, (iv) target speed: 1 m min−1, (v) traverse speed: 0.5 cm min−1. Fiber mats were annealed overnight to remove any remaining solvent and stored until required. After electrospinning, 2 mL of algae solution in its exponential phase of growth with a total chlorophyll content (Chl a and b) of ∼2 mg L−1, was placed onto a cut out rectangular chitosan-mat (3 × 2 cm) and left at room temperature for approximately 48 hours to allow sufficient attachment of algae cells to the surface of the mat. Microalgae cell walls contain various polysaccharides, which are compatible with the surface of the chitosan nanofibers.18,19 The presence of negative surface charge on the surface of Chlorella cells, arising from dissociation of uronic acid groups, and/or the presence of sulfate groups for example,18 provide electrostatic attraction to the positively charged primary amine groups of chitosan not involved in the above cross linking. Moreover, the negatively charged surface of the microalgae can also result in binding metal ions, thereby providing an opportunity for biosorption applications, along with the removal of nitrate ions as functional algal cells.18–20 This bionano-composite material was then placed into nitrate containing artificial growth medium which contained mainly phosphates, nitrates, carbonate buffer, micronutrients and vitamins,21 with an initial nitrate-nitrogen concentration of around 30 mg L−1. Nitrate-nitrogen (NO3−-N) term refers to the amount of nitrogen (N) in liquid solutions coming from nitrate ions (NO3−). This nitrate-nitrogen concentration is within the range of other algal nitrate removal studies.2,22 Furthermore, it simulates the range of nitrate content present in ground-waters (∼0.1–50 mg L−1 NO3−-N)23 and sewage treatment plant effluents. The regulatory limit for the maximum contaminant levels of [NO3−-N] in public drinking water is 10 mg L−1, as established by the United States Environmental Protection Agency (EPA).24,25 During the algal growth, the pH of the medium was kept around 6.5–7.0 by the addition of dilute hydrochloric acid (HCl) when necessary. The colorimetric “cadmium reduction method” was employed for the nitrate-nitrogen analysis, using chemicalkits in the form of powder pillows (HACH®, NitraVer Nitrate Reagent) and a colorimeter (HACH® DR/870).26 For comparative purposes, a control experiment with a chitosan nanofiber mat devoid of algae culture was also treated under the same conditions. Scanning electron microscopy (SEM) analyses were acquired using a Zeiss 1555 VP-FESEM, while the accelerating voltage was changed between 3 to 5 kV. The air-dried samples were coated with approximately 3 nm layer of platinum before imaging. A NanoMan AFM system (Veeco Instruments Inc.) was used for the atomic force microscopy (AFM) analysis, operating under the tapping mode. Chlorophyll content of the cells was analyzed using spectrophotometric measurements of methanol extracts obtained from the algal culture pellets.27 A challenge in the present work was to fabricate an insoluble, fibrous structure with sufficient porosity, which can facilitate the diffusion of nutrients and cellular products between the environment and the algae. Fig. 1a and b show scanning electron This journal is © The Royal Society of Chemistry 2012 Fig. 1 SEM images of as prepared electrospun chitosan nanofibers at low and high magnifications with scale bars of (a) 1 μm, and (b) 10 μm, respectively. Fig. 2 (a) SEM images of porous and swollen chitosan nanofiber mats, after exposure to nitrate containing media for two days, (b) SEM images of chitosan nanofibers surrounding individual, and (c) multiple algae cells. Scale bars are given as 1 μm. microscopic (SEM) images of the nanofiber structure of the chitosan mats after the electrospinning process. The diameter of the fibers was between 50 to 180 nm, with an average diameter of around 91 nm. After placing the nanofiber mat into aqueous solution, the fibers gradually swelled with a significant increase in porosity of the material, Fig. 2a, and became an effective support matrix for the C. vulgaris cells, which have the expected diameter from SEM images around 3–4 μm, Fig. 2b and c. This porous structure has an advantage for facilitating the diffusion of materials such as nutrients and waste products between the environment and the algae, while the replication of algal cells is accomplished on the surface of the nanofiber mat, Fig. 2c. The height profile measurements obtained by AFM analyses established the thickness of the chitosan mat at close to 400 nm, Fig. 3a, with the overall thickness for the algae attached chitosan mat at 4.3 μm, Fig. 3b. The difference in height is consistent with the attachment of a single layer of individual C. vulgaris cells on the nanofiber mats. Fig. 4 shows the optical images of algae cells attached to the surface of chitosan mats (a) initially, (b) after 3 days, and (c) after 10 days from the start of the growth experiments in 40 mL liquid media. Note the increasing green color on the surface is due to the increased concentration of algal cells on the chitosan mat with respect to time. Detailed imaging is given as SEM images in Fig. 5. The amount of algae cells on a chitosan mat with same dimensions yielded around 4 cells per 100 μm2 for a 3 day old sample, Fig. 5c, whereas this increased to around Green Chem., 2012, 14, 2682–2685 | 2683 Published on 08 August 2012. Downloaded by University of Western Australia on 09/10/2014 10:19:02. View Article Online Fig. 3 AFM topographic mapping of chitosan nanofiber mats (4 × 4 μm) without, (a) and with, (b) algal cells. Fig. 6 Nitrate-nitrogen (NO3−-N) concentration (mg L−1) of algal medium versus time. Chitosan nanofibers devoid of algal cells are represented with triangles, whereas those with immobilized algal cells are shown as rectangles. Fig. 4 Progress of the algal growth on the surface of chitosan mats: (a) initially, (b) after 3 days, (c) after 10 days of the growth experiment. Fig. 5 SEM images of immobilized C. vulgaris cells on the surface of chitosan nanofiber mats after different time intervals. (a and c) are for mats after 3 days, and (b and d) are for mats 10 days old. 20 cells per 100 μm2 by the 10th day of the treatment process, Fig. 5d. Fig. 6 shows the nitrate-nitrogen concentration versus time in the absence of algae (triangles) or algae attached (rectangles) nanofiber mats. After the insertion of this bionano-composite into the liquid media (Vtotal: 40 mL), around 30% of the initial nitrate value was decreased within the first 2 days. This reduction in nitrate is mainly caused through the uptake by the chitosan nanofibers rather than the algal cultures, as a physicochemical adsorption process. A similar pattern was also observed for the mats devoid of algal cells where after the second day there was no further nitrate removal. In contrast the algae containing mats continued their nitrate uptake, being used in their cellular metabolism for replication, building more biomass and energy products. 2684 | Green Chem., 2012, 14, 2682–2685 Amino groups in chitosan are protonated at acidic to neutral pH conditions,28 which enhance the adhesive properties of chitosan by increasing its tendency to attach negatively charged entities which in this case are algal cell walls and nitrates. Chlorella vulgaris cell walls are known to be highly negatively charged with a zeta potential of around −30 mV in neutral water.29 On the other hand, the zeta potential of the positively charged chitosan nanofibers is +20 mV at neutral pH. Matsumoto et al.30 reported the zeta potential values of chitosan nanofibers to be highly dependent on the pH of the media, with it increasing to +30 mV in pH around 5–6, whereas it drops to zero for pH values above 8.30 Algal growth tends to alkalify its medium, as the cellular uptake of anions (such as nitrates, phosphates, carbonates, etc.) is stabilized with equivalent amounts of hydroxyl (OH−) anion efflux.31 For this reason, we maintained the pH of the culture around 6.5–7 by the regular addition of dilute HCl during the current study. Due to the nature of HCl, several other acidifying agents (such as CO2) can be considered for any future developments and advanced scale-up processes for municipal and/or industrial wastewater samples. Clearly the presence of the nanofiber mat in the liquid environment is responsible for the initial removal of nitrate while the continued growth of algae subsequently consumes the remaining nitrate in further stages with a slower rate. Overall nitrate removal rates were calculated as 32 ± 3%, and 87 ± 4%, for the “microalgae-absent” and “microalgae-attached” chitosan mats, respectively. Several studies have already been reported on wastewater treatment with immobilized microorganisms. Fierro et al.22 investigated the effect of nitrate removal by Scenedesmus spp. cyanobacterial cells immobilized within spherical chitosan beads. They achieved 70% nitrate removal for the immobilized cultures, while 20% of the initial nitrate content was removed by the blank chitosan beads. In another study, Mallick and Rai32 also achieved relatively higher nitrate removal rates (73%) by Anabaena doliolum and Chlorella vulgaris cells immobilized in chitosan beads compared to the cells immobilized in other types of gels made of alginate, carrageenan or agar. De-Bashan et al.33 achieved only 15% nitrate removal for co-immobilized This journal is © The Royal Society of Chemistry 2012 Published on 08 August 2012. Downloaded by University of Western Australia on 09/10/2014 10:19:02. View Article Online microorganisms (Chlorella vulgaris with a growth-promoting bacterium Azospirillum brasilense) within alginate beads. At the other end of the scale, Tam and Yong34 reported complete nitrate removal using immobilized C. vulgaris cells within calcium alginate beads. The treatment efficiency of our current method is comparable with that of these aforementioned methods, although large variations among the experimental parameters; including wastewater composition, microbial species, duration of the process, type of bioreactor, chemical composition and shape of the immobilization matrix, make direct comparison difficult. In summary, we have established the use of cross-linked chitosan nanofiber mat as a water-insoluble and non-toxic support for algal growth and nitrate removal from waters. Algal growth on a support material can lead to combine algal harvesting, dewatering, and processing steps in a single stage. This bionano-composite material is potentially an attractive, simple and highly durable polymer, with the mats still retaining their integrity after six months in contact with an aqueous solution, and has promise for industrial and/or municipal wastewater treatment processes. Acknowledgements This work has been supported by The University of Western Australia and the Australian Research Council. We would like to acknowledge the facilities of the Australian Microscopy & Microanalysis Research Facility at the Centre for Microscopy, Characterization & Analysis, The University of Western Australia, which was funded by the University, State and Commonwealth Governments. AFM images were performed at Curtin University. Notes and references 1 Metcalf and Eddy, Inc, Wastewater Engineering: Treatment and Reuse, McGraw-Hill, New York, 4th edn, 2003. 2 N. Mallick, BioMetals, 2002, 15, 377. 3 L. Hall-Stoodley, J. W. Costerton and P. Stoodley, Nat. Rev. Microbiol., 2004, 2(2), 95. 4 Y. Liu, M. H. Rafailovich, R. Malal, D. Cohn and D. Chidambaram, Proc. Natl. Acad. Sci. U. S. A., 2009, 106(34), 14201. 5 M. S. A. Hameed and O. H. Ebrahim, Int. J. Agri. Biol., 2007, 9, 183. This journal is © The Royal Society of Chemistry 2012 6 K. Abe, E. Takahashi and M. Hirano, J. Appl. Phycol., 2008, 20, 283. 7 G. Schumacher and I. Sekoulov, Water Sci. Technol., 2002, 46, 83. 8 E. Delahaye, R. Boussahel, T. Petitgand, J. P. Duguet and A. Montiel, Desalination, 2005, 177, 273. 9 A. Lavoie and J. de La Noue, J. World Maricul. Soc., 1983, 14, 685. 10 A. Zamani, L. Edebo, B. Sjöström and M. J. Taherzadeh, Biomacromolecules, 2007, 8, 3786. 11 W. A. Bough, Process Biochem., 1976, 11, 13. 12 E. Wasik, J. Bohdziewcz and M. Blaszczyk, Process Biochem., 2001, 37, 57. 13 H.-Y. Hu, N. Goto and K. Fujie, Water Res., 2001, 35, 2789. 14 A. Elmidaoui, F. Elhannouni, S. M. A. Menkouchi Sahli, L. Chay, E. Elabbassi, M. Hafsi and D. Largeteau, Desalination, 2001, 136, 325. 15 Z. Feleke and Y. Sakakibara, Water Sci. Technol., 2001, 43, 25. 16 K. Ohkawa, D. I. Cha, H. Kim, A. Nishida and H. Yamamoto, Macromol. Rapid Commun., 2004, 25, 1600. 17 M. Bradshaw, J. Zou, L. Byrne, K. S. Iyer, S. G. Stewart and C. L. Raston, Chem. Commun., 2011, 47, 12292. 18 D. Kaplan, D. Christiaen and S. M. Arad, Appl. Environ. Microbiol., 1987, 53, 2953. 19 R. H. Crist, K. Oberholser, N. Shank and M. Nguyen, Environ. Sci. Technol., 1981, 15, 1212. 20 B. Volesky and Z. R. Holan, Biotechnol. Prog., 1995, 11, 235. 21 C. J. S. Bolch and S. I. Blackburn, J. Appl. Phycol., 1996, 8, 5. 22 S. Fierro, M. del P. Sanchez-Saavedra and C. Copalcua, Bioresour. Technol., 2008, 99, 1274. 23 U. N. Dwivedi, S. Mishra, P. Singh and R. D. Tripathi, in Environmental Bioremediation Technologies, ed. S. N. Singh and R. D. Tripathi, Springer, New York, 2007, ch. 16, pp. 353–389. 24 EPA, National Pesticide Survey: Project Summary, U.S. Environmental Protection Agency, Washington DC, 1990. 25 S. Ghafari, M. Hasan and M. K. Aroua, Bioresour. Technol., 2008, 99, 3965. 26 APHA, Standard Methods for the Examination of Water and Wastewater, American Public Health Association, Washington, DC, 18th edn, 1992. 27 H. K. Lichtenthaler and C. Buschmann, in Current Protocols in Food Analytical Chemistry, ed. R. E. Wrolstad, John Wiley & Sons Inc., New York, 2001, pp. F4.3.1–F4.3.8. 28 L.-Q. Wu, P. Gadre Anand, H. Yi, M. J. Kastantin, W. Rubloff Gary, E. Bentley William, F. Gregory Payne and R. Ghodssi, Langmuir, 2002, 18, 8620. 29 B.-M. Hsu, Parasitol. Res., 2006, 99, 357. 30 H. Matsumoto, H. Yako, M. Minagawa and A. Tanioka, J. Colloid Interface Sci., 2007, 310, 678. 31 J. Naus and A. Melis, Plant Cell Physiol., 1991, 32, 569. 32 N. Mallick and L. C. Rai, World J. Microbiol. Biotechnol., 1994, 10, 439. 33 L. E. de-Bashan, Y. Bashan, M. Moreno, V. K. Lebsky and J. J. Bustillos, Can. J. Microbiol., 2002, 48, 514. 34 N. F. Y. Tam and Y. S. Wong, Environ. Pollut., 2000, 107, 145. Green Chem., 2012, 14, 2682–2685 | 2685 ) ARTICLE Hierarchical Patterning of Multifunctional Conducting Polymer Nanoparticles as a Bionic Platform for Topographic Contact Guidance Dominic Ho,†,‡ Jianli Zou,§ Xianjue Chen,†,0 Alaa Munshi,† Nicole M. Smith,†, Vipul Agarwal,† Stuart I. Hodgetts,‡ Giles W. Plant,^ Anthony J. Bakker,‡ Alan R. Harvey,‡ Igor Luzinov,# and K. Swaminathan Iyer*,† School of Chemistry and Biochemistry, The University of Western Australia, Crawley, Western Australia 6009, Australia, ‡School of Anatomy, Physiology and Human Biology, The University of Western Australia, Crawley, Western Australia 6009, Australia, §Institute for Integrated Cell-Material Sciences (iCeMS), iCeMS Complex 2, Kyoto University, Yoshida-Honmachi, Sakyo-ku, Kyoto, 606-8501, Japan, Experimental and Regenerative Neurosciences, School of Animal Biology, The University of Western Australia, Crawley, Western Australia 6009, Australia, ^Stanford Partnership for Spinal Cord Injury and Repair, Department of Neurosurgery, Stanford University School of Medicine, Stanford, California 94305, United States, and #School of Materials Science and Engineering, Clemson University, Clemson, South Carolina 29634, United States. 0Present address: Centre for NanoScale Science and Technology, School of Chemical and Physical Sciences, Flinders University, Bedford Park, Adelaide, SA 5042, Australia. ) † ABSTRACT The use of programmed electrical signals to influence biological events has been a widely accepted clinical methodology for neurostimulation. An optimal biocompatible platform for neural activation efficiently transfers electrical signals across the electrode cell interface and also incorporates large-area neural guidance conduits. Inherently conducting polymers (ICPs) have emerged as frontrunners as soft biocompatible alternatives to traditionally used metal electrodes, which are highly invasive and elicit tissue damage over long-term implantation. However, fabrication techniques for the ICPs suffer a major bottleneck, which limits their usability and medical translation. Herein, we report that these limitations can be overcome using colloidal chemistry to fabricate multimodal conducting polymer nanoparticles. Furthermore, we demonstrate that these polymer nanoparticles can be precisely assembled into large-area linear conduits using surface chemistry. Finally, we validate that this platform can act as guidance conduits for neurostimulation, whereby the presence of electrical current induces remarkable dendritic axonal sprouting of cells. KEYWORDS: multimodal nanoparticles . conducting polymers . capillary force lithography . neurostimulation E xogenous electrical stimulation has been effectively used both in clinical practice and in laboratory research to regulate cell-type-dependent adhesion, differentiation, and growth.1 This phenomenon of introducing programmed electrical signals locally to influence biological events has resulted in the pursuit of sophisticated medical bionic devices.2 An important property that dictates the performance of most bionic electrodes is the electrode/cellular interface and its ability to transmit charge across the biointerface.3 Traditionally metallic electrodes made of platinum, gold, iridium oxide, tungsten, their alloys, and more recently carbon fibers have been effectively HO ET AL. employed in bionic devices.4 They have been employed for deep brain stimulation, as cochlear implants, for vagus nerve stimulation to treat epilepsy, and for stimulating regeneration in the central nervous system.2 However, stiff metal electrodes suffer a major drawback of eliciting tissue damage over long-term implantation.2 Importantly, it is now recognized that nanoscale patterns provide topographic guidance cues for cells. This has been widely exploited to engineer sophisticated regenerative platforms for nerves, muscles, skin, and bones.5 The need to incorporate large-area nanoscale patterns for bionic applications coupled with the demand toward miniaturization of VOL. 9 ’ NO. 2 ’ * Address correspondence to swaminatha.iyer@uwa.edu.au. Received for review November 20, 2014 and accepted January 26, 2015. Published online January 26, 2015 10.1021/nn506607x C 2015 American Chemical Society 1767–1774 ’ 2015 1767 www.acsnano.org HO ET AL. sustained release from the nanoparticles once patterned and multimodal imaging of the nanoparticle constructs once implanted. RESULTS AND DISCUSSION Patterned Multifunctional PEDOT:PSS Nanoparticle Arrays. In this study poly(glycidal methacrylate) (PGMA) is used as a reactive macromolecular anchoring platform both on the substrate as a nanoscale layer and as a colloidal nanoparticle to enable multilayer assembly (Figure 1). A polymer with epoxy functionality was chosen, since the reactions of epoxy groups are universal and easily transferable to various substrates, affording ease of attachment of functional molecules. Furthermore, the epoxy groups of the polymer can cross-link to provide structural integrity to the pattern and nanoparticle constructs.11 The mobility of the reactive loops of PGMA ensures greater access to anchoring, resulting in a 23-fold greater grafting density when compared to a monolayer of epoxy groups on a nanoparticle surface of similar dimension, enabling high loading using a layer by layer approach that is adopted in the current study.11 Polymer nanospheres were initially prepared using an oil in water emulsion methodology from PGMA modified with a rhodamine-B (RhB) dye, encapsulated with magnetite (Fe3O4) nanoparticles to form the core platform (Figure 1a,b). Not only does the incorporation of magnetite and RhB render these constructs multimodal for both MRI and fluorescence imaging, but importantly in the present case magnetite provides a means to separate, wash, and purify the nanoparticles using a magnetic fractionation column during each step of layered assembly. Polyethylenimine (PEI) was then covalently bound to the RhB-PGMA core to facilitate a cationic layer for electrostatic conjugation of an anionic conducting polymer, PEDOT:PSS (Figure 1c,d). Capillary force lithography (CFL) was then used to generate large-area nanoscale conduits in which PEDOT:PSS nanoparticles are electrostatically directed to self-assemble as linear channels from solution (Figure e,f). Capillarity allows the polymer melt to fill up the void space between the polymer and the applied mold when the temperature is above the glass-transition temperature (Tg), thereby generating a large-area pattern that depends on the size of stamp. Importantly, the technique needs no specialized instrumentation for generation of large-area patterns. Patterns can easily be generated using polydimethylsiloxane (PDMS) stamps, which in turn can be fabricated using the ubiquitous optical storage discs as a master. An optical data storage disc is typically made of a polymer (polycarbonate) disc, on which a single spiral track is drilled. The typical width and depth of each line in the spiral track are 800 and 130 nm, respectively, and the periodicity of the track is ∼1.5 μm (Figure S1). In the present study, an indium tin oxide (ITO) substrate was VOL. 9 ’ NO. 2 ’ 1767–1774 ’ ARTICLE biocompatible implantable devices has resulted in the emergence of inherently conducting polymers as frontrunners for fabricating flexible organic electrode materials. However, advances in the applicability of patterned surfaces of inherently conducting polymers in bionic devices have been limited due to the difficulties of transferring printing techniques and their integration under physiological conditions. In this article, we report a transferable method to fabricate multifunctional poly(3,4-ethylenedioxythiophene)poly(styrenesulfonate) (PEDOT:PSS) nanoparticles and direct their self-assembly by electrostatic interactions into large-area patterns. Using the rat pheochromocytoma cell line (PC12), we demonstrate the suitability of the assembly as a bionic platform for exogenous electrical stimulation. The three primary classes of conducting polymers that have been studied are polyanilines, polypyrroles, and polythiophenes.6 The ease of functionalization of polythiophenes and maintenance of conductivity under physiological conditions has made them primary candidates for multifunctional organic bionic devices.7 The most widely explored processes for the fabrication of organic conducting polymer patterns are electropolymerization, extrusion printing, inkjet printing, microcontact printing, electrospinning, and more recently high-precision Dip Pen Nanolithography (DPN).4,6 Electropolymerization has been widely used for coating metal/carbon substrates, following which patterning is achieved by top-down lithography on polymer thin films covering larger area electrodes. This technique results in controlled, high-resolution nanoscale patterns but is limited by the ability to regulate polymerization of monomers on nanoscale implantable electrodes.8 Similarly, printing techniques have achieved significant advances in recent years, reaching high-throughput patterns, but are limited in resolution by the liquid dispensing techniques, which operate within the limit of tens of micrometers.9 Electrospinning techniques have offered simple processable solutions to generate 3D scaffolds at resolutions mimicking the extracellular matrix architecture but are limited by the inability to generate patterned conducting conduits for the development of bionic guidance channels.4 The aforementioned shortfalls have been recently overcome by the advances of DPN, which enables precise deposition, patterning down to nanoscale resolution, and most importantly applicability over a wide range of substrates.10 However, advances are limited by their cost, need for specialized equipment, and low throughput. In the present paper we adopt a bottom-up self-assembly process to precisely pattern conducting polymer nanoparticles into patterns as conduits for guidance. The approach is easily adoptable over multiple substrates, needs no specialized equipment, and affords large-area patterns. Importantly, this approach enables drug encapsulation and 1768 2015 www.acsnano.org ARTICLE Figure 1. Schematic illustration of the fabrication protocol to pattern multifunctional PEDOT:PSS nanoparticle arrays for exogenous electrical stimulation. (ad) Multilayer assembly of conducting PEDOT:PSS nanoparticle fabrication via nonspontaneous emulsification. (a) An organic phase is initially formed by dissolving RhB-modified PGMA (yellow) and Fe3O4 (purple) in a 1:3 mixture of CHCl3 and MEK. (b) Colloidal fluorescent PGMA-Fe3O4 nanoparticles are fabricated upon dropwise addition of the organic phase to an aqueous solution of Pluronic F-108. (c) Cationic second layer via covalent attachment of PEI (green) to the PGMA-Fe3O4 core. (d) Anionic conducting polymer layer via electrostatic attachment of PEDOT:PSS (blue). (eg) Patterning of the multilayered PEDOT:PSS nanoparticles for exogenous electrical stimulation of PC12 cells. (e) Linear nanoparticle conduits patterned on a substrate via capillary force lithography (CFL) using charge complementarity. A detailed schematic of the CFL procedure can be found in Figure S2. (f) PC12 cells (green) were cultured onto the biocompatible platform, followed by (g) exogenous electrical modulation. modified first by spin coating a thin film of PGMA followed by a second spin-coated layer of polystyrene (PS) using previously reported conditions.12 The PS layer acts as a chemical resist to selectively react the epoxy groups of PGMA following patterning. The PS/PGMA bilayer was annealed with the PDMS mask at 130 C (T > Tg of PS) to induce patterning via capillary flow. The reusable PDMS stamp was peeled off following heat treatment to obtain a patterned surface resulting in alternating PGMA and PS stripes. HO ET AL. Ethylenediamine (EA) was then grafted to PGMA to result in cationic linear patterns. PEDOT:PSS nanoparticles were then electrostatically assembled onto the patterned surface, followed by washing steps to remove PS to obtain linear arrays of assembled PEDOT: PSS nanoparticles. A detailed schematic of the fabrication process is shown in Figure S2. The nanoparticle and the patterns were characterized at each step of the assembly (Figure 2). The PEDOT:PSS nanoparticles were an average size of 200 nm (Z-average) with a VOL. 9 ’ NO. 2 ’ 1767–1774 ’ 1769 2015 www.acsnano.org ARTICLE Figure 2. Characterization of the multilayered PEDOT:PSS conducting nanoparticles and their assembly as linear conduits. (a) TEM micrograph of the multilayered PEDOT:PSS nanoparticles. Scale bar = 200 nm. (Inset: high-magnification TEM image of PEDOT:PSS-coated nanoparticles showing encapsulated Fe3O4 nanoparticles. Scale bar = 10 nm.) (b) DLS particle size distributions of the PEDOT:PSS nanoparticles in solution. (c) Zeta potential distributions of the nanoparticles: PGMA-Fe3O4 core (black) with an average zeta potential of 3.9 ( 1.3 mV, cationic PEI-coated (red) with an average zeta potential of 37 ( 1.2 mV, and anionic PEDOT:PSS-coated (blue) with an average zeta potential of 29 ( 6.15 mV. (d) Current vs voltage response for the nonconducting PEI-coated nanoparticles (red) and conducting PEDOT:PSS-coated (black) nanoparticles. (eg) Tapping mode AFM topography images of the nanoparticle patterns at each stage of fabrication: PGMA and PS stripes (e), EA-modified PGMA and PS stripes (f), PEDOT:PSS nanoparticle patterns (g). (hj) Corresponding height profiles of the nanoparticle patterns at each stage of fabrication.: PGMA and PS stripes (h), EA-grafted PGMA and PS stripes (i), PEDOT:PSS nanoparticles patterns (j). The AFM line scans corresponding to the height profiles are indicated on the topography images in (e)(g). (k, l) SEM micrographs of the nanoparticle patterns at a magnification of 25k (k) and 11k (l) indicating the formation of tightly packed and highly ordered nanoparticle arrays. (m) Confocal fluorescence image of the RhBfunctionalized PEDOT:PSS nanoparticle arrays at 20 magnification. HO ET AL. VOL. 9 ’ NO. 2 ’ 1767–1774 ’ 1770 2015 www.acsnano.org ARTICLE Figure 3. Biocompatibility of the PEDOT:PSS nanoparticle arrays with PC12 cells. (a) Cell viability determined using MTS calorimetric assay obtained at 72 h after an initial exogenous electrical stimulation for 2 h and in the absence of stimulation showing no significant changes. (b) SEM micrograph demonstrating preferential cell adhesion to the pattern area (yellow box). Image acquired at 323 magnification 72 h after the addition of NGF without exogenous electrical stimulation. (c) Highmagnification (12k magnification) SEM images demonstrating specific and preferential interactions of neurites (white arrows) with the PEDOT:PSS linear conduits (red arrows). polydispersity index (PDI) of 0.07, a zeta potential of 29 ( 6.15 mV, and a conductivity of 2.5 1012 S/cm. The measured conductivity is in accordance with other values reported in the literature for polymer blends.13 Importantly, this low conductivity is important under physiological conditions to induce local cellular stimulation and avoid tissue damage due to toxic overstimulation.14 The final self-assembled linear arrays of PEDOT:PSS nanoparticles were of large-area highdensity packing, as confirmed at various length scales using AFM, SEM, and fluorescence imaging. Biocompatibility Assessment of the PEDOT:PSS Nanoparticle Arrays. Topographic modulation of tissue response is one of the most important considerations in developing bionic implants. Topographic contact guidance using micropatterns has been widely exploited to influence cell migration, adhesion, and proliferation.15,16 One of the pivotal first steps in the present study was to establish biocompatibility of the patterned structures. PC12 cells were chosen in the present case, as they have been demonstrated to show HO ET AL. enhanced neurite outgrowth and spreading upon exogenous stimulation on a conducting polymer substrate.17 MTS (3-(4,5-dimethylthiazol-2-yl)-5-(3carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium, inner salt) assays, cell viability assays, and SEM imaging were performed after exogenous electrical stimulation and in the absence of electrical stimulation to determine effects on cell viability and cell adhesion (Figure 3a,b and Figure S3). The stimulation conditions used in the present study involved a monophasic pulsed current at a frequency of 250 Hz with a 2 ms pulse width and an amplitude of 1 mA for 2 h, similar to protocols previously reported for similar cell lines.18,19 Importantly, we observed no changes in cell viability upon exogenous stimulation and observed preferential adhesion of the PC12 cells to the patterned surface over a nonpatterned surface in both cases (( stimulation). High-magnification SEM imaging (no stimulation) further revealed preferential interaction of the PC12 cells to the PEDOT:PSS nanoparticle arrays, confirming not only biocompatibility with the large VOL. 9 ’ NO. 2 ’ 1767–1774 ’ 1771 2015 www.acsnano.org ARTICLE Figure 4. Exogenous electrical stimulation induced dendritic sprouting of the PC12 cells guided by the PEDOT:PSS linear conduits. (a) Significant increase in the average cell area is observed 72 h after exogenous electrical stimulation on the nanoparticle platform in comparison to unstimulated and nonpatterned controls. (b) Corresponding decrease in PC12 cell proliferation observed 72 h after exogenous electrical stimulation on the nanoparticle platform in comparison to unstimulated and nonpatterned controls. (cf) Representative confocal images (40 magnification) of β-III tubulin immunohistochemically stained cells 72 h after the following treatments: {(þ) pattern, () stimulation} (c); {(þ) pattern, (þ) stimulation} (d); {() pattern, () stimulation} (e); {() pattern, (þ) stimulation} (f), demonstrating modulation of cell morphology. (g) High-magnification SEM image (magnification 8k) indicating the formation of extensive dendritic networks (white arrows) guided by the PEDOT:PSS arrays. Inset: The corresponding low-magnification image of the area (yellow box) analyzed (magnification 3k). area of the pattern but also potential applicability of the nanoscale linear arrays as guidance conduits (Figure 3c). Exogenous Electrical Stimulation Induced Dendritic Sprouting of the PC12 Cells. Electrical stimulation has been effectively used to modulate growth and differentiation of HO ET AL. anchorage-dependent cells such as neurons, fibroblasts, and epithelium cells.17,20,21 In the central nervous system, brief stimulation to the proximal end of transected peripheral nerves has been shown to augment preferential motor reinnervation,22 improve the specificity of sensory reinnervation,23 and accelerate VOL. 9 ’ NO. 2 ’ 1767–1774 ’ 1772 2015 www.acsnano.org METHODS SUMMARY Nanoparticle Synthesis. The conducting nanoparticles were prepared via a nonspontaneous emulsification route. Briefly, rhodamine B was attached to PGMA in MEK at 80 C under N2 for 5 h. The modified PGMA was then precipitated in diethyl ether and dried under N2. This was dispersed in a 1:3 mixture of CHCl3 and MEK along with 25 mg of Fe3O4 to form the organic phase. This organic phase was added dropwise into a rapidly stirring aqueous solution of Pluronic F-108. The emulsion was homogenized with a probe-type ultrasonic wand for 1 min. The organic solvents were then evaporated off under N2. Large aggregates of Fe3O4 and excess polymer were separated via centrifugation. The nanoparticles in the supernatant were then mixed with PEI and heated to 80 C for 16 h to facilitate attachment. The PEI-coated nanoparticles were isolated and washed on a magnetic separation column. Next, a diluted solution of PEDOT:PSS was added dropwise under rapid stirring to nanoparticles at a concentration of 0.5 mg/mL to facilitate electrostatic attachment. This was followed by sonication for 10 min and stirring for 18 h. The nanoparticles were then washed multiple times in water before being stored at 4 C for further use. Platform Fabrication. To direct the self-assembly of the nanoparticles, a template was fabricated by CFL. A 0.2% w/v PGMA in CHCl3 solution was spin coated on ITO coverslips and annealed at 120 C for 20 min. Next, 1.3% w/v PS in toluene was spin coated onto the PGMA surface. A PDMS stamp was then placed onto the PS layer, followed by heat treatment in an oven at 130 C for 1 h. Once cooled, the stamp was peeled off. This was followed by exposure to EA at room temperature for 5 h. The pattern was next washed multiple times with water to remove unreacted EA. A 50 μL amount of 4 mg/mL nanoparticle suspension was drop casted onto the patterned area of the coverslip. The setup was then placed in a sealed vial, facilitating HO ET AL. extension during NGF-induced PC12 cell differentiation.29,30 Using immunohistochemical staining for β-III tubulin it was determined that stimulation on the patterned surface resulted in a significant increase in the cell area and lower number of cells per unit area, indicating exogenous electrical stimulation induced differentiation of PC12 cells (Figure 4af, Figure S4). High-magnification SEM (Figure 4g) also revealed that stimulation resulted in an extensive dendritic network guided by the linear conduits of PEDOT:PSS nanoparticles. ARTICLE the reinnervation of distal target tissues.24 These have been reported to depend on depolarization of the neuronal soma and its axon, involvement of axon guidance factors such as polysilylated neural cell adhesion molecule,25 the L2/HNK-1 carbohydrate,26 and brain-derived neurotrophic factor.27 Finally electrical stimulation induced neurite outgrowth was recently reported to be dependent on calcium influx through L- and N-type voltage-dependent calcium channels and calcium mobilization from IP3R and RYR-sensitive calcium stores.28 In the present case, we analyzed the morphological modulation of PC12 cells following electrical stimulation having determined no change in cell viability using the MTS assay. Nerve growth factor (NGF) induces PC12 cells to change their phenotype and acquire a number of properties that are similar to sympathetic neurons. Importantly, although they can acquire properties similar to sympathetic neurons upon NGF treatment, they do not develop definitive dendritic axons or form true synapses with each other in the absence of exogenous stimulation.29 This change in phenotype upon NGF treatment is associated with a retardation in proliferation, the extension of neurites making them electrically excitable. Monitoring the cell numbers and cell area can assess this change from the proliferation state to a differentiation state. Furthermore, microtubule levels correlate precisely with the neurite CONCLUSION In summary, we have demonstrated a practical and transferable protocol to fabricate self-assembled largearea patterns of conducting polymers from solution. This overcomes some of the shortfalls in the current fabrication techniques in developing patterned organic bionic devices. The patterns generated have demonstrated excellent biocompatibility. At the same time, they have been shown to induce exogenous electrical stimulation under physiological conditions to elicit a measurable and consistent cellular response. Importantly the methodology permits the design of bionic devices capable of inducing local electrical stimulation for in vivo applications while integrating multimodal imaging and simultaneous drug delivery capabilities of nanoparticles. controlled evaporation, which allowed for electrostatic nanoparticle attachment onto the EA surface. The PS mask was then removed by washing with toluene. The resulting patterned PEDOT:PSS nanoparticle array was then used for further experimentation. Electrical Stimulation Protocol. For electrical stimulation experiments, two silver epoxy electrodes were painted onto the ends of the patterned nanoparticle arrays. Prior to cell culture, the whole platform was UV and ethanol sterilized. Wells were coated with poly(L-lysine) and 15 μg/mL of laminin followed by cell seeding at a density of 50 000 cells/well. Cells were left to adhere for 18 h. Immediately prior to stimulation, the proliferation media was replaced with low-serum nerve growth factor containing differentiation media. For stimulation, the cells were subjected to a monophasic pulsed current at a frequency of 250 Hz with a 2 ms pulse width and an amplitude of 1 mA for 2 h, after which they were left for an additional 72 h before analysis. Cell Viability Assessment. Cell viability was measured using the MTS assay as per the manufacturer protocols (Invitrogen, UK). For measurements, 80 μL from each well was transferred into a new 96-well plate and read under a plate reader at 490 nm excitation wavelength. To analyze cell morphology, cells were immunohistochemically stained for β-III tubulin. Material Characterization. AFM was performed on a Dimension 3100 AFM system with a Nanoscope IV controller used to obtain the AFM images in tapping mode, using Pt/Ir-coated contact mode probes with a spring constant of 0.2 N/m (type SCM-PIC, Bruker). TEM was performed on a JEOL 2100 transmission electron microscope at an accelerating voltage of 80 kV. SEM was performed on a Zeiss 1555 VP-FESEM, and all samples were coated with 5 nm of Pt. Biological samples were initially fixed in 2.5% glutaraldehyde and dehydrated in increasing concentrations of ethanol followed by critical point drying prior to Pt coating. Immunohistochemically stained samples were VOL. 9 ’ NO. 2 ’ 1767–1774 ’ 1773 2015 www.acsnano.org Conflict of Interest: The authors declare no competing financial interest. Supporting Information Available: Detailed materials and methods: synthesis, characterization (TEM, SEM, AFM), cell culture, and electrical stimulation experiments. This material is available free of charge via the Internet at http://pubs.acs.org. Acknowledgment. D.H., I.L., and K.S.I. designed the experiments, developed the concept, and analyzed the data. D.H., J.Z., and N.M.S. optimized the capillary force lithography experiments. D.H., X.C., V.A., and A.M. performed image acquisition using confocal microscopy, transmission electron microscopy, scanning electron microscopy, and atomic force microscopy. D.H., A.R.H., G.W.P., S.I.H., and A.B. optimized and designed the electrical stimulation experiments. This work was funded by the Australian Research Council (ARC), the National Health & Medical Research Council (NHMRC) of Australia, and the National Science Foundation (CBET-0756457). The authors acknowledge the Australian Microscopy & Microanalysis Research Facility at the Centre for Microscopy, Characterization & Analysis, and The University of Western Australia, funded by the University, State and Commonwealth Governments. The authors also wish to thank Margaret Pollett and Chrisna LeVaillant for their invaluable contribution in assisting with the PC12 cell cultures and immunohistochemistry, and Ella Marushchenko (www.scientificillustrations.com) for assistance with Figure 1. REFERENCES AND NOTES 1. Ciofani, G.; Danti, S.; D'Alessandro, D.; Ricotti, L.; Moscato, S.; Bertoni, G.; Falqui, A.; Berrettini, S.; Petrini, M.; Mattoli, V.; et al. Enhancement of Neurite Outgrowth in Neuronal-Like Cells Following Boron Nitride Nanotube-Mediated Stimulation. ACS Nano 2010, 4, 6267–6277. 2. Moulton, S. E.; Higgins, M. J.; Kapsa, R. M. I.; Wallace, G. G. Organic Bionics: A New Dimension in Neural Communications. Adv. Funct. Mater. 2012, 22, 2003–2014. 3. Wallace, G. G.; Moulton, S. E.; Clark, G. M. Electrode-Cellular Interface. Science 2009, 324, 185–186. 4. Wallace, G. G.; Higgins, M. J.; Moulton, S. E.; Wang, C. Nanobionics: The Impact of Nanotechnology on Implantable Medical Bionic Devices. Nanoscale 2012, 4, 4327– 4347. 5. Khademhosseini, A.; Langer, R.; Borenstein, J.; Vacanti, J. P. Microscale Technologies for Tissue Engineering and Biology. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 2480–2487. 6. Guimard, N. K.; Gomez, N.; Schmidt, C. E. Conducting Polymers in Biomedical Engineering. Prog. Polym. Sci. 2007, 32, 876–921. 7. Kim, D. H.; Richardson-Burns, S. M.; Hendricks, J. L.; Sequera, C.; Martin, D. C. Effect of Immobilized Nerve Growth Factor on Conductive Polymers: Electrical Properties and Cellular Response. Adv. Funct. Mater. 2007, 17, 79–86. 8. Lee, J. I.; Cho, S. H.; Park, S.-M.; Kim, J. K.; Kim, J. K.; Yu, J.-W.; Kim, Y. C.; Russell, T. P. Highly Aligned Ultrahigh Density Arrays of Conducting Polymer Nanorods Using Block Copolymer Templates. Nano Lett. 2008, 8, 2315–2320. 9. Wang, J.; Zheng, Z.; Li, H.; Huck, W.; Sirringhaus, H. Dewetting of Conducting Polymer Inkjet Droplets on Patterned Surfaces. Nat. Mater. 2004, 3, 171–176. 10. Nakashima, H.; Higgins, M. J.; O'Connell, C.; Torimitsu, K.; Wallace, G. G. Liquid Deposition Patterning of Conducting Polymer Ink onto Hard and Soft Flexible Substrates via Dip-Pen Nanolithography. Langmuir 2011, 28, 804–811. 11. Iyer, K. S.; Zdyrko, B.; Malz, H.; Pionteck, J.; Luzinov, I. Polystyrene Layers Grafted to Macromolecular Anchoring Layer. Macromolecules 2003, 36, 6519–6526. 12. Zou, J.; Zdyrko, B.; Luzinov, I.; Raston, C. L.; Swaminathan Iyer, K. Regiospecific Linear Assembly of Pd Nanocubes for Hydrogen Gas Sensing. Chem. Commun. 2012, 48, 1033–1035. HO ET AL. 13. Choi, J.; Lee, J.; Jung, D.; Shim, S. E. Electrospun PEDOT: PSS/PVP Nanofibers as the Chemiresistor in Chemical Vapour Sensing. Synth. Met. 2010, 160, 1415–1421. 14. Merrill, D. R.; Bikson, M.; Jefferys, J. G. R. Electrical Stimulation of Excitable Tissue: Design of Efficacious and Safe Protocols. J. Neurosci. Methods 2005, 141, 171–198. 15. Chen, C. S.; Mrksich, M.; Huang, S.; Whitesides, G. M.; Ingber, D. E. Micropatterned Surfaces for Control of Cell Shape, Position, and Function. Biotechnol. Prog. 1998, 14, 356–363. 16. Ito, Y. Surface Micropatterning to Regulate Cell Functions. Biomaterials 1999, 20, 2333–2342. 17. Schmidt, C. E.; Shastri, V. R.; Vacanti, J. P.; Langer, R. Stimulation of Neurite Outgrowth Using an Electrically Conducting Polymer. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 8948–8953. 18. Richardson, R. T.; Thompson, B.; Moulton, S.; Newbold, C.; Lum, M. G.; Cameron, A.; Wallace, G.; Kapsa, R.; Clark, G.; O'Leary, S. The Effect of Polypyrrole with Incorporated Neurotrophin-3 on the Promotion of Neurite Outgrowth from Auditory Neurons. Biomaterials 2007, 28, 513–523. 19. Weng, B.; Liu, X.; Shepherd, R.; Wallace, G. G. Inkjet Printed Polypyrrole/Collagen Scaffold: A Combination of Spatial Control and Electrical Stimulation of PC12 Cells. Synth. Met. 2012, 162, 1375–1380. 20. Wong, J. Y.; Langer, R.; Ingber, D. E. Electrically Conducting Polymers Can Noninvasively Control the Shape and Growth of Mammalian Cells. Proc. Natl. Acad. Sci. U.S.A. 1994, 91, 3201–3204. 21. Bourguignon, G.; Bourguignon, L. Electric Stimulation of Protein and DNA Synthesis in Human Fibroblasts. FASEB J. 1987, 1, 398–402. 22. Al-Majed, A. A.; Neumann, C. M.; Brushart, T. M.; Gordon, T. Brief Electrical Stimulation Promotes the Speed and Accuracy of Motor Axonal Regeneration. J. Neurosci. 2000, 20, 2602–2608. 23. Brushart, T. M.; Jari, R.; Verge, V.; Rohde, C.; Gordon, T. Electrical Stimulation Restores the Specificity of Sensory Axon Regeneration. Exp. Neurol. 2005, 194, 221–229. 24. Brushart, T. M.; Hoffman, P. N.; Royall, R. M.; Murinson, B. B.; Witzel, C.; Gordon, T. Electrical Stimulation Promotes Motoneuron Regeneration without Increasing Its Speed or Conditioning the Neuron. J. Neurosci. 2002, 22, 6631– 6638. 25. Franz, C. K.; Rutishauser, U.; Rafuse, V. F. Intrinsic Neuronal Properties Control Selective Targeting of Regenerating Motoneurons. Brain 2008, 131, 1492–1505. 26. Eberhardt, K. A.; Irintchev, A.; Al-Majed, A. A.; Simova, O.; Brushart, T. M.; Gordon, T.; Schachner, M. BDNF/TrkB Signaling Regulates HNK-1 Carbohydrate Expression in Regenerating Motor Nerves and Promotes Functional Recovery after Peripheral Nerve Repair. Exp. Neurol. 2006, 198, 500–510. 27. Geremia, N. M.; Gordon, T.; Brushart, T. M.; Al-Majed, A. A.; Verge, V. M. Electrical Stimulation Promotes Sensory Neuron Regeneration and Growth-Associated Gene Expression. Exp. Neurol. 2007, 205, 347–359. 28. Yan, X.; Liu, J.; Huang, J.; Huang, M.; He, F.; Ye, Z.; Xiao, W.; Hu, X.; Luo, Z. Electrical Stimulation Induces CalciumDependent Neurite Outgrowth and Immediate Early Genes Expressions of Dorsal Root Ganglion Neurons. Neurochem. Res. 2014, 39, 129–141. 29. Das, K. P.; Freudenrich, T. M.; Mundy, W. R. Assessment of PC12 Cell Differentiation and Neurite Growth: A Comparison of Morphological and Neurochemical Measures. Neurotoxicol. Teratol. 2004, 26, 397–406. 30. Drubin, D. G.; Feinstein, S. C.; Shooter, E. M.; Kirschner, M. W. Nerve Growth Factor-Induced Neurite Outgrowth in PC12 Cells Involves the Coordinate Induction of Microtubule Assembly and Assembly-Promoting Factors. J. Cell Biol. 1985, 101, 1799–1807. VOL. 9 ’ NO. 2 ’ 1767–1774 ’ ARTICLE analyzed using a Leica TCS SP2 AOBS multiphoton confocal microscope. 1774 2015 www.acsnano.org Hierarchical Patterning of Multifunctional Conducting Polymer Nanoparticles as a Bionic Platform for Topographic Contact Guidance Dominic Ho,1,2 Jianli Zou,3 Xianjue Chen,1,‡, Alaa Munshi,1 Nicole M. Smith,1,4 Vipul Agarwal,1 Stuart I. Hodgetts,2 Giles W. Plant,5 Anthony J. Bakker,2 Alan R. Harvey,2 Igor Luzinov6 & K. Swaminathan Iyer1* 1 School of Chemistry and Biochemistry, The University of Western Australia, Crawley, WA 6009, Australia; 2 School of Anatomy, Physiology and Human Biology, The University of Western Australia, Crawley, WA 6009, Australia; 3 Institute for Integrated Cell-Material Sciences (iCeMS), iCeMS Complex 2, Kyoto University, Yoshida-Honmachi, Sakyo-ku, Kyoto, 606-8501, Japan; 4 Experimental and Regenerative Neurosciences, School of Animal Biology, The University of Western Australia, Crawley, WA 6009, Australia; 5 Stanford Partnership for Spinal Cord Injury and Repair, Department of Neurosurgery, Stanford University School of Medicine, Stanford, CA 94305, USA; 6 School of Materials Science and Engineering, Clemson University, Clemson, South Carolina, 29634-0971, USA. ‡ Present Address: Centre for NanoScale Science and Technology, School of Chemical and Physical Sciences, Flinders University, Bedford Park, Adelaide, SA 5042, Australia. * Correspondance: swaminatha.iyer@uwa.edu.au S1 Supplementary Information Materials. All chemicals were purchased from Sigma-Aldrich unless otherwise stated: iron(III) acetylacetonate (97 %), benzyl ether (98 %), oleic acid (90 %), oleyl amine (70 %), 1,2tetradecanediol (90 %), rhodamine B (Fluka), methyl ethyl ketone (99 %, Fisher), chloroform (99 %, merck), toluene (99 %, Fisher), diethyl ether (90 %, Asia Pacific Speciality Chemicals), polyethylenimine (50 % solution, Mn 1200, Mw 1300), Poly(3,4-ethylenedioxythiophene) Polystyrene sulfonate (1.3% solution, Mw 10355), ethylenediamine (99.5 %, Fluka) and Pluronic F-108. All tissue culture reagents were purchased from Gibco unless otherwise stated. Dulbecco's Modified Eagle's medium (DMEM), PBS, foetal calf serum (Sigma), horse serum (Sigma), penicillin/streptomycin, L-glutamine, non-essential amino acids (NEAA), trypsin/EDTA (Sigma), laminin (#L2020, Sigma) and nerve growth factor (β-NGF, PeproTech). Magnetite Synthesis. Fe3O4 was synthesized by the organic decomposition of Fe(acac)3 in benzyl ether at 300 oC, in the presence of oleic acid, oleyl amine, and 1,2- tetradecanediol, as previously described by Sun et al.1 The method to synthesise 6 nm Fe3O4 nanoparticles was followed. Synthesis of RhB-Modified PGMA: PGMA was synthesized by radical polymerization according to a published procedure.2 Briefly, glycidyl methacrylate was polymerized in methyl ethyl ketone (MEK) to give PGMA (Mn = 220515, Mw = 433730), using azobisisobutyronitrile as initiator. The polymer was purified by multiple precipitations from MEK solution using diethyl ether. To attach the dye to the polymer, a solution of rhodamine B (RhB, 20 mg) and PGMA (100 mg) in MEK (20 mL) was heated to reflux under N2 for 18 h. The solution was reduced in vacuo before the modified polymer was precipitated with diethyl ether (20 mL). The S2 polymer was redissolved in MEK and precipitated with ether twice to remove ungrafted RhB. PEDOT:PSS Multilayer Nanoparticle (NP) Synthesis. To prepare the organic phase of the emulsion, the dried RhB-PGMA polymer was initially dissolved in 2 mL of CHCl3 and dried under N2, leaving a sticky residue. This was redissolved in a 1:3 mixture of CHCl3 (2 mL) and MEK (6 mL) along with 25 mg of Fe3O4. This organic phase was added drop wise to a rapidly stirring aqueous solution of Pluronic F-108 (12.5 mg/mL, 30 mL). The emulsion was homogenised with a probe-type ultrasonic wand at the lowest setting for 1 min. The organic solvents were evaporated off overnight under a slow flow of N2. The suspension was purified via centrifugation at 3000 g for 45 mins. The supernatant was transferred to a 50 mL flask containing PEI (50 wt % solution, 100 mg) and heated to 80 oC for 16 h. The magnetic polymer nanoparticles were collected on a magnetic separation column (LS, Miltenyi Biotec) in 3 mL batches, washed with water (5 mL) and then flushed with water until the filtrate ran clear. This purified product produced 10 mL of nanoparticle suspension at a concentration of 1 mg/mL. Next, PEDOT:PSS was electrostatically attached to the nanoparticles. 60 µL of PEDOT:PSS (1.3 wt % dispersion in H2O) was diluted in 2 mL of water and added drop wise under rapid stirring to NPs at a concentration of 0.5 mg/mL. The PEDOT:PSS was further dispersed under sonication for 10 mins to ensure complete dispersion and then left to stir for 18 h. After 18 h, the mixture was again sonicated for 2 mins. Excess PEDOT:PSS was then removed via centrifugation 16800 g for 20 mins). NPs were then washed twice in water before being stored at 4 oC at a concentration of 4 mg/mL for further use. Nanoparticle Conductivity Measurement. The nanoparticle conductivity was determined using 4-point probe measurements. 80 µL of nanoparticle solution with a concentration of 1 mg/mL S3 was selectively dried on a square area 0.5 cm x 0.5 cm. Electrodes were placed at the 4 corners of the square and subject to current-voltage sweeps. Fabrication of PDMS Stamp. The metal layer of a blank compact disc (CD) was peeled off and the CD washed with ethanol. The remaining polycarbonate structure was used as a master for the PDMS stamp. The polymer base and curing agent from a Sylgard® 184 (Dow Corning) silicone elastomer kit were mixed at a 10:1 ratio by weight in a glass vial. The glass vial was placed in a vacuum desiccator to remove trapped bubbles from the mixture. Following vacuum treatment, the elastomer was restored to atmospheric pressure slowly several times until it was free of bubbles. The PDMS mixture was then cast onto the surface of the grooved side of CD and cured at 80 ºC for 2 hours. CFL Procedure. Prior to the CFL procedure, the indium tin oxide (ITO) coverslips were first clean in acetone and isopopanol under sonication. 0.2 % w/v PGMA in CHCl3 was spin coated onto the conducting surface of the ITO coverslips. Coverslips were then placed in an oven at 120 ºC for 20 mins to anneal the PGMA. Unreacted PGMA on the coverslip surface was removed by washing in CHCl3. Next, 1.3 % w/v PS in toluene was spin coated onto the PGMA surface. A PDMS stamp was then placed onto the PS layer, followed by heat treatment in an oven at 130 ºC for 1 hr. The assembly was then cooled down at room temperature for another hour before the PDMS stamp was peeled off. Next, the substrate was exposed to ethylenediamine (EA) and left at room temperature for 5 h. The substrate was then wasted with water to remove unreacted EA. Next, 50 µL of 4 mg/mL nanoparticle solution was drop casted onto the patterned area of the coverslip. The setup was then placed in a sealed vial, facilitating controlled evaporation which allowed for electrostatic nanoparticle attachment onto the EA surface. The PS mask was then S4 removed by toluene, leaving the patterned nanoparticle array. Cell Culture. The rat pheochromocytoma cells (PC12 cells) used here were obtained from Flinders University (Adelaide, Australia) courtesy of Professor Jacqueline Phillips (Macquarie University, Sydney, Australia). PC12 cells were cultured in P75 flasks in a humidified atmosphere containing proliferation media: 5 % CO2 at 37 oC and maintained in DMEM medium containing horse serum (10 % v/v), fetal calf serum (5 % v/v), penicillin/streptomycin (0.5 % v/v), L-glutamine (1 % v/v) and nonessential amino acids (1 % v/v). For PC12 differentiation, cells were cultured in differentiation media consisting of DMEM, L-glutamine (1 % v/v), horse serum (1 % v/v) and nerve growth factor (50 ng/mL). Electrical Stimulation Experiments. Prior to stimulation experiments, two silver epoxy electrodes were painted onto the ends of the prepared NP array and platinum wires attached to allow for connections with the stimulator. Next, a cell culture well was created by first cutting a 1.5 mL microcentrifuge tube in half and then sealing the capped end with silicon vacuum grease. This was stuck onto the ITO glass with the patterned arrays in the centre of the well. This was done to ensure that the electrodes did not come into direct contact with the cell culture media. The array was then placed in a Petri dish to maintain sterility throughout the course of the experiment (Fig S5a). Prior to culturing cells on the arrays, the coverslips were UV sterilised (20 mins) and then washed with 70 % ethanol three times. Wells were then coated with poly-(Llysine) and 15 µg/mL of laminin. Cells were then seeded at a density of 50 000 cells/well and left to adhere for 18 h. Prior to stimulation, the proliferation media was replaced with differentiation media. The cells were then stimulated according to protocols as listed below. Following stimulation, the cells were left for a further 72 h with fresh differentiating media added every 48 S5 h. Photographs of the electrical stimulation setup are described in Figure S5. Stimulation Protocol. The electrical signals were supplied by Grass S44 Stimulator (Quincy, Massachusetts, USA). The stimulation regime is similar to that used by Wallace et al.3-5 Briefly, the cells were subjected to a monophasic pulsed current at a frequency of 250 Hz with a 2 ms pulse width and an amplitude of 1 mA for 2 h. Cell Viability Assays. Cell viability was measured using the (3-(4,5-dimethylthiazol-2-yl)-5-(3carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium, inner salt) (MTS) assay as per the manufacturer protocols (Invitrogen, UK). Cells were plated as per “electrical stimulation protocol” stated above. Viability was to be determined at 3 time points: (i) 0 h (immediately prior to electrical stimulation), (ii) 72 h after the addition of differentiation media and (iii) 72 h after the addition of differentiation media and electrical stimulation. For measurements, 80 µL from each well was transferred into a new 96 well plate and read under a plate reader at 490 nm excitation wavelength. The same protocol was followed for every sample and each measurement was carried out in triplicate. Immunohistochemical Staining. The PC12 cells were immunohistochemically stained for ß-III tubulin. The cells were fixed in 4 % paraformaldehyde for 10 mins. Cells were first incubated with a primary antibody solution containing PBS, 10 % Normal Goat Serum, 0.1 % Triton X-100 and the anti-β-III tubulin antibody (1:1000, anti-rabbit, Covance) at room temperature for 30 mins. After 3 PBS washes, the antibody binding was visualised with anti-rabbit FITC (1:100, Sigma) following incubation for 30 mins at room temperature. Coverslips were mounted on glass slides covered with Dako Fluorescent Mounting Medium (Dako, USA). All experiments were S6 performed in triplicate. Confocal and Fluorescence Microscopy Analysis. Immunohistochemically stained samples were analysed using confocal and fluorescence microscopy. Confocal microscopy was carried out using a Leica TCS SP2 AOBS Multiphoton Confocal microscope and fluorescence microscopy with a Diaplan fluorescence microscope. Image and Statistical Analysis. To determine the effects both stimulation and the NP arrays had on the PC12 cells, the average area of each cell was determined. 3 randomly selected areas on each sample was visualised at 40 x magnification. The average area covered by each cell was assessed using Image J analysis software (version 1.48a, NIH). All immunohistochemical analyses were conducted by a single investigator, ensuring constant selection criteria, and results expressed as means ± SD. Data were analysed using Origin data management software to conduct ANOVA on groups of data. Statistically significant differences between each treatment were determined using Bonferroni/Dunn post hoc tests (p≤0.05). Scanning Electron Microscopy (SEM). Prior to SEM imaging, samples without cells were coated with 5 nm of Pt. Samples with cells were fixed in 2.5 % glutaraldehyde for 2 h at 4 oC and dehydrated. Samples were washed with deionized water and dehydrated in a microwave in serial concentrations of ethanol (50 %, 70 % and 90 % once then 3x in absolute ethanol), before critical point drying with carbon dioxide for 1h and then coating with 5 nm of Pt. Samples were imaged using a Zeiss 1555 VP-FESEM. Transmission Electron Microscopy (TEM). Synthesized polymer nanoparticles were dropcasted on carbon coated TEM grids and imaged with an accelerating voltage of 100 kV on a S7 JEOL 2100 transmission electron microscope. Atomic Force Microscopy. A Dimension 3100 AFM system (Bruker) with a Nanoscope IV controller (Bruker) was used to obtain the AFM images in Contact Mode, using Pt/Ir coated contact mode probes with a spring constant of 0.2 N/m (type SCM-PIC, Bruker). The scan parameters were adjusted to ensure reliable imaging with the smallest possible contact force setpoint. Data analysis was performed using the SPM analysis freeware Gwyddion (http://gwyddion.net). Dynamic light scattering (DLS) and zeta potential measurements. DLS experiments were performed using a Malvern Zetasizer Nano series. For measuring the size distribution, 5 measurements were taken and in each measurement there were 10 data acquisitions. Zeta potential (ζ) measurements were performed using the same instrument. Measurements for each sample were recorded in triplicate and 100 data acquisitions were recorded in each measurement. All measurements were recorded at 25 oC in Malvern disposable clear Folded Capillary Cells. S8 Supplementary Figures (a) (b) Figure S1. (a) The PDMS stamp used in the study. SEM micrograph of the grooved structure of the PDMS. Image taken at 6k x magnification; (b) Photograph of the polycarbonate disc peeled from a CD. PDMS was cast on the grooved surface and stamps of the desired size were cut out. S9 T > Tg (PS) (130 oC) Peel off PDMS Stamp EA grafting onto PGMA Conducting NPs Electrostatic attachment PS removal Figure S2. Schematic of CFL procedure. Briefly, ITO substrate was modified with a thin layer of PGMA followed by second layer of PS; a PDMS stamp was placed over the PS film and heat treated at 130 oC; PDMS stamp was peeled off after cooling; EA was selective reacted to the exposed PGMA stripes to produce cationic stripes to enable charge complementarity to assemble the anionic PEDOT:PSS nanoparticles. The PS mask was removed by washing with toluene, to obtain linear PEDOT:PSS conduits. S10 Figure S3. SEM micrograph of PC12 cells 18 h after plating and immediately prior to electrical stimulation. PC12 cells on the patterned surface (yellow box) were evenly spread out, in comparison to the rounded cells on non-patterned areas of the substrate demonstrating preferential adhesion. Image taken at 1000 x magnification. S11 (a) (b) Figure S4. Representative low magnification (magnification 400 x) SEM micrographs of PC12 cells 72 h after the following treatments: (a) (+) pattern, (-) stimulation and (b) (+) pattern, (+) stimulation demonstrating lower coverage due to reduction in proliferation upon stimulation. . S12 (b) (a) (c) Figure S5. Photographs of the electrical stimulation setup: (a) A sterile Petri dish containing the modified cell culture well (red arrow) and the Platinum wires which allow for connections to the stimulator (green arrows), (b) The stimulator (yellow arrow) was placed next to an incubator and the wires from the machine leading into the stimulator (blue arrows), (c) The wires from the stimulator were connected to the platinum wires via alligator clips. S13 References 1. Sun, S.; Zeng, H.; Robinson, D. B.; Raoux, S.; Rice, P. M.; Wang, S. X.; Li, G. Monodisperse MFe2O4 (M = Fe, Co, Mn) Nanoparticles. J. Am. Chem. Soc. 2004, 126, 273-279. 2. Tsyalkovsky, V.; Klep, V.; Ramaratnam, K.; Lupitskyy, R.; Minko, S.; Luzinov, I. Fluorescent Reactive Core–Shell Composite Nanoparticles With a High Surface Concentration of Epoxy Functionalities. Chem. Mat. 2007, 20, 317-325. 3. Liu, X.; Gilmore, K. J.; Moulton, S. E.; Wallace, G. G. Electrical Stimulation Promotes Nerve Cell Differentiation on Polypyrrole/Poly (2-Methoxy-5 Aniline Sulfonic Acid) Composites. J. Neural Eng. 2009, 6, 065002. 4. Weng, B.; Liu, X.; Shepherd, R.; Wallace, G. G. Inkjet Printed Polypyrrole/Collagen Scaffold: A Combination of Spatial Control and Electrical Stimulation of PC12 Cells. Synt. Met. 2012, 162, 1375-1380. 5. Richardson, R. T.; Thompson, B.; Moulton, S.; Newbold, C.; Lum, M. G.; Cameron, A.; Wallace, G.; Kapsa, R.; Clark, G.; O’Leary, S. The Effect of Polypyrrole with Incorporated Neurotrophin-3 on the Promotion of Neurite Outgrowth from Auditory Neurons. Biomaterials 2007, 28, 513-523. S14 References 1. G. G. Gauglitz, H. C. Korting, T. Pavicic, T. Ruzicka and M. G. Jeschke, Mol. Med., 2011, 17, 113. 2. A. Bayat, D. A. McGrouther and M. W. J. Ferguson, The BMJ, 2003, 326, 88-92. 3. R. F. Pereira, C. C. Barrias, P. L. Granja and P. J. Bartolo, Nanomedicine, 2013, 8, 603-621. 4. L. Yildirimer, N. T. K. Thanh and A. M. Seifalian, Trends Biotechnol., 2012, 30, 638-648. 5. T. Dvir, B. P. Timko, D. S. Kohane and R. Langer, Nat. Nanotechnol., 2011, 6, 1322. 6. R. Langer and J. Vacanti, Science, 1993, 260, 920-926. 7. E. S. Place, N. D. Evans and M. M. Stevens, Nat. Mater., 2009, 8, 457-470. 8. F. Edalat, I. Sheu, S. Manoucheri and A. Khademhosseini, Curr. Opin. Chem. Biol., 2012, 23, 820-825. 9. F. J. O’Brien, B. Harley, I. V. Yannas and L. J. Gibson, Biomaterials, 2005, 26, 433-441. 10. Y. C. Chen, R. Z. Lin, H. Qi, Y. Yang, H. Bae, J. M. Melero‐Martin and A. Khademhosseini, Adv. Funct. Mater., 2012, 22, 2027-2039. 11. T. Freyman, I. Yannas and L. Gibson, Prog. Mater Sci., 2001, 46, 273-282. 12. H. K. Kleinman and G. R. Martin, Seminars in cancer biology, 2005. 13. J. F. Mano, G. A. Silva, H. S. Azevedo, P. B. Malafaya, R. A. Sousa, S. S. Silva, L. F. Boesel, J. M. Oliveira, T. C. Santos, A. P. Marques, N. M. Neves and R. L. Reis, J. R. Soc. Interface, 2007, 4, 999-1030. 14. P. X. Ma, Adv. Drug Delivery Rev., 2008, 60, 184-198. 15. D. Li and Y. Xia, Nano Lett., 2004, 4, 933-938. 16. D. Wahl and J. Czernuszka, Eur. Cell. Mater., 2006, 11, 43-56. 147 17. C. Rodrigues, P. Serricella, A. Linhares, R. Guerdes, R. Borojevic, M. Rossi, M. Duarte and M. Farina, Biomaterials, 2003, 24, 4987-4997. 18. D. A. Wahl, E. Sachlos, C. Liu and J. T. Czernuszka, J. Mater. Sci. Mater. Med., 2007, 18, 201-209. 19. J. Venugopal, S. Low, A. T. Choon, T. S. Kumar and S. Ramakrishna, J. Mater. Sci. Mater. Med., 2008, 19, 2039-2046. 20. M. W. Tibbitt and K. S. Anseth, Biotechnol. Bioeng., 2009, 103, 655-663. 21. S. R. Shin, H. Bae, J. M. Cha, J. Y. Mun, Y.-C. Chen, H. Tekin, H. Shin, S. Farshchi, M. R. Dokmeci and S. Tang, ACS nano, 2011, 6, 362-372. 22. N. Annabi, S. M. Mithieux, E. A. Boughton, A. J. Ruys, A. S. Weiss and F. Dehghani, Biomaterials, 2009, 30, 4550-4557. 23. D. Putnam, Nat. Mater., 2006, 5, 439-451. 24. D. H. Reneker and I. Chun, Nanotechnology, 1996, 7, 216. 25. S. Zhang, Nat. Biotechnol., 2003, 21, 1171-1178. 26. A. Khademhosseini, R. Langer, J. Borenstein and J. P. Vacanti, Proc. Natl. Acad. Sci. U.S.A., 2006, 103, 2480-2487. 27. X. Yang, J. Loos, S. C. Veenstra, W. J. Verhees, M. M. Wienk, J. M. Kroon, M. A. Michels and R. A. Janssen, Nano Lett., 2005, 5, 579-583. 28. S. Nemir and J. L. West, Ann. Biomed. Eng., 2010, 38, 2-20. 29. K. K. Parker and D. E. Ingber, Phil. Trans. R. Soc. B, 2007, 362, 1267-1279. 30. C. Fink, S. Ergun, D. Kralisch, U. Remmers, J. Weil and T. Eschenhagen, FASEB J., 2000, 14, 669-679. 31. M. Radisic, H. Park, H. Shing, T. Consi, F. J. Schoen, R. Langer, L. E. Freed and G. Vunjak-Novakovic, Proc. Natl. Acad. Sci., 2004, 101, 18129-18134. 32. N. Bursac, K. Parker, S. Iravanian and L. Tung, Circ. Res., 2002, 91, e45-e54. 33. D.-H. Kim, E. A. Lipke, P. Kim, R. Cheong, S. Thompson, M. Delannoy, K.-Y. Suh, L. Tung and A. Levchenko, Proc. Natl. Acad. Sci., 2010, 107, 565-570. 34. M. Lovett, K. Lee, A. Edwards and D. L. Kaplan, Tissue Eng Part B Rev., 2009, 15, 353-370. 35. S. Chung and M. W. King, Biotechnol. Appl. Biochem., 2011, 58, 423-438. 36. T. Karande, J. Ong and C. M. Agrawal, Ann. Biomed. Eng., 2004, 32, 1728-1743. 148 37. B. J. Papenburg, J. Liu, G. A. Higuera, A. M. C. Barradas, J. de Boer, C. A. van Blitterswijk, M. Wessling and D. Stamatialis, Biomaterials, 2009, 30, 6228-6239. 38. H. Ko, C. Mcfarland and B. Milthorpe, European Cells and Materials, 2007, 14, 119. 39. S. Yang, K.-F. Leong, Z. Du and C.-K. Chua, Tissue Eng., 2001, 7, 679-689. 40. L. E. Freed, F. Guilak, X. E. Guo, M. L. Gray, R. Tranquillo, J. W. Holmes, M. Radisic, M. V. Sefton, D. Kaplan and G. Vunjak-Novakovic, Tissue Eng., 2006, 12, 3285-3305. 41. A. J. Salgado, O. P. Coutinho and R. L. Reis, Macromol. Biosci., 2004, 4, 743-765. 42. K. F. Leong, C. M. Cheah and C. K. Chua, Biomaterials, 2003, 24, 2363-2378. 43. T. Ma, Y. Li, S.-T. Yang and D. A. Kniss, Biotechnol. Bioeng., 2000, 70, 606-618. 44. J. Zeltinger, J. K. Sherwood, D. A. Graham, R. Müeller and L. G. Griffith, Tissue Eng., 2001, 7, 557-572. 45. J. H. Brauker, V. E. Carr-Brendel, L. A. Martinson, J. Crudele, W. D. Johnston and R. C. Johnson, J. Biomed. Mater. Res., 1995, 29, 1517-1524. 46. A. Rosengren and L. M. Bjursten, J. Biomed. Mater. Res. A, 2003, 67A, 918-926. 47. B.-S. Kim and D. J. Mooney, J. Biomed. Mater. Res., 1998, 41, 322-332. 48. L. E. Freed, G. Vunjak-Novakovic, R. J. Biron, D. B. Eagles, D. C. Lesnoy, S. K. Barlow and R. Langer, Nat. Biotechnol., 1994, 12, 689-693. 49. S. H. Lee, B. S. Kim, S. H. Kim, S. W. Kang and Y. H. Kim, Macromol. Biosci., 2004, 4, 802-810. 50. K. C. Ang, K. F. Leong, C. K. Chua and M. Chandrasekaran, Rapid Prototyping J., 2006, 12, 100-105. 51. D. L. Butler, S. A. Goldstein and F. Guilak, J. Biomech. Eng., 2000, 122, 570-575. 52. D. Dado and S. Levenberg, Sem Cell Dev. Biol., 2009, 20, 656-664. 53. D. E. Discher, P. Janmey and Y.-l. Wang, Science, 2005, 310, 1139-1143. 54. M. B. Keogh, F. J. O’Brien and J. S. Daly, Acta Biomater., 2010, 6, 4305-4313. 55. A. J. Engler, S. Sen, H. L. Sweeney and D. E. Discher, Cell, 2006, 126, 677-689. 56. P. Roach, T. Parker, N. Gadegaard and M. R. Alexander, Surf. Sci. Rep., 2010, 65, 145-173. 57. A. Curtis and C. Wilkinson, Biomaterials, 1997, 18, 1573-1583. 149 58. V. Raghunathan, C. McKee, W. Cheung, R. Naik, P. F. Nealey, P. Russell and C. J. Murphy, Tissue Eng. Pt. A, 2013, 19, 1713-1722. 59. Y. Tamada and Y. Ikada, J. Biomed. Mater. Res., 1994, 28, 783-789. 60. J. Lee, M. J. Cuddihy and N. A. Kotov, Tissue Eng Part B Rev., 2008, 14, 61-86. 61. C. Sangwon, G. Mike P. and K. Martin W., Biomed. Mater., 2011, 6, 045001. 62. M. M. Stevens and J. H. George, Science, 2005, 310, 1135-1138. 63. Z.-Z. Wu, W. S. Kisaalita, L. Wang, A. L. Zachman, Y. Zhao, K. Hasneen, D. Machacek and S. L. Stice, Biotechnol. Bioeng., 2010, 106, 649-659. 64. A. Graziano, R. d'Aquino, M. G. C.-D. Angelis, F. De Francesco, A. Giordano, G. Laino, A. Piattelli, T. Traini, A. De Rosa and G. Papaccio, J. Cell. Physio., 2008, 214, 166-172. 65. M. Singer, R. H. Nordlander and M. Egar, J. Comp. Neurol., 1979, 185, 1-21. 66. P. Martin, Science, 1997, 276, 75-81. 67. B. Zavan, V. Vindigni, R. Cortivo and G. Abatangelo, in Tissue Eng., ed. D. Eberli, InTech, 2010, pp. 509-524. 68. A. D. Metcalfe and M. W. J. Ferguson, Biomaterials, 2007, 28, 5100-5113. 69. S. MacNeil, Nature, 2007, 445, 874-880. 70. A. J. Singer and R. A. F. Clark, N. Engl. J. Med., 1999, 341, 738-746. 71. V. Falanga, Lancet, 2005, 366, 1736-1743. 72. E. J. Brown, Bioessays, 1995, 17, 109-117. 73. D. A. Rappolee, D. Mark, M. J. Banda and Z. Werb, Science, 1988, 241, 708-712. 74. S. Leibovich and R. Ross, Am. J. Pathol., 1975, 78, 71. 75. T. J. Koh and L. A. DiPietro, Expert Rev. Mol. Med., 2011, 13, e23. 76. M. L. Usui, R. A. Underwood, J. N. Mansbridge, L. A. Muffley, W. G. Carter and J. E. Olerud, Wound Repair Regen., 2005, 13, 468-479. 77. W. S. Krawczyk, J. Cell Biol., 1971, 49, 247-263. 78. R. D. Paladini, K. Takahashi, N. S. Bravo and P. A. Coulombe, J. Cell Biol., 1996, 132, 381-397. 79. G. P. Radice, Dev. Biol., 1980, 76, 26-46. 80. R. Clark, The molecular and cellular biology of wound repair, Springer, 1996. 81. J. A. Goliger and D. L. Paul, Mol. Biol. Cell, 1995, 6, 1491-1501. 150 82. G. Gabbiani, C. Chaponnier and I. Hüttner, J. Cell Biol., 1978, 76, 561-568. 83. R. A. F. Clark, J. Invest. Dermatol., 1990, 94, 128s-134s. 84. H. Larjava, T. Salo, K. Haapasalmi, R. Kramer and J. Heino, J. Clin. Invest., 1993, 92, 1425. 85. R. Clark, G. Ashcroft, M. J. Spencer, H. Larjava and M. Ferguson, Br. J. Dermatol., 1996, 135, 46-51. 86. T. H. Bugge, K. W. Kombrinck, M. J. Flick, C. C. Daugherty, M. J. S. Danton and J. L. Degen, Cell, 1996, 87, 709-719. 87. S. Werner, H. Smola, X. Liao, M. T. Longaker, T. Krieg, P. H. Hofschneider and L. T. Williams, Science, 1994, 266, 819-822. 88. R. A. Clark, J. M. Lanigan, P. DellaPelle, E. Manseau, H. F. Dvorak and R. B. Colvin, J. Invest. Dermatol., 1982, 79, 264-269. 89. M. Robson, L. Phillips, L. Robson, A. Thomason and G. Pierce, Lancet, 1992, 339, 23-25. 90. D. L. Steed and S. Group, J. Vasc. Surg., 1995, 21, 71-81. 91. M. Vaalamo, L. Mattila, N. Johansson, A.-L. Kariniemi, M.-L. KarjalainenLindsberg, V.-M. Kähäri and U. Saarialho-Kere, J. Invest. Dermatol., 1997, 109, 96101. 92. R. A. Clark, L. D. Nielsen, M. P. Welch and J. M. McPherson, J. Cell Sci., 1995, 108, 1251-1261. 93. M. P. Welch, G. F. Odland and R. Clark, J. Cell Biol., 1990, 110, 133-145. 94. A. Weiss and L. Attisano, Wiley Interdiscip Rev Dev Biol., 2013, 2, 47-63. 95. S. O'Kane and M. W. J. Ferguson, Int. J. Biochem. Cell Biol., 1997, 29, 63-78. 96. J. P. Annes, J. S. Munger and D. B. Rifkin, J. Cell Sci., 2003, 116, 217-224. 97. H. J. You, T. How and G. C. Blobe, Carcinogenesis, 2009, 30, 1281-1287. 98. G. C. Blobe, W. P. Schiemann and H. F. Lodish, N. Engl. J. Med., 2000, 342, 13501358. 99. S. Radaev, Z. Zou, T. Huang, E. M. Lafer, A. P. Hinck and P. D. Sun, J. Biol. Chem., 2010, 285, 14806-14814. 100. E. P. Bottinger, J. J. Letterio and A. B. Roberts, Kidney. Int., 1997, 51, 1355-1360. 151 101. M. C. Dickson, J. S. Martin, F. M. Cousins, A. B. Kulkarni, S. Karlsson and R. J. Akhurst, Development, 1995, 121, 1845-1854. 102. H. Hayashi and T. Sakai, Front. Physiol., 2012, 3. 103. L. P. Sanford, I. Ormsby, A. C. Gittenberger-de Groot, H. Sariola, R. Friedman, G. P. Boivin, E. L. Cardell and T. Doetschman, Development, 1997, 124, 2659-2670. 104. G. Proetzel, S. A. Pawlowski, M. V. Wiles, M. Yin, G. P. Boivin, P. N. Howles, J. Ding, M. W. Ferguson and T. Doetschman, Nat. Genet., 1995, 11. 105. J. S. Munger, J. G. Harpel, P.-E. Gleizes, R. Mazzieri, I. Nunes and D. B. Rifkin, Kidney. Int., 1997, 51, 1376-1382. 106. A. M. Brunner, H. Marquardt, A. R. Malacko, M. N. Lioubin and A. F. Purchio, J. Biol. Chem., 1989, 264, 13660-13664. 107. J. Massagué, Cell, 2008, 134, 215-230. 108. R. Ebner, R. Chen, S. Lawler, T. Zioncheck and R. Derynck, Science, 1993, 262, 900-902. 109. R. Ebner, R. Chen, L. Shum, S. Lawler, T. Zioncheck, A. Lee, A. Lopez and R. Derynck, Science, 1993, 260, 1344-1348. 110. A. Moustakas and C.-H. Heldin, Development, 2009, 136, 3699-3714. 111. C. Bernabeu, J. M. Lopez-Novoa and M. Quintanilla, BBA-Mol. Basis Dis., 2009, 1792, 954-973. 112. G. C. Blobe, X. Liu, S. J. Fang, T. How and H. F. Lodish, J. Biol. Chem., 2001, 276, 39608-39617. 113. L. Attisano and J. L. Wrana, Science, 2002, 296, 1646-1647. 114. F. Verrecchia and A. Mauviel, J. Invest. Dermatol., 2002, 118, 211-215. 115. C. S. Hill, Cell Res., 2009, 19, 36-46. 116. K. H. Wrighton, X. Lin and X.-H. Feng, Cell Res., 2009, 19, 8-20. 117. A. Leask and D. J. Abraham, FASEB J., 2004, 18, 816-827. 118. Y. Shi and J. Massagué, Cell, 2003, 113, 685-700. 119. S. Van Vlierberghe, P. Dubruel and E. Schacht, Biomacromolecules, 2011, 12, 1387-1408. 120. R. O. Hynes, Cell, 1992, 69, 11-25. 121. R. O. Hynes, Cell, 2002, 110, 673-687. 152 122. K. Minton, Nat. Rev. Mol. Cell Biol., 2013, 14, 401-401. 123. J. S. Munger, X. Huang, H. Kawakatsu, M. J. D. Griffiths, S. L. Dalton, J. Wu, J.-F. Pittet, N. Kaminski, C. Garat, M. A. Matthay, D. B. Rifkin and D. Sheppard, Cell, 1999, 96, 319-328. 124. M. Shi, J. Zhu, R. Wang, X. Chen, L. Mi, T. Walz and T. A. Springer, Nature, 2011, 474, 343-349. 125. S. E. Crawford, V. Stellmach, J. E. Murphy-Ullrich, S. M. F. Ribeiro, J. Lawler, R. O. Hynes, G. P. Boivin and N. Bouck, Cell, 1998, 93, 1159-1170. 126. S. M. F. Ribeiro, M. Poczatek, S. Schultz-Cherry, M. Villain and J. E. MurphyUllrich, J. Biol. Chem., 1999, 274, 13586-13593. 127. S. Schultz-Cherry, S. Ribeiro, L. Gentry and J. E. Murphy-Ullrich, J. Biol. Chem., 1994, 269, 26775-26782. 128. J. E. Murphy-Ullrich and M. Poczatek, Cytokine Growth Factor Rev., 2000, 11, 5969. 129. Q. Yu and I. Stamenkovic, Genes Dev., 2000, 14, 163-176. 130. Y. Sato and D. B. Rifkin, J. Cell Biol., 1989, 109, 309-315. 131. R. M. Lyons, J. Keski-Oja and H. L. Moses, J. Cell Biol., 1988, 106, 1659-1665. 132. Y. Sato, R. Tsuboi, R. Lyons, H. Moses and D. Rifkin, J. Cell Biol., 1990, 111, 757763. 133. I. Nunes, R. L. Shapiro and D. B. Rifkin, J. Immunol., 1995, 155, 1450-1459. 134. N. Khalil, S. Corne, C. Whitman and H. Yacyshyn, Am J Respir Cell Mol Biol, 1996, 15, 252-259. 135. C. Margadant and A. Sonnenberg, EMBO Rep, 2010, 11, 97-105. 136. T. Chan, A. Ghahary, J. Demare, L. Yang, T. Iwashina, P. Scott and E. E. Tredget, Wound Repair Regen., 2002, 10, 177-187. 137. E. B. Tredget, J. Demare, G. Chandran, E. E. Tredget, L. Yang and A. Ghahary, Wound Repair Regen., 2005, 13, 61-67. 138. L. Yang, T. Chan, J. Demare, T. Iwashina, A. Ghahary, P. G. Scott and E. E. Tredget, Am. J. Pathol., 2001, 159, 2147-2157. 139. S. Fang, N. Pentinmikko, M. Ilmonen and P. Salven, Angiogenesis, 2012, 15, 511519. 153 140. G. Ferrari, B. D. Cook, V. Terushkin, G. Pintucci and P. Mignatti, J. Cell. Physio., 2009, 219, 449-458. 141. A. B. Roberts, Wound Repair Regen., 1995, 3, 408-418. 142. A. F. Purchio, J. A. Cooper, A. M. Brunner, M. N. Lioubin, L. E. Gentry, K. S. Kovacina, R. A. Roth and H. Marquardt, J. Biol. Chem., 1988, 263, 14211-14215. 143. A. Gosiewska, C.-F. Yi, O. Blanc-Brude and J. Geesin, Methods Cell Sci., 1999, 21, 47-56. 144. S. J. Bates, E. Morrow, A. Y. Zhang, H. Pham, M. T. Longaker and J. Chang, J. Bone Joint Surg. Am., 2006, 88, 2465-2472. 145. Xia C, Yang X, Yingzhen W, Sun K and T. S., Orthopedics, 2010, 33, 727. 146. S. Vidal, C. Vidil, A. Morère, M. Garcia and J.-L. Montero, Eur. J. Org. Chem., 2000, 2000, 3433-3437. 147. P. Y. Tong and S. Kornfeld, J. Biol. Chem., 1989, 264, 7970-7975. 148. P. Y. Tong, W. Gregory and S. Kornfeld, J. Biol. Chem., 1989, 264, 7962-7969. 149. L. J. Olson, N. M. Dahms and J.-J. P. Kim, J. Biol. Chem., 2004, 279, 34000-34009. 150. J. J. Distler, J. F. Guo, G. W. Jourdian, O. P. Srivastava and O. Hindsgaul, J. Biol. Chem., 1991, 266, 21687-21692. 151. S. Vidal, M. Garcia, J.-L. Montero and A. Morère, Bioorgan. Med. Chem., 2002, 10, 4051-4056. 152. UK Pat., 2012. 153. Renovo, in http://clinicaltrials.gov/ct2/show/study/NCT00664352, ClinicalTrials.gov [Internet], UK, 2009, vol. 2014. 154. A. I. Arno, G. G. Gauglitz, J. P. Barret and M. G. Jeschke, Burns, 2014. 155. C. Vidil, A. Morère, M. Garcia, V. Barragan, B. Hamdaoui, H. Rochefort and J.-L. Montero, Eur. J. Org. Chem., 1999, 1999, 447-450. 156. D. B. Berkowitz, G. Maiti, B. D. Charette, C. D. Dreis and R. G. MacDonald, Org. Lett., 2004, 6, 4921-4924. 157. A. Jeanjean, M. Gary-Bobo, P. Nirdé, S. Leiris, M. Garcia and A. Morère, Bioorg. Med. Chem. Lett., 2008, 18, 6240-6243. 158. A. Jeanjean, M. Garcia, A. Leydet, J.-L. Montero and A. Morère, Bioorgan. Med. Chem., 2006, 14, 3575-3582. 154 159. M. Gary-Bobo, P. Nirde, A. Jeanjean, A. Morere and M. Garcia, Curr. Med. Chem., 2007, 14, 2945-2953. 160. M. K. Christensen, M. Meldal, K. Bock, H. Cordes, S. Mouritsen and H. Elsner, J. Chem. Soc., Perkin Trans. 1, 1994, 1299-1310. 161. H. Franzyk, M. K. Christensen, R. M. Jørgensen, M. Meldal, H. Cordes, S. Mouritsen and K. Bock, Bioorgan. Med. Chem., 1997, 5, 21-40. 162. G. P. Sidgwick and A. Bayat, J. Eur. Acad. Dermatol. Venereol., 2012, 26, 141-152. 163. P. Durani, D. A. McGrouther and M. W. Ferguson, Plast. Reconstr. Surg., 2009, 123, 1481-1489. 164. T. a. Sullivan, J. Smith, J. Kermode, E. Mclver and D. Courtemanche, J. Burn Care Res., 1990, 11, 256-260. 165. M. J. Baryza and G. A. Baryza, J. Burn Care Res., 1995, 16, 535-538. 166. A. E. Slemp and R. E. Kirschner, Curr. Opin. Pediatr., 2006, 18, 396-402. 167. T.-L. Tuan and L. S. Nichter, Mol. Med. Today, 1998, 4, 19-24. 168. M. Babu, R. Diegelmann and N. Oliver, J. Invest. Dermatol., 1992, 99, 650-655. 169. O. Koese and A. Waseem, Dermatol. Surg., 2008, 34, 336-346. 170. F. Groeber, M. Holeiter, M. Hampel, S. Hinderer and K. Schenke-Layland, Adv. Drug Delivery Rev., 2011, 63, 352-366. 171. R. V. Shevchenko, S. L. James and S. E. James, J. R. Soc. Interface, 2010, 7, 229258. 172. A. E. Van der Merwe and W. Steenkamp, Ann. Burns Fire Disasters, 2012, 25, 188. 173. B. S. Atiyeh, S. N. Hayek and S. W. Gunn, Burns, 2005, 31, 944-956. 174. B. A. Rubis, D. Danikas, M. Neumeister, W. G. Williams, H. Suchy and S. M. Milner, Burns, 2002, 28, 752-759. 175. S. E. James, S. Booth, P. Gilbert, I. Jones and R. Shevchenko, in Strategies in Regenerative Medicine, ed. M. Santin, Springer New York, 2009, pp. 1-33. 176. H. Bannasch, M. Föhn, T. Unterberg, A. D. Bach, B. Weyand and G. B. Stark, Clin. Plast. Surg., 2003, 30, 573-579. 177. J. G. Rheinwatd and H. Green, Cell, 1975, 6, 331-343. 178. N. O'Connor, J. Mulliken, S. Banks-Schlegel, O. Kehinde and H. Green, Lancet, 1981, 317, 75-78. 155 179. D. Chester, D. Balderson and R. Papini, J. Burn Care Res., 2004, 25, 266-275. 180. R. J. Pye, Eye, 1988, 2, 172-178. 181. M. Rouabhia, Transplantation, 1996, 61, 1290-1300. 182. L. W. Rue III, W. G. Cioffi, W. F. McManus and B. A. Pruitt Jr, J. Trauma Acute Care, 1993, 34, 662-668. 183. S. Böttcher-Haberzeth, T. Biedermann and E. Reichmann, Burns, 2010, 36, 450460. 184. F. Braye, P. Pascal, M. Bertin-Maghit, J.-J. Colpart, E. Tissot and O. Damour, Med. Biol. Eng. Comp., 2000, 38, 248-252. 185. H. Kaiser, G. Stark, J. Kopp, A. Balcerkiewicz, G. Spilker and H. Kreysel, Burns, 1994, 20, 23-29. 186. M. Ghosh, S. Boyce, E. Freedlander and S. MacNeil, J. Burn Care Res., 1995, 16, 407-417. 187. J. Prenosil and M. Kino‐Oka, Ann. N Y Acad. Sci., 1999, 875, 386-397. 188. G. Gravante, M. C. Di Fede, A. Araco, M. Grimaldi, B. De Angelis, A. Arpino, V. Cervelli and A. Montone, Burns, 2007, 33, 966-972. 189. C. Zweifel, C. Contaldo, C. Köhler, A. Jandali, W. Künzi and P. Giovanoli, J. Plast. Reconstr. Aesthet. Surg., 2008, 61, e1-e4. 190. N. Zhu, R. Warner, C. Simpson, M. Glover, C. Hernon, J. Kelly, S. Fraser, T. Brotherston, D. Ralston and S. MacNeil, Eur. J. Plast. Surg., 2005, 28, 319-330. 191. D.-h. Hu, Z.-f. Zhang, Y.-g. Zhang, W.-f. Zhang, H.-t. Wang, W.-x. Cai, X.-z. Bai, H.-y. Zhu, J.-h. Shi and C.-w. Tang, Burns, 2012, 38, 702-712. 192. R. Renner, W. Harth and J. C. Simon, International Wound Journal, 2009, 6, 226232. 193. A. Khachemoune, Cws, Y. M. Bello and T. J. Phillips, Dermatol. Surg., 2002, 28, 274-280. 194. M. Moustafa, C. Simpson, M. Glover, R. A. Dawson, S. Tesfaye, F. M. Creagh, D. Haddow, R. Short, S. Heller and S. MacNeil, Diabetic Medicine, 2004, 21, 786-789. 195. A.-K. Tausche, M. Skaria, L. Böhlen, K. Liebold, J. Hafner, H. Friedlein, M. Meurer, R. J. Goedkoop, U. Wollina, D. Salomon and T. Hunziker, Wound Repair Regen., 2003, 11, 248-252. 156 196. A. Rejzek, F. Weyer, R. Eichberger and W. Gebhart, Cell Tissue Bank., 2001, 2, 95102. 197. K. Gajiwala and A. L. Gajiwala, Cell Tissue Bank., 2004, 5, 73-80. 198. M. Haberal, Z. Oner, U. Bayraktar and N. Bilgin, Burns, 1987, 13, 159-163. 199. M. R. Kesting, K.-D. Wolff, B. Hohlweg-Majert and L. Steinstraesser, J. Burn Care Res., 2008, 29, 907-916. 200. M. Ueta, M. N. KWEON, Y. Sano, C. Sotozono, J. Yamada, N. Koizumi, H. Kiyono and S. Kinoshita, Clin. Exp. Immunol., 2002, 129, 464-470. 201. W. C. Quinby Jr, H. C. Hoover, M. Scheflan, P. T. Walters, S. A. Slavin and C. C. Bondoc, Plast. Reconstr. Surg., 1982, 70, 711-716. 202. C. Sawhney, Burns, 1989, 15, 339-342. 203. E. S. Garfein, D. P. Orgill and J. J. Pribaz, Clin. Plast. Surg., 2003, 30, 485-498. 204. J. F. Hansbrough, D. W. Mozingo, G. P. Kealey, M. Davis, A. Gidner and G. D. Gentzkow, J. Burn Care Res., 1997, 18, 43-51. 205. E. Freedlander, S. Boyce, M. Ghosh, D. Ralston and S. MacNeil, Burns, 1998, 24, 19-24. 206. J. Vuola and D. Pipping, Burns, 2002, 28, 31-33. 207. M. Obeng, R. McCauley, J. Barnett, J. Heggers, K. Sheridan and S. Schutzler, Burns, 2001, 27, 267-271. 208. J. B. I. Yannas, W. Quinby Jr, C. Bondoc and W. Jung, Ann. Surg., 1981, 194, 413. 209. V. Jayarama Reddy, S. Radhakrishnan, R. Ravichandran, S. Mukherjee, R. Balamurugan, S. Sundarrajan and S. Ramakrishna, Wound Repair Regen., 2013, 21, 1-16. 210. K. A. Wright, K. B. Nadire, P. Busto, R. Tubo, J. M. McPherson and B. M. Wentworth, Burns, 1998, 24, 7-17. 211. S. R. Myers, V. N. Partha, C. Soranzo, R. D. Price and H. A. Navsaria, Tissue Eng., 2007, 13, 2733-2741. 212. J. Noordenbos, C. Doré and J. F. Hansbrough, J. Burn Care Res., 1999, 20, 275-281. 213. W. A. Marston, Expert Rev. Mol. Med., 2004, 1, 21-31. 214. V. Falanga and M. Sabolinski, Wound Repair Regen., 1999, 7, 201-207. 157 215. V. Falanga, D. Margolis, O. Alvarez, M. Auletta, F. Maggiacomo, M. Altman, J. Jensen, M. Sabolinski and J. Hardin-Young, Arch. Dermatol., 1998, 134, 293-300. 216. R. Guo, S. Xu, L. Ma, A. Huang and C. Gao, Biomaterials, 2011, 32, 1019-1031. 217. C. Philandrianos, L. Andrac-Meyer, S. Mordon, J.-M. Feuerstein, F. Sabatier, J. Veran, G. Magalon and D. Casanova, Burns, 2012, 38, 820-829. 218. C. M. Kelleher and J. P. Vacanti, J. R. Soc. Interface, 2010, rsif20100345. 219. G. D. Prestwich, J. Cell. Biochem., 2007, 101, 1370-1383. 220. S. Heydarkhan-Hagvall, C.-H. Choi, J. Dunn, S. Heydarkhan, K. Schenke-Layland, W. R. MacLellan and R. E. Beygui, CELL Commun. Adhes., 2007, 14, 181-194. 221. C. E. Ayres, B. S. Jha, S. A. Sell, G. L. Bowlin and D. G. Simpson, Wiley Interdiscip Rev Nanomed Nanobiotechnol., 2010, 2, 20-34. 222. R. Jaeger, M. M. Bergshoef, C. M. I. Batlle, H. Schönherr and G. Julius Vancso, Macromolecular symposia, 1998. 223. M. M. Hohman, M. Shin, G. Rutledge and M. P. Brenner, Phys. Fluids, 2001, 13, 2201-2220. 224. J. Venugopal, S. Low, A. T. Choon and S. Ramakrishna, J. Biomed. Mater. Res. Part B Appl. Biomater, 2008, 84, 34-48. 225. N. G. Rim, C. S. Shin and H. Shin, Biomed. Mater., 2013, 8, 014102. 226. Y. Yang, X. Li, M. Qi, S. Zhou and J. Weng, Eur. J. Pharm. Biopharm., 2008, 69, 106-116. 227. H. S. Yoo, T. G. Kim and T. G. Park, Adv. Drug Delivery Rev., 2009, 61, 10331042. 228. Y. Z. Zhang, X. Wang, Y. Feng, J. Li, C. T. Lim and S. Ramakrishna, Biomacromolecules, 2006, 7, 1049-1057. 229. S. Ahn, H. Yoon, G. Kim, Y. Kim, S. Lee and W. Chun, Tissue Eng. Part C Methods, 2009, 16, 813-820. 230. R. Parenteau-Bareil, R. Gauvin, S. Cliche, C. Gariépy, L. Germain and F. Berthod, Acta Biomater., 2011, 7, 3757-3765. 231. P. X. Ma and J.-W. Choi, Tissue Eng., 2001, 7, 23-33. 232. Y. Yang, T. Xia, W. Zhi, L. Wei, J. Weng, C. Zhang and X. Li, Biomaterials, 2011, 32, 4243-4254. 158 233. H. Chen, J. Huang, J. Yu, S. Liu and P. Gu, Int. J. Biol. Macromolec., 2011, 48, 1319. 234. Y. Zhou, D. Yang, X. Chen, Q. Xu, F. Lu and J. Nie, Biomacromolecules, 2007, 9, 349-354. 235. B. Dhandayuthapani, U. M. Krishnan and S. Sethuraman, J. Biomed. Mater. Res. Part B Appl. Biomater, 2010, 94B, 264-272. 236. B. Veleirinho, D. S. Coelho, P. F. Dias, M. Maraschin, R. M. Ribeiro-do-Valle and J. A. Lopes-da-Silva, Int. J. Biol. Macromolec., 2012, 51, 343-350. 237. H. M. Powell, K. L. McFarland, D. L. Butler, D. M. Supp and S. T. Boyce, Tissue Eng. Pt. A, 2009, 16, 1083-1092. 238. S. G. Kumbar, S. P. Nukavarapu, R. James, L. S. Nair and C. T. Laurencin, Biomaterials, 2008, 29, 4100-4107. 239. A. R. Chandrasekaran, J. Venugopal, S. Sundarrajan and S. Ramakrishna, Biomed. Mater., 2011, 6, 015001. 240. W. Cui, X. Zhu, Y. Yang, X. Li and Y. Jin, Mat. Sci. Eng. C, 2009, 29, 1869-1876. 241. E. J. Chong, T. T. Phan, I. J. Lim, Y. Z. Zhang, B. H. Bay, S. Ramakrishna and C. T. Lim, Acta Biomater., 2007, 3, 321-330. 242. G. Jin, M. P. Prabhakaran and S. Ramakrishna, Acta Biomater., 2011, 7, 3113-3122. 243. A. Martins, A. R. C. Duarte, S. Faria, A. P. Marques, R. L. Reis and N. M. Neves, Biomaterials, 2010, 31, 5875-5885. 244. H. Cao, X. Jiang, C. Chai and S. Y. Chew, J. Controlled Release, 2010, 144, 203212. 245. T. Liu, J. Xu, B. P. Chan and S. Y. Chew, J. Biomed. Mater. Res. A, 2012, 100, 236242. 246. H. Nie and C.-H. Wang, J. Controlled Release, 2007, 120, 111-121. 247. K. Numata and D. L. Kaplan, Adv. Drug Delivery Rev., 2010, 62, 1497-1508. 248. М. Ignatova, I. Rashkov and N. Manolova, Exp. Opin. Drug Deliv., 2013, 10, 469483. 249. X. Hu, S. Liu, G. Zhou, Y. Huang, Z. Xie and X. Jing, J. Controlled Release, 2014, 185, 12-21. 159 250. E.-R. Kenawy, G. L. Bowlin, K. Mansfield, J. Layman, D. G. Simpson, E. H. Sanders and G. E. Wnek, J. Controlled Release, 2002, 81, 57-64. 251. K. Shalumon, K. Anulekha, S. V. Nair, S. Nair, K. Chennazhi and R. Jayakumar, Int. J. Biol. Macromolec., 2011, 49, 247-254. 252. R. Dave, P. Jayaraj, P. K. Ajikumar, H. Joshi, T. Mathews and V. P. Venugopalan, J. Biomater. Sci., Polym. Ed., 2013, 24, 1305-1319. 253. K. Kim, Y. K. Luu, C. Chang, D. Fang, B. S. Hsiao, B. Chu and M. Hadjiargyrou, J. Controlled Release, 2004, 98, 47-56. 254. S. T. Yohe, Y. L. Colson and M. W. Grinstaff, J. Am. Chem. Soc., 2012, 134, 20162019. 255. S. T. Yohe, V. L. Herrera, Y. L. Colson and M. W. Grinstaff, J. Controlled Release, 2012, 162, 92-101. 256. H. Jiang, D. Fang, B. Hsiao, B. Chu and W. Chen, J. Biomater. Sci., Polym. Ed., 2004, 15, 279-296. 257. J. Zou, Y. Yang, Y. Liu, F. Chen and X. Li, Polymer, 2011, 52, 3357-3367. 258. P. Gupta, S. R. Trenor, T. E. Long and G. L. Wilkes, Macromolecules, 2004, 37, 9211-9218. 259. X. Xu, J.-F. Zhang and Y. Fan, Biomacromolecules, 2010, 11, 2283-2289. 260. Y. Jin, D. Yang, Y. Zhou, G. Ma and J. Nie, J. Appl. Polym. Sci., 2008, 109, 33373343. 261. V. Leszczak, L. W. Place, N. Franz, K. C. Popat and M. J. Kipper, ACS Appl. Mater. Interfaces, 2014. 262. D. Newton, R. Mahajan, C. Ayres, J. R. Bowman, G. L. Bowlin and D. G. Simpson, Acta Biomater., 2009, 5, 518-529. 263. K. S. Rho, L. Jeong, G. Lee, B.-M. Seo, Y. J. Park, S.-D. Hong, S. Roh, J. J. Cho, W. H. Park and B.-M. Min, Biomaterials, 2006, 27, 1452-1461. 264. S. A. Sell, P. S. Wolfe, K. Garg, J. M. McCool, I. A. Rodriguez and G. L. Bowlin, Polymers, 2010, 2, 522-553. 265. C. P. Barnes, C. W. Pemble IV, D. D. Brand, D. G. Simpson and G. L. Bowlin, Tissue Eng., 2007, 13, 1593-1605. 160 266. G. P. Huang, S. Shanmugasundaram, P. Masih, D. Pandya, S. Amara, G. Collins and T. L. Arinzeh, J. Biomed. Mater. Res. A, 2014. 267. Z. Meng, X. Xu, W. Zheng, H. Zhou, L. Li, Y. Zheng and X. Lou, Colloids Surf., B, 2011, 84, 97-102. 268. L. Cheng, X. Sun, C. Hu, R. Jin, B. Sun, Y. Shi, L. Zhang, W. Cui and Y. Zhang, Acta Biomater., 2013, 9, 9461-9473. 269. S. Y. Chew, J. Wen, E. K. Yim and K. W. Leong, Biomacromolecules, 2005, 6, 2017-2024. 270. P.-o. Rujitanaroj, Y.-C. Wang, J. Wang and S. Y. Chew, Biomaterials, 2011, 32, 5915-5923. 271. A. Saraf, L. S. Baggett, R. M. Raphael, F. K. Kasper and A. G. Mikos, J. Controlled Release, 2010, 143, 95-103. 272. A. Mickova, M. Buzgo, O. Benada, M. Rampichova, Z. Fisar, E. Filova, M. Tesarova, D. Lukas and E. Amler, Biomacromolecules, 2012, 13, 952-962. 273. C.-L. Zhang and S.-H. Yu, Chem. Soc. Rev., 2014, 43, 4423-4448. 274. S. L. Edwards, J. S. Church, J. A. Werkmeister and J. A. Ramshaw, Biomaterials, 2009, 30, 1725-1731. 275. R. Jayakumar, D. Menon, K. Manzoor, S. Nair and H. Tamura, Carbohydrate Polymers, 2010, 82, 227-232. 276. J. Wang, H.-B. Yao, D. He, C.-L. Zhang and S.-H. Yu, ACS Appl. Mater. Interfaces, 2012, 4, 1963-1971. 277. D. Li, J. T. McCann, M. Gratt and Y. Xia, Chem. Phys. Lett., 2004, 394, 387-391. 278. G.-M. Kim, A. Wutzler, H.-J. Radusch, G. H. Michler, P. Simon, R. A. Sperling and W. J. Parak, Chem. Mater., 2005, 17, 4949-4957. 279. P.-o. Rujitanaroj, N. Pimpha and P. Supaphol, Polymer, 2008, 49, 4723-4732. 280. E. Formo, E. Lee, D. Campbell and Y. Xia, Nano Lett., 2008, 8, 668-672. 281. Q. Shi, N. Vitchuli, J. Nowak, J. Noar, J. M. Caldwell, F. Breidt, M. Bourham, M. McCord and X. Zhang, J. Mater. Chem., 2011, 21, 10330-10335. 282. A. C. Patel, S. Li, C. Wang, W. Zhang and Y. Wei, Chem. Mater., 2007, 19, 12311238. 161 283. J. Bai, Y. Li, S. Yang, J. Du, S. Wang, C. Zhang, Q. Yang and X. Chen, Nanotechnology, 2007, 18, 305601. 284. M. Bognitzki, M. Becker, M. Graeser, W. Massa, J. H. Wendorff, A. Schaper, D. Weber, A. Beyer, A. Gölzhäuser and A. Greiner, Adv. Mater., 2006, 18, 2384-2386. 285. Y. Yu, L. Gu, C. Zhu, P. A. van Aken and J. Maier, J. Am. Chem. Soc., 2009, 131, 15984-15985. 286. R. Solasi, Y. Zou, X. Huang, K. Reifsnider and D. Condit, J. Power Sources, 2007, 167, 366-377. 287. A. Kusoglu, A. M. Karlsson, M. H. Santare, S. Cleghorn and W. B. Johnson, J. Power Sources, 2006, 161, 987-996. 288. X. Huang, R. Solasi, Y. Zou, M. Feshler, K. Reifsnider, D. Condit, S. Burlatsky and T. Madden, J. Polym. Sci. Part B Polym. Phys., 2006, 44, 2346-2357. 289. Y. K. Sung, B. W. Ahn and T. J. Kang, J. Magn. Magn. Mater., 2012, 324, 916-922. 290. F. Meng, Y. Zhan, Y. Lei, R. Zhao, J. Zhong and X. Liu, J. Appl. Polym. Sci., 2012, 123, 1732-1739. 291. N. Sharma, G. H. Jaffari, S. I. Shah and D. J. Pochan, Nanotechnology, 2010, 21, 085707. 292. A. Garreau, F. Massuyeau, S. Cordier, Y. Molard, E. Gautron, P. Bertoncini, E. Faulques, J. Wery, B. Humbert and A. Bulou, ACS nano, 2013, 7, 2977-2987. 293. K. Yin, L. Zhang, C. Lai, L. Zhong, S. Smith, H. Fong and Z. Zhu, J. Mater. Chem., 2011, 21, 444-448. 294. G. Dong, X. Xiao, Y. Chi, B. Qian, X. Liu, Z. Ma, S. Ye, E. Wu, H. Zeng, D. Chen and J. Qiu, J. Phys. Chem. C, 2009, 113, 9595-9600. 295. K. Friedemann, A. Turshatov, K. Landfester and D. Crespy, Langmuir, 2011, 27, 7132-7139. 296. Z. Hou, C. Li, P. Ma, G. Li, Z. Cheng, C. Peng, D. Yang, P. Yang and J. Lin, Adv. Funct. Mater., 2011, 21, 2356-2365. 297. S. Aloom, J. Polak and E. Lindenlaub, Nucleus, 1984, 4, 570-575. 298. J. W. Baish and R. K. Jain, Cancer res., 2000, 60, 3683-3688. 299. V. Tsyalkovsky, V. Klep, K. Ramaratnam, R. Lupitskyy, S. Minko and I. Luzinov, Chem. Mater., 2007, 20, 317-325. 162 300. C. Xia, J. Zuo, C. Wang and Y. Wang, Orthopedics, 2012, 35, 572. 301. P. A. Dennis and D. B. Rifkin, Proc. Natl. Acad. Sci., 1991, 88, 580-584. 302. M. Prabaharan, R. Jayakumar and S. Nair, in Biomedical applications of polymeric nanofibers, Springer, 2012, pp. 241-262. 303. X. Zong, K. Kim, D. Fang, S. Ran, B. S. Hsiao and B. Chu, Polymer, 2002, 43, 4403-4412. 304. S. Lee and S. Kay Obendorf, J. Appl. Polym. Sci., 2006, 102, 3430-3437. 305. S. Hoogendoorn, G. H. van Puijvelde, J. Kuiper, G. A. van der Marel and H. S. Overkleeft, Angew. Chem. Int. Ed., 2014, 126, 11155-11158. 306. K. McGourty, T. L. Thurston, S. A. Matthews, L. Pinaud, L. J. Mota and D. W. Holden, Science, 2012, 338, 963-967. 307. P. Saftig and J. Klumperman, Nat. Rev. Mol. Cell Biol., 2009, 10, 623-635. 308. T. Braulke and J. S. Bonifacino, BBA-Mol. Cell Res., 2009, 1793, 605-614. 309. M. W. J. Ferguson, J Interferon Res., 1994, 14, 303-304. 310. C. A. Gabel, C. E. Costello, V. N. Reinhold, L. Kurz and S. Kornfeld, J. Biol. Chem., 1984, 259, 13762-13769. 311. H. H. Freeze, Arch Biochem Biophys, 1985, 243, 690-693. 312. D. A. Smith, C. Allerton, H. Kubinyi, H. Walker and D. K. Walker, Pharmacokinetics and metabolism in drug design, John Wiley & Sons, 2012. 313. U. Pradere, E. C. Garnier-Amblard, S. J. Coats, F. Amblard and R. F. Schinazi, Chem. Rev., 2014. 314. K. Hofman, B. Hall, H. Cleaver and S. Marshall, Anal. Biochem., 2011, 417, 289291. 315. C. Z. C. Chen, Y. X. Peng, Z. B. Wang, P. V. Fish, J. L. Kaar, R. R. Koepsel, A. J. Russell, R. R. Lareu and M. Raghunath, Br. J. Pharmacol., 2009, 158, 1196-1209. 316. S. E. D'Souza, M. H. Ginsberg and E. F. Plow, Trends Biochem. Sci., 1991, 16, 246250. 317. S. L. Bellis, Biomaterials, 2011, 32, 4205-4210. 318. J. S. Munger and D. Sheppard, Cold Spring Harb. Perspect. Biol., 2011, 3, a005017. 319. D. T. Worster, T. Schmelzle, N. L. Solimini, E. S. Lightcap, B. Millard, G. B. Mills, J. S. Brugge and J. G. Albeck, Sci. Signal., 2012, 5, ra19-ra19. 163 320. M. Hajimiri, S. Shahverdi, G. Kamalinia and R. Dinarvand, J. Biomed. Mater. Res. A, 2014, n/a-n/a. 321. T. N. Demidova-Rice, M. R. Hamblin and I. M. Herman, Adv. Skin Wound Care, 2012, 25, 349. 322. S. Benzaria, H. Pélicano, R. Johnson, G. Maury, J.-L. Imbach, A.-M. Aubertin, G. Obert and G. Gosselin, J. Med. Chem., 1996, 39, 4958-4965. 164