PDF hosted at the Radboud Repository of the Radboud University Nijmegen The following full text is a publisher's version. For additional information about this publication click this link. http://hdl.handle.net/2066/130279 Please be advised that this information was generated on 2017-06-16 and may be subject to change. Bone Regeneration: From Intramembranous to Endochondral Pathway Wanxun Yang Colofon Thesis Radboud University Medical Center, with summary in English and Dutch. Bone Regeneration: From Intramembranous to Endochondral Pathway The research presented in this thesis was financially supported by China Scholarship Council (No. 2010627030) and The Royal Netherlands Academy of Arts and Sciences (KNAW). Publication of this thesis was financially supported by Radboud University Medical Center (Radboud UMC), Radboud Institute for Molecular Life Sciences (RIMLS), Netherlands Society for Biomaterials and Tissue Engineering (NBTE) and Corbion Purac Biomaterials. Layout: Arnold Nijhuis & Wanxun Yang Cover design: Wanxun Yang & Proefschriftmaken.nl || Uitgeverij BOXPress Printed by: Proefschriftmaken.nl || Uitgeverij BOXPress Published by: Uitgeverij BOXPress, ‘s‐Hertogenbosch ISBN: 978-90-8891-962-6 Copyright © Wanxun Yang, 2014 All rights reserved. Bone Regeneration: From Intramembranous to Endochondral Pathway Proefschrift ter verkrijging van de graad van doctor aan de Radboud Universiteit Nijmegen op gezag van de rector magnificus prof. dr. Th.L.M. Engelen, volgens besluit van het college van decanen in het openbaar te verdedigen op dinsdag 21 oktober 2014 om 14.30 uur precies door Wanxun Yang Geboren op 15 juni 1985 te Bengbu, China Promotor Prof. dr. J.A. Jansen Copromotoren Dr. F. Yang Dr. ing. S.K. Both (Universiteit Twente) Manuscriptcommissie Prof. dr. P. Buma Prof. dr. J. Klein-Nulend (Vrije Universiteit) Prof. dr. M. Karperien (Universiteit Twente) Paranimfen Eva Urquía Edreira Alexey Klymov Bone Regeneration: From Intramembranous to Endochondral Pathway Doctoral Thesis to obtain the degree of doctor from Radboud University Nijmegen on the authority of the Rector Magnificus prof. dr. Th.L.M. Engelen, according to the decision of the Council of Deans to be defended in public on Tuesday, October 21, 2014 at 14.30 hours by Wanxun Yang born on June 15, 1985 in Bengbu, China Supervisor Prof. dr. J.A. Jansen Co‐supervisors Dr. F. Yang Dr. ing. S.K. Both (University of Twente) Doctoral Thesis Committee Prof. dr. P. Buma Prof. dr. J. Klein-Nulend (VU University) Prof. dr. M. Karperien (University of Twente) Ushers Eva Urquía Edreira Alexey Klymov “Don’t be too timid and squeamish about your actions. All life is an experiment.” - Ralph Waldo Emerson To my beloved family 谨以此书献给我的父母和家人 Contents Chapter 1 | Page 12 General introduction Chapter 2 | Page 20 Combining osteochondral stem cells and biodegradable materials for bone regeneration Chapter 3 | Page 40 Biological evaluation of porous aliphatic polyurethane/hydroxyapatite composite scaffolds for bone tissue engineering Chapter 4 | Page 60 In vivo bone generation via the endochondral pathway on three-dimensional electrospun fibers Chapter 5 | Page 80 Performance of different three-dimensional scaffolds for in vivo endochondral bone generation Chapter 6 | Page 106 Effects of in vitro chondrogenic priming time of bone marrow derived mesenchymal stem cells on in vivo endochondral bone formation Chapter 7 | Page 130 Roles of cartilage templates in mediating mesenchymal stem cell behavior in endochondral bone tissue engineering Chapter 8 | Page 148 Summary, closing remarks and future perspectives Chapter 9 | Page 158 Samenvatting, slotopmerkingen en toekomstperspectieven Acknowledgements | Page 168 List of publications | Page 176 Curriculum Vitae | Page 178 Chapter 1 General Introduction General Introduction 1. Background Bone is a dynamic and vascularized tissue responsible for providing structural support to the body and protection of the internal organs. In addition, it plays a key role in supporting muscle contraction, blood production and the storage of mineral components. Clinically, the occurrence of bone defects is common and can be caused by a variety of reasons, including congenital deformity, trauma/ injury, infection and tumorectomy. Although bone possesses a remarkable ability to heal and remodel, in many clinical cases the regeneration of bone defects exceeds its self-healing potential. It has been reported that bone is the second most transplanted tissue after blood with over 2.2 million bone grafting procedure being performed annually worldwide [1]. In addition, the treatment for fractures and other musculoskeletal injuries will continue to boost as a result of increasing life expectancy and dynamism in daily activities. Current clinical treatments for bone defects involve the transplantation of autograft (from the patient) or allograft (from the donor), which can be harvested from the iliac crest, fibula, scapula or radius [2]. So far, this approach has been established as the “gold standard” for bone reconstructive surgery. However, the inherent drawbacks of this approach, such as limited source of available tissue, donor-site morbidity, risk of rejection or disease transmission (when using allograft), and relevant surgical complications, hinder its applications in the clinical practice, especially for the treatment of large bone defects [3, 4]. This has strongly urged the clinicians and researches to explore alternative treatment options. 2. Bone tissue engineering Since the concept of tissue engineering has been introduced by Langer and Vacanti [5], it has emerged as a promising approach for regenerating bone tissue. The essential rationale behind this is to utilize the body’s own capacity for bone healing in conjunction with engineering principles. Three important elements are involved in bone tissue engineering, i.e. scaffolds, cells and biomolecules. A conventional tissue engineering strategy starts with loading progenitor cells onto proper designed biomaterials to induce cell differentiation in a specific pathway, followed by transplantation of this cell-scaffold complex into the defect site to regenerate bone tissue. In this procedure, cells are the main source to form the new bone tissue, among which the progenitor/stem cells isolated from bone marrow have been recognized as the most efficient cell source [6]. 15 1 Chapter 1 3. Intramembranous and endochondral approach for bone tissue engineering During osteogenesis, bone is essentially formed through two different mechanisms: intramembranous ossification and endochondral ossification. Accordingly, this has inspired researchers to implement bone tissue engineering through two specific pathways of cell differentiation. In the first approach, bone is directly from osteoprogenitor cells in a manner akin to intramembranous ossification [7]. This method is the most commonly used strategy in bone tissue engineering for years, a standard procedure of which involves the differentiation of osteoprogenitor cells into osteoblasts. The second approach resembles endochondral ossification, whereby a cartilage intermediate is created and used as an osteoinductive template to transform into bone in vivo [8]. Such endochondral approach has been recently proposed and shows superiority to generate vascularized bone tissue, as cartilage is hypoxia tolerant and chondrocytes can secrete angiogenic factors to promote vascular invasion into the engineered constructs [9, 10]. 4. Scaffolds for bone tissue engineering Besides the cells, the choice of biomaterial as scaffold is also critical to the success of tissue engineering. It provides a three-dimensional environment for the cells and plays an imperative role in regulating the cellular behaviors, i.e. proliferation and differentiation [11]. The shape and architecture of the scaffold define the ultimate structure of newly formed tissue [12]. Since the introduction of the tissue engineering concept, a variety of biomaterials of natural derivation or synthetic origin has been suggested as potential scaffold candidates. Among them, ceramic-based materials, such as hydroxyapatite (HA) and calcium phosphate (CaP), and polymer-based materials, such as collagen, polycaprolactone (PCL), polyurethanes (PUs), polyglycolide (PGA), polylactide (PLA) and their copolymers, are currently the most widely used. Along with research on biomaterials, techniques to manufacture scaffolds have also been intensively studied, such as gas foaming and electrospinning. In general, scaffolds for application in bone tissue engineering have to meet the following requirements, 1) provide a biologically compatible and porous framework to facilitate nutrients supply and waste removal, and 2) possess sufficient mechanical strength to function from the time of cell seeding up to the point when the host tissue begins to remodel the seeded construct after implantation [11]. More importantly, the cellular growth and osteogenic capacity of the seeded cells should be supported/promoted by the scaffolds to form the ultimate bone tissue. 16 General Introduction When the endochondral approach is applied, the scaffolds should also be able to support cells in differentiating into the chondrogenic lineage, as the creation of an in vitro cartilage template is necessary. 5. Objectives of this thesis With the potential to overcome the drawbacks of the current autografts and allografts, bone tissue engineering undoubtedly has been suggested as a promising technique for reconstructing bone defects. However, until recently, the clinical translation of this strategy is still unsatisfactory. Lack of timely and sufficient vascular supply, resulting in immediate cell death after implantation, is generally thought to be the cause of failure of the cell-scaffold constructs in patients [4]. To overcome this issue, a potential strategy is to differentiate cells in a metabolically less-demanding and angiogenic pathway, i.e. along the endochondral ossification pathway, as this seems to hold promise for achieving better cell survival and vascularization of the implanted constructs in vivo. Nevertheless, compared to the available knowledge on bone regeneration via the intramembranous pathway, such an endochondral approach is still in its infancy. Therefore, the overall objective of thesis was to contribute to a better understanding of endochondral bone tissue engineering. In particular, the studies focused on 1) exploiting new porous scaffolds for bone regeneration applications, 2) establishing a cell-scaffold model for studying endochondral bone formation, 3) testing the applicability of endochondral approach with different scaffold materials, 4) investigating the optimal in vitro priming time for the creation of cartilage templates, and 5) exploring the underlying mechanism of the transformation from cartilage into bone. More specifically, the following research questions were addressed: 1. What is the current state of the art of combining stem cells and biodegradable materials for bone regeneration via the endochondral pathway? (Chapter 2) 2. Is the porous aliphatic polyurethane/hydroxyapatite composite a suitable scaffold for bone tissue engineering applications? (Chapter 3) 3. Is it possible to fabricate a 3D scaffold by electrospinning technique and utilize it as cell carrier to enable effective in vivo endochondral bone generation? (Chapter 4) 4. Can in vivo endochondral bone generation be achieved in various scaffold materials and how is this process influenced by the scaffold properties? (Chapter 5 and Chapter 6) 5. What is the effect of in vitro chondrogenic priming time of MSCs on in vivo endochondral bone formation? (Chapter 6) 17 1 Chapter 1 6. What is the role of cartilage templates in mediating MSC behavior in the process of endochondral bone formation? (Chapter 7) References 1. Giannoudis, P.V., Dinopoulos, H., and Tsiridis, E., Bone substitutes: an update. Injury, 2005. 36 Suppl 3: p. S20-7. 2. Buck, D.W., 2nd and Dumanian, G.A., Bone biology and physiology: Part II. Clinical correlates. Plast Reconstr Surg, 2012. 129(6): p. 950e-956e. 3. Dawson, J.I. and Oreffo, R.O., Bridging the regeneration gap: stem cells, biomaterials and clinical translation in bone tissue engineering. Arch Biochem Biophys, 2008. 473(2): p. 124-31. 4. Meijer, G.J., de Bruijn, J.D., Koole, R., and van Blitterswijk, C.A., Cell-based bone tissue engineering. PLoS Med, 2007. 4(2): p. e9. 5. Langer, R. and Vacanti, J.P., Tissue engineering. Science, 1993. 260(5110): p. 920-6. 6. Bianco, P. and Robey, P.G., Stem cells in tissue engineering. Nature, 2001. 414(6859): p. 118-21. 7. Caplan, A.I., Review: mesenchymal stem cells: cell-based reconstructive therapy in orthopedics. Tissue Eng, 2005. 11(7-8): p. 1198-211. 8. Gawlitta, D., Farrell, E., Malda, J., Creemers, L.B., Alblas, J., and Dhert, W.J., Modulating endochondral ossification of multipotent stromal cells for bone regeneration. Tissue Eng Part B Rev. 16(4): p. 385-95. 9. Farrell, E., van der Jagt, O.P., Koevoet, W., Kops, N., van Manen, C.J., Hellingman, C.A., Jahr, H., O’Brien, F.J., Verhaar, J.A., Weinans, H., and van Osch, G.J., Chondrogenic priming of human bone marrow stromal cells: a better route to bone repair? Tissue Eng Part C Methods, 2009. 15(2): p. 285-95. 10. Scotti, C., Tonnarelli, B., Papadimitropoulos, A., Scherberich, A., Schaeren, S., Schauerte, A., Lopez-Rios, J., Zeller, R., Barbero, A., and Martin, I., Recapitulation of endochondral bone formation using human adult mesenchymal stem cells as a paradigm for developmental engineering. Proc Natl Acad Sci U S A, 2010. 107(16): p. 7251-6. 11. Hutmacher, D.W., Schantz, J.T., Lam, C.X., Tan, K.C., and Lim, T.C., State of the art and future directions of scaffold-based bone engineering from a biomaterials perspective. J Tissue Eng Regen Med, 2007. 1(4): p. 245-60. 12. Scaglione, S., Giannoni, P., Bianchini, P., Sandri, M., Marotta, R., Firpo, G., Valbusa, U., Tampieri, A., Diaspro, A., Bianco, P., and Quarto, R., Order versus Disorder: in vivo bone formation within osteoconductive scaffolds. Sci Rep, 2012. 2: p. 274. 18 General Introduction 19 1 Chapter 2 Combining osteochondral stem cells and biodegradable materials for bone regeneration Wanxun Yang*, Matilde Bongio*, Fang Yang, Sanne K. Both, Sander C.G. Leeuwenburgh, Jeroen J.J.P. van den Beucken and John A. Jansen *These authors contribute equally. Combining Osteochondral Stem Cells and Biomaterials for Bone Regeneration 1. Introduction Musculoskeletal malconditions (e.g. arthritis, back pain, spinal disorders, osteoporosis, bone fractures, trauma, and congenital problems) affect hundreds of millions of people across the world. According to ‘The Burden of Musculoskeletal Diseases in the United States’ (BMUS), over 130 million patient visits to healthcare providers occur annually, and the economic impact of these conditions is staggering: in 2004, the annual costs for bone and joint health has been estimated to be $849 billion dollars, and the burden of musculoskeletal conditions is expected to escalate in the next 10-20 years as the population ages and lifestyles become more active (http://www.boneandjointburden.org). The most favorable choice for treatment of bone defects is the graft of the patient’s own healthy bone (so-called autograft) harvested from a donor site, usually iliac crest; or, alternatively, from a cadaveric bone typically sourced from a bone bank (allograft). In addition to the respective benefits of these approaches, several drawbacks limit their clinical applications. Autografts reduce the risk of rejection but require an additional surgical site, thus increasing post-operative pain and complications for the patient. Allograft eliminates the risk of donor site morbidity and is hence associated with less pain and discomfort for the recipient; however, rejection or disease transmission cannot be ruled out. In view of these drawbacks associated with autografts and allografts, the multidisciplinary field of regenerative medicine (RM) holds the promise of creating appealing new possibilities regarding the quality and duration of life. Currently, RM-based treatment strategies for critical-size bone defects (i.e. defects that will not heal spontaneously) propose the design of biomaterials either alone or in combination with stem cells and signaling molecules, which resemble the three dimensional in vivo tissue arrangements and promote restoration of normal tissue functions following grafting in the body [1]. To that end, a deep knowledge of bone biology will lead to better understanding of the mechanisms of interaction between cells and scaffolds, including cell adhesion, proliferation, and differentiation, which are all important steps for the success of a man-made bone graft. Up to now, thousands of scaffolds with variable properties and chemical composition, including ceramics, natural and synthetic polymers, as well as cell culture strategies have been proposed. Nevertheless, as for most of these approaches, the developmental stage is in its infancy and the challenge is to obtain proof of clinical efficacy in order to proceed towards routine clinical implementation. The present chapter will firstly describe bone tissue and its developmental mechanisms, in order to reach fundamental understanding of the processes involved in RM-based approaches for bone regeneration utilizing biomaterials 23 2 Chapter 2 and cell culture strategies. Subsequently, current results in basic research on biomaterials for bone repair, with emphasis on the endochondral ossification approach, will be discussed. 2. Bone tissue Bone is a dynamic and vascularized tissue with a remarkable ability to heal and remodel, and it is responsible for providing structural support to the body and protection of the internal organs. In addition, bone plays a key role in supporting muscle contraction, blood production and the storage of mineral components [2]. Bone is not a uniformly solid structure, but is rather composed of two distinct parts, namely compact and trabecular bone. Compact bone (also called cortical bone) forms the hard outer component of the bones, whereas trabecular bone (also called cancellous bone) fills the inner side. Compact bone contributes approximately 80% of the weight of a human skeleton with a very dense and stiff structure and is usually located in the shaft of the long bones and in the peripheral layer of flat bones. In contrast, trabecular bone has a lower density, larger surface area, and is composed of a highly porous network that creates space for bone marrow and blood vessels. From a material perspective, bone is composed of an inorganic and organic component. The inorganic matrix represents ~70% of bone mass and is mainly composed of hydroxyapatite, which provides stiffness and strength. The organic matrix, which represents ~30% of bone mass, provides flexibility and tensile strength, and is composed of proteins and cells. Collagen type I is the main component of the organic matrix; whereas other proteins, such as osteocalcin, bone sialoprotein and osteonectin, play important signaling functions and are involved in mineralization process. The three types of cells that populate bone tissue are osteoblasts, osteocytes and osteoclasts. Osteoblasts are bone-forming cells that originate from osteoprogenitor cells located in the deeper layer of periosteum and bone marrow, which produce an osteoid matrix that successively mineralizes into bone. Although the term osteoblast implies an immature cell type, osteoblasts are the mature cells of bone that generate bone tissue. At a later stage, osteoblasts can either become bone lining cells or differentiate into osteocytes. Bone lining cells are inactive osteoblasts that reside along the quiescent bone surface forming a protective layer. Although their function is not well understood, they may play a role in the activation of bone remodeling [3]. Osteocytes are embedded within bone matrix in so-called bone lacunae. These mature bone cells are involved in different functions, including matrix maintenance, calcium homeostasis and regulation of bone response to 24 Combining Osteochondral Stem Cells and Biomaterials for Bone Regeneration mechanical stimuli. Osteoclasts are responsible for bone resorption by (locally) decreasing the pH and activating lytic enzymes, which in turn break down the mineralized matrix. 2.1 Osteogenesis During osteogenesis, two different processes can result in the formation of normal, healthy bone tissue, i.e. intramembranous ossification and endochondral ossification (Figure 1). Intramembranous ossification Intramembranous ossification occurs in a vascularized environment and involves the direct differentiation of mesenchymal stem cells (MSCs) into osteoblasts. Initially, MSCs are widely dispersed within an extracellular matrix that is devoid of collagen. Groups of adjacent MSCs proliferate and form a dense nodule of cell aggregation, which undergoes morphological changes toward the osteoblastic phenotype. Subsequently, cells start to produce new organic matrix (i.e. osteoid), mainly consisting of type I collagen fibrils, and eventually develop into mature osteoblasts. Some osteoblasts align along the periphery of the nodule and keep forming osteoid, whereas others are incorporated into the nodule and become osteocytes. At this point, the osteoid becomes mineralized and forms the rudimentary bone tissue. Endochondral ossification Endochondral ossification is the type of osteogenesis that accounts for the development of most bones of the vertebrate skeleton (e.g. long bones) and is the most commonly occurring process in fracture healing. As such, endochondral ossification takes place in a poorly or non-vascularized environment, in which loosely connected MSCs condense at the sites pre-determined for bone formation. The increased cell-cell contacts and low oxygen condition (i.e. hypoxia) trigger the differentiation of MSCs into the chondrogenic lineage. The cartilage template takes shape through the deposition of abundant cartilaginous matrix, and the chondrocytes gradually differentiate until reaching a hypertrophic stage, during which the chondrocytes mineralize and eventually become apoptotic. Simultaneously, these chondrogenic cells release vascular endothelial growth factor (VEGF), which induces blood vessel formation [4]. At this point, osteoclasts and osteoblasts, which receive nutrition and oxygen supply from the blood vessels, reach the mineralized matrix and promote the transformation of the cartilage into bone. 25 2 Chapter 2 Figure 1. Schematic representation of the two types of osteogenesis. 1) Intramenbranous ossification involves bone formation through the direct osteoblastic differentiation; 2) endochondral ossification generates bone via transition from an avascular hypertrophic cartilage to a vascularized and mineralized matrix. 3. Basic tools for RM purposes - stem cells and biomaterial scaffolds 3.1 Stem cells Stem cells are undifferentiated cells with a high proliferation capability, which have the capacity of self-renewal or differentiate into specialized cell types of our body [5]. The central focus of RM are human stem cells, for which two main types of stem cells exist based on their origin: embryonic stem cells (ESCs) and adult stem cells. ESCs are derived from the inner cell mass of the blastocyst (i.e. an early-stage embryo) and are able to differentiate into all cell types of the adult organism. The multi-differentiation potential of ESCs has been reported by several researchers [6, 7], and has attracted wide interest for many medical applications, including regeneration of bone tissue. Indeed, ESCs have demonstrated the capacity to differentiate into osteoblasts when cultured in the presence of osteogenic differentiation-inducing factors, such as dexamethasone [8]. However, many issues regarding the clinical use of ESCs need to be addressed, including the risk of tumorogenesis, the labor-intensive procedures, and important ethical 26 Combining Osteochondral Stem Cells and Biomaterials for Bone Regeneration considerations [9]. Adult stem cells, also called somatic stem cells, are present in the body at many sites and are responsible for growth and maintenance of normal tissues, as well as restoration of injured tissues. Initially, adult stem cells were discovered in the bone marrow and, successively, in many other different tissues, including heart, liver, pancreas, brain, epidermis, dental pulp and spinal cord. In the last 30 years, substantial efforts have been devoted to studies of adult stem cells for a broad range of applications for tissue regeneration. Among the different sources for adult stem cells, mesenchymal stem cells (MSCs) are of great interests in RM applications. MSCs have been initially identified in bone marrow as nonhematopoietic stem cells that can differentiate into tissues of mesodermal origin, such as adipocytes, osteoblasts, chondrocytes, tenocytes, skeletal myocytes and visceral stromal cells [10]. Friedenstein and co-workers were the first to reveal the bone-forming capacity of MSCs [11]. 3.2 Biomaterial scaffolds As anticipated earlier in the introduction, a promising strategy to optimize and simplify the treatment of muscoloskeletal disorders is represented by bone grafts from man-made source. Several decades of progress in techniques and understanding of the mechanisms involved in the interaction between synthetic materials and living surrounding tissues resulted in the establishment of ‘biomaterials’ as a scientific discipline, and a rapid progression in the design of increasingly sophisticated materials that provide a template for tissue regeneration. As a result of this evolution, the definition of “biomaterial” has constantly changed over the years. Currently, the most appropriate and used definition depicts biomaterial as “a substance that has been engineered to take a form which, alone or as part of a complex system, is used to direct, by control of interactions with components of living systems, the course of any therapeutic or diagnostic procedure, in human or veterinary medicine” [12]. The history of biomaterials for bone repair is marked by three generations, namely, from tolerant and bioinert (the first generation), to bioactive and osteoconductive (the second generation), and to osteoinductive biomaterials (the third generation) that induce neighboring tissue to become bone [13]. Primordial biomaterials were meant to be tolerated by the body and to passively replace missing or damaged tissue. The difficult adaptation of the human body to these foreign materials raised the need to create improved artificial substitutes. These scaffolds reproduce the complex three dimensional environment to which cells are exposed to and actively participate in the dialog with the natural environment (i.e. bioactive). To this end, bone healing requires scaffolds that provide a porous and biologically compatible framework onto which (precursors of) bone-forming cells migrate 27 2 Chapter 2 (i.e. osteoconduction) and form new bone tissue, or even direct differentiation of stem cells into mature osteogenic cells (i.e. osteoinduction) [14]. Last, but not least, biomaterials should degrade over time in concert with the growth of new bone, and the degradation products should be non-toxic and completely eliminated from the body. Table 1 lists the most important properties and respective definitions for appropriate bone graft materials. Table 1. Properties and definitions for a bone graft material Properties Biocompatibility Bioactivity Definitions Behavior of a biomaterial and its byproducts to elicit little or no immune response into the body Ability of a biomaterial to interact with the natural environment of the body and determine a biological effect The full process of obtaining firm anchoring and incorporation of a Osteointegration biomaterial into living bone without fibrous tissue formation at the interface Osteoconductivity The ability of a biomaterial to stimulate the migration of adjacent boneforming cells in order to induce the growth of bone tissue The ability of a biomaterial to stimulate the migration of undifferentiated Osteoinductivity cells and induce their differentiation into active osteoblasts, in order to promote de novo bone formation Biodegradability The ability of a biomaterial to degrade in concert with the natural bone remodeling process 3.2.1 Ceramic-based materials In the past three decades, ceramics, such as hydroxyapatite (HA), tricalcium phosphate (TCP) and bioglass, were the main scaffolds studied for bone regeneration, being similar in their chemical composition and crystallographic structure to the mineral component of bone tissue. These characteristics govern the bioactive and osteoconductive behavior, which in turn favor the formation of new bone; although the brittleness and weak mechanical properties represent the major drawbacks that still require optimization [15]. 3.2.2 Polymer-based materials An alternative approach in RM is to engineer new bone tissues by using selective cell transplantation on polymer scaffolds [16]. Some of the most commonly explored polymers as scaffold materials include poly(lactic acid) (PLA), poly(lactic-co-glycolic 28 Combining Osteochondral Stem Cells and Biomaterials for Bone Regeneration acid) (PLGA), poly(2-hydroxyethyl methacrylate) (pHEMA), poly(ε-caprolactone) (PCL), poly(ethylene glycol) (PEG), polyglycolic acid (PGA), polyurethanes and composites [17]. Polymers exhibit biocompatibility, degradability and have good processability. These properties confer to polymers a great potential to serve as artificial ECM, thus mimicking and recreating in some extent the structural organization of bone. 4. Cell culture strategies for bone tissue regeneration Initial in vitro culture set-up, including selection of cell types (mono- or coculture), growth factors, medium composition and timing are pivotal to determine the possible benefits of well-timed implantation [4]. Figure 2 depicts all the steps required for the synthesis of a cell-scaffold construct for bone tissue regeneration, which include (1) cell isolation and expansion in complete medium, (2) cell seeding and cell maturation (via intramembranous or endochondral ossification), either prior to or after loading into 3D scaffolds, and (3) in vivo grafting. 4.1 Isolation and expansion As described above, natural bone regeneration is a complex process, during which multiple cell types, dispersed in a three dimensional environment, are spatially and temporally regulated by several environmental stimuli. In order to re-create the natural situation, millions of cells are necessary for incorporation into an artificial matrix. From a clinical point of view, it is possible to isolate osteoblasts from biopsies taken from the patients, however, this procedure is time consuming and allows to obtain only a limited number of cells. Therefore, stem cells represent a favorable alternative. Stem cells, derived from bone marrow or other tissues, are usually expanded with complete medium in culture plates or flasks, where they adhere, start proliferating and form colony-forming cell clusters. Before these cells become completely confluent, they are passaged and expanded in multiple flasks, thus becoming a more and more homogenous adherent cell population that may proliferate without differentiating up to several generations [18]. 4.2 Seeding and differentiation Intramembranous ossification pathway Except for a few reports, regenerative medicine efforts in the field of bone tissue repair have focused on generating bone directly from osteoprogenitor cells in a manner akin to intramembranous ossification in vivo. The standard procedure of this strategy involves stem cell differentiation into osteoblasts. To that end, MSCs are exposed to specific culture supplements, such as dexamethasone, ascorbic 29 2 Chapter 2 acid, and beta glycerolphosphate [19]. Cells can either be differentiated in flasks (2D-environment) and loaded into scaffolds just prior to grafting in vivo, or be loaded into scaffolds when they are still uncommitted and undergo maturation in a 3D-environment. Although the second approach is more appealing, as it resembles the in vivo situation, its success has not been satisfactory to date. A major issue with this approach is the limited long-term cell survival and necrosis, arising from lack of nutrient delivery to and waste removal from the center of tissue-engineered constructs, and lack of blood supply [20]. First, blood vessels formation is not an instantaneous process and usually occurs too slowly for cells to survive without nutrients. Second, culturing a cell-laden scaffold in vitro for several weeks can lead to the formation of an extensive extracellular matrix, which can prevent or seriously hamper blood vessel infiltration in vivo by sealing the pores of a scaffold [21]. Endochondral ossification pathway Because chondrocytes can survive in a hypoxic environment and they produce anabolic and catabolic factors important for the conversion of avascular tissue into vascularized tissue, it seems straightforward to exploit the chondrocytemediated route (i.e. endochondral) for ossification in RM-based approaches, rather than intramembranous ossification [4]. Either human ESCs and MSCs are able to form bone via the endochondral pathway, i.e. differentiate to chondrocytes in vitro and turn into bone when implanted in vivo [20, 22]. Chondrogenic priming of mesenchymal stem cells can be obtained by culturing cells in differentiation medium containing a mixture of specific supplements, such as dexamethasone, ascorbic acid, insulin, transferrin, selenious acid, sodium pyruvate, transforming growth factor-beta (TGF-β), and a phosphate donor [23]. Three-dimensional conformation of the cells in aggregates with high cell density and cell-cell interaction play an important role in the mechanism of cartilage formation. The main idea would be to induce the secretion of the angiogenic factors from the hypertrophic chondrocytes or in vitro pre-vascularization of the templates in order to facilitate blood vessel formation when implanted in vivo. However, no in vitro system is known to support vascularization necessary for the formation of bone following the endochondral ossification pathway. Different culture systems can offer a solution for the development of in vitro tissue engineering approaches for the creation of vascularized bone in vivo. Among these strategies, the use of bioreactors [24] and an environment with low oxygen tension have demonstrated to support growth and differentiation of cartilage that forms bone [25]. 30 Combining Osteochondral Stem Cells and Biomaterials for Bone Regeneration 2 Figure 2. Illustration of all the fundamental steps required for the synthesis and application of a cell-material system for bone tissue regeneration. (1) cell isolation and expansion, (2) cell seeding and differentiation, and (3) in vivo grafting. 31 Chapter 2 4.3 In vivo grafting Osteo- or chondroprogenitor cells, previously differentiated from stem cells under prolonged conventional culture conditions, are implanted in vivo in conjunction with natural or synthetic scaffolds. Osteoprogenitor cells differentiate directly into osteoblasts that will deposit new bone matrix, while chondrogenic-committed cells fulfill hypertrophy and further deposit cartilaginous extracellular matrix that becomes calcified, and eventually undergo apoptosis. Meanwhile, blood vessels and osteoblasts differentiated from the surrounding stem/progenitor cells, invade the hypertrophic cartilage and give rise to bone [26]. 5. Preliminary investigations on cell-biomaterial constructs for endochondral bone formation Regarding endochondral ossification, it is a very unique transition process, which includes two distinguished stages: chondrogenesis and osteogenesis. Cartilage matrix consists of highly hydrated proteoglycan embedded into a type II collagen network. On the other hand, bone tissue is more rigid than cartilage and impregnated mainly with hydroxyapatite (HA) and type I collagen. Regarding the endochondral bone tissue engineering, the scaffolds should be able to support cells to form transient cartilage-like tissue in vitro and also promote the calcification of such cartilage template into bone after in vivo implantation. The following sections describe the cell-scaffold systems that have been explored so far for the generation of bone tissue via endochondral ossification approach. 5.1 Calcium phosphate (CaP) ceramics Before being used as scaffold materials, CaP ceramics have been widely applied as bone substitutes and implant coatings because of their resemblance to the mineral composition of the natural bone. Commercially available CaP-based biomaterials include HA, β-tricalcium phosphate (β-TCP), and biphasic CaP (BCP, containing HA and β-TCP). The CaP ceramics are generally considered to be bioactive and osteoconductive, which posses the ability to guide formation of the new bone tissue tightly along their surfaces. This has made them the superior scaffold candidates for bone tissue engineering. However, the development of three-dimensional in vitro models of bone using bioceramics remains an important challenge that is still to be overcome. CaP ceramics are reactive and their reactivity depends on their characteristics (such as composition, dissolution, sintering temperature, microstructure). When ceramics are cultured in vitro, calcium and phosphates are released in the medium and, as a result, can hamper cellular function [27]. Since testing the in vitro cellular activity is a pertinent step, and scaffolds and cells are in 32 Combining Osteochondral Stem Cells and Biomaterials for Bone Regeneration contact for a certain period in vitro before being implanted, a number of studies have been focusing on solving the difficulties of removal of excessive ions, as well as oxygen and nutrients supply to cells within ceramic scaffolds [28]. One solution is using bioreactors to convey the medium directly throughout the interconnected pores to continuously introduce nutrients and remove wastes, so that the cell viability and activities can be favored in long-term culture within 3D ceramic scaffolds [29]. Teixeira et al. reported that a BCP scaffold, consisting of 60% HA and 40% β-TCP, had the potential to support the chondrocytes isolated from chick embryonic tibia to form bone via the endochondral route [15]. Tortelli et al. compared the biological responses between murine MSCs and mature osteoblasts, when seeded in HA scaffolds and subsequently implanted in immunocompromised mice. The results showed that new vascularized bone was formed through the activation of an endochondral ossification process in the MSCs-scaffolds complex. Conversely, osteoblasts directly formed bone via an intramembranous ossification, without however any signs of vascularization [30]. In another study, mouse embryonic stem cells (ESCs) seeded onto ceramic scaffolds and cultured in osteogenic medium, failed to form bone tissue upon implantation in subcutaneous pockets of nude mice. Differently, when ESCs-ceramic constructs were cultured in chondrogenic differentiation medium, they showed cartilage maturation toward the hypertrophic stage, calcification and ultimately bone formation in vivo [22]. All together, these studies indicated that the origin of bone cells and the ossification type are closely related to the nature and commitment of the seeded cells. 5.2 Hydrogels Hydrogels are networks of hydrophilic, crosslinked polymer chains with a distinctive quality to absorb water in aqueous fluids, and hence to exhibit mechanical and visco-elastic properties that closely resemble the structural three dimensional environment of natural extracellular matrix (ECM). These attributes make hydrogels appealing substrates for cells to adhere, proliferate, secrete new ECM and eventually restore the damage tissue [31]. Most importantly, from a clinical perspective, hydrogels are available ready-to-use grafts, can be injected into the wound site using minimally invasive techniques and allow complete filling of irregularly shaped defects, thus avoiding an open surgery procedure. In combination with stem cells, these scaffolds have been evaluated for their suitability as potential replacements for bone and cartilage. Hydrogels can be broadly classified into natural and synthetic, based on their origin [32]. Natural hydrogels (e.g. agarose, alginate, chitosan, collagen/gelatin, fibrin, hyaluronan, and silk), exhibit excellent bioactivity, as they mimic the native extracellular environment structurally, have 33 2 Chapter 2 unique mechanical properties and are biodegradable by enzymatic or hydrolytic mechanisms. Synthetic hydrogels can be formed by crosslinking of polymer chains through physical (temperature, UV light or pH) or chemical reactions under controlled conditions. Representative synthetically derived polymers explored in combination with stem cells for bone regeneration purposes include poly(ethylene glycol) (PEG), poly(vinyl alcohol) (PVA), oligo(poly(ethylene glycol) fumarate) (OPF) hydrogels, and poly(ethylene glycol)-diacrylate (PEG-DA) hydrogels [33-35]. In regenerative medicine and tissue engineering of bone, the vast majority of studies have been focused on the formation of new bone directly from stem cells or mature osteogenic cells encapsulated within hydrogel matrices. However, more recently, stimulation of cells toward the endochondral ossification using cell-laden hydrogel systems has shown to be not only appealing for the understanding of the regulatory mechanisms of cartilage and bone formation, but also a potential approach for bone repair. 5.2.1 Alginate Alginate is a linear heteropolysaccharides composed of D-mannuronic acid and L-guluronic acid, and are derived primarily from brown algae. Due to the presence of the carboxyl groups along the polymer chain, alginates can form gels in the presence of divalent ions such as calcium ions [36]. Due to easy preparation under gentle conditions, low toxicity and readily availability, alginate hydrogels have been extensively investigated as matrices for encapsulating cells and delivering drugs. Alginate hydrogels have been shown to be suitable matrices for generation of bone tissue via endochondral ossification when associated with committed stem cells and appropriate combinations of regulatory signals (e.g. BMP-2, VEGF, TGF-β3) [37]. Chang et al. followed the progression of osteogenically induced MSCs along the endochondral pathway [38]. At 2 weeks after subcutaneous implantation, cell-alginate constructs presented islands of cartilage formation and cells had the typical round morphology of the chondrocyte phenotype. By 6 weeks, endochondral ossification with trabecular bone deposition started to appear, whereas osteoblasts and osteocytes were seen in the new bone at week 8 and at week 12, respectively. Simultaneously, increasing calcification was detected over time. Similarly, in another study, hMSC were suspended in alginate beads with a chondrogenesis-induction medium containing transforming growth factor TGF-β3. During the first stage of culture, specific chondrogenic markers, such as collagen type II, type X and proteoglycan, were detectable, and their expression decreased over time in concert with the increase of osteogenic markers, such as osteocalcin [39] . 34 Combining Osteochondral Stem Cells and Biomaterials for Bone Regeneration 5.2.2 Chitosan Chitosan is a linear polymer of (1-4) β-linked d-glucosamine residues with N-acetyl glucosamine groups derived from chitin, a naturally occurring polysaccharide which forms the outer shell of crustaceans, insect exoskeletons, and fungal cell walls. Chitosan is completely soluble in aqueous solutions with pH lower than 5.0 and it undergoes biodegradation in vivo enzymatically by lysozyme to nontoxic products [40] . Biocompatibility, biodegradability and structural similarity to natural GAGs make crosslinked chitosans hydrogels attractive materials for tissue engineering applications. Moreover, the feasibility of forming porous scaffolds showed to promote osteoblastic differentiation in vitro and might support angiogenesis and osteogenesis in vivo [41]. Oliveira et al. fabricated the chitosan sponges via freezedrying chitosan acidic solutions and further demonstrated that these chitosan sponges with adequate pore structure and mechanical properties could serve as a support for hypertrophic chondrocytes to induce endochondral ossification [42]. 5.2.3 Collagen Collagen represents the most abundant protein in the human body, being the major component of bone, skin, ligament, cartilage, and tendon. It also forms the structural framework for other tissues such as blood vessels. There are at least 19 different types of collagen, but the basic unit of all is a polypeptide consisting of a three-amino-acid sequence (glycine, proline, and hydroxyproline) forming polypeptide chains, which wrap around one another to form the lefthanded triple helix structure [43]. Due to its excellent biocompatibility, enzymatic degradability and resemblance to the organic composition of natural bone, collagen-based matrices find wide application in tissue engineering and have been extensively investigated in combination with cells and growth factors [44]. Abrahamsson et al. seeded human MSCs suspended collagen gel on a non-woven PCL scaffold. Cultured in chondrogenic condition, cartilaginous tissue was formed by day 21, and hypertrophic mineralization was observed in the newly formed ECM at the interface with underlying scaffold by day 45 [37]. Collagen hydrogels could also be fabricated into porous sponges before cell seeding. MSCs-collagen/ glycosaminoglycan constructs pre-cultured in chondrogenic culture medium and subsequently implanted subcutaneously in the dorsum of nude mice showed good cell survival and progression along the endochondral ossification route, and more interesting, blood vessels formation. Conversely, the osteogenically primed scaffolds showed a mineralized matrix of poor quality, no vascularization, likely due to too much matrix deposition or a lack of release of inductive factors, and few surviving cells, as a result of absence of oxygen and nutrients [4, 20]. In another study, chondrocytes isolated from chick embryo were cultured in collagen sponges treated with retinoic acid to induce chondrocyte maturation 35 2 Chapter 2 and extracellular matrix deposition. Biological properties and stiffness of collagen matrices supported chondrocytes attachment, proliferation and endochondral bone maturation after implantation [45]. 6. Conclusion and future perspective Regenerative medicine is a multidisciplinary field aiming at restoration of normal function to injured tissues, by the development of increasingly sophisticated biomaterials, either alone or in combination with cells and/or stimulating factors. The enormous progress achieved so far at the laboratory stage and the continuous worldwide effort towards new insights and technologies, direct researchers to face new challenges. In order to translate basic scientific advantages into approved medical products, all the materials requirements (e.g. safety and efficacy) must comply the guidelines established by regulatory agencies, including the Food and Drug Administration Agency (FDA) and the European Medicinal Agency (EMEA), and clinical trials must be carried out within acceptable clinical practice and due ethical considerations. In conclusion, although great challenges still remain, research on combining stem cells with biomaterial for the creation of bone grafts is at the precipice of major breakthroughs that likely will change and simplify the treatment of bone disorders, thus to bring major benefits to patients, including reduced need and cost of medical attention, and short-term recovery. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 36 Dominici, M., Hofmann, T.J., and Horwitz, E.M., Bone marrow mesenchymal cells: biological properties and clinical applications. J Biol Regul Homeost Agents, 2001. 15(1): p. 28-37. Sommerfeldt, D.W. and Rubin, C.T., Biology of bone and how it orchestrates the form and function of the skeleton. Eur Spine J, 2001. 10 Suppl 2: p. S86-95. Bord, S., Horner, A., Hembry, R.M., Reynolds, J.J., and Compston, J.E., Production of collagenase by human osteoblasts and osteoclasts in vivo. Bone, 1996. 19(1): p. 35-40. Farrell, E., van der Jagt, O.P., Koevoet, W., Kops, N., van Manen, C.J., Hellingman, C.A., Jahr, H., O’Brien, F.J., Verhaar, J.A., Weinans, H., and van Osch, G.J., Chondrogenic priming of human bone marrow stromal cells: a better route to bone repair? Tissue Eng Part C Methods, 2009. 15(2): p. 285-95. Blau, H.M., Brazelton, T.R., and Weimann, J.M., The evolving concept of a stem cell: entity or function? Cell, 2001. 105(7): p. 829-41. Nakayama, N., Duryea, D., Manoukian, R., Chow, G., and Han, C.Y., Macroscopic cartilage formation with embryonic stem-cell-derived mesodermal progenitor cells. J Cell Sci, 2003. 116(Pt 10): p. 2015-28. Wiles, M.V. and Keller, G., Multiple hematopoietic lineages develop from embryonic stem (ES) cells in culture. Development, 1991. 111(2): p. 259-67. zur Nieden, N.I., Kempka, G., and Ahr, H.J., In vitro differentiation of embryonic stem cells into mineralized osteoblasts. Differentiation, 2003. 71(1): p. 18-27. Wobus, A.M., Potential of embryonic stem cells. Mol Aspects Med, 2001. 22(3): p. 149-64. Combining Osteochondral Stem Cells and Biomaterials for Bone Regeneration 10. Krampera, M., Pizzolo, G., Aprili, G., and Franchini, M., Mesenchymal stem cells for bone, cartilage, tendon and skeletal muscle repair. Bone, 2006. 39(4): p. 678-83. 11. Friedenstein, A.J., Piatetzky, S., II, and Petrakova, K.V., Osteogenesis in transplants of bone marrow cells. J Embryol Exp Morphol, 1966. 16(3): p. 381-90. 12. Williams, D.F., On the nature of biomaterials. Biomaterials, 2009. 30(30): p. 5897-909. 13. Bongio, M., Van Den Beucken, J.J.J.P., Leeuwenburgh, S.C.G., and Jansen, J.A., Development of bone substitute materials: From ‘biocompatible’ to ‘instructive’. Journal of Materials Chemistry, 2010. 20(40): p. 8747-8759. 14. Hutmacher, D.W., Schantz, J.T., Lam, C.X., Tan, K.C., and Lim, T.C., State of the art and future directions of scaffold-based bone engineering from a biomaterials perspective. J Tissue Eng Regen Med, 2007. 1(4): p. 245-60. 15. Teixeira, C.C., Nemelivsky, Y., Karkia, C., and Legeros, R.Z., Biphasic calcium phosphate: a scaffold for growth plate chondrocyte maturation. Tissue Eng, 2006. 12(8): p. 2283-9. 16. Langer, R. and Vacanti, J.P., Tissue engineering. Science, 1993. 260(5110): p. 920-6. 17. Nair, L.S. and Laurencin, C.T., Polymers as biomaterials for tissue engineering and controlled drug delivery. Adv Biochem Eng Biotechnol, 2006. 102: p. 47-90. 18. Lennon, D.P. and Caplan, A.I., Isolation of human marrow-derived mesenchymal stem cells. Exp Hematol, 2006. 34(11): p. 1604-5. 19. Petite, H., Viateau, V., Bensaid, W., Meunier, A., de Pollak, C., Bourguignon, M., Oudina, K., Sedel, L., and Guillemin, G., Tissue-engineered bone regeneration. Nat Biotechnol, 2000. 18(9): p. 959-63. 20. Farrell, E., Both, S.K., Odorfer, K.I., Koevoet, W., Kops, N., O’Brien, F.J., de Jong, R.J., Verhaar, J.A., Cuijpers, V., Jansen, J., Erben, R.G., and van Osch, G.J., In-vivo generation of bone via endochondral ossification by in-vitro chondrogenic priming of adult human and rat mesenchymal stem cells. BMC Musculoskelet Disord, 2011. 12: p. 31. 21. Meijer, G.J., de Bruijn, J.D., Koole, R., and van Blitterswijk, C.A., Cell-based bone tissue engineering. PLoS Med, 2007. 4(2): p. e9. 22. Jukes, J.M., Both, S.K., Leusink, A., Sterk, L.M., van Blitterswijk, C.A., and de Boer, J., Endochondral bone tissue engineering using embryonic stem cells. Proc Natl Acad Sci U S A, 2008. 105(19): p. 6840-5. 23. Mackay, A.M., Beck, S.C., Murphy, J.M., Barry, F.P., Chichester, C.O., and Pittenger, M.F., Chondrogenic differentiation of cultured human mesenchymal stem cells from marrow. Tissue Eng, 1998. 4(4): p. 415-28. 24. Montufar-Solis, D., Nguyen, H.C., Nguyen, H.D., Horn, W.N., Cody, D.D., and Duke, P.J., Using cartilage to repair bone: an alternative approach in tissue engineering. Ann Biomed Eng, 2004. 32(3): p. 504-9. 25. Robins, J.C., Akeno, N., Mukherjee, A., Dalal, R.R., Aronow, B.J., Koopman, P., and Clemens, T.L., Hypoxia induces chondrocyte-specific gene expression in mesenchymal cells in association with transcriptional activation of Sox9. Bone, 2005. 37(3): p. 313-22. 26. Solursh, M., Cartilage stem cells: regulation of differentiation. Connect Tissue Res, 1989. 20(14): p. 81-9. 27. Habibovic, P., Woodfield, T., de Groot, K., and van Blitterswijk, C., Predictive value of in vitro and in vivo assays in bone and cartilage repair--what do they really tell us about the clinical performance? Adv Exp Med Biol, 2006. 585: p. 327-60. 28. Tortelli, F. and Cancedda, R., Three-dimensional cultures of osteogenic and chondrogenic cells: a tissue engineering approach to mimic bone and cartilage in vitro. Eur Cell Mater, 2009. 17: p. 1-14. 29. Du, D., Furukawa, K., and Ushida, T., Oscillatory perfusion seeding and culturing of osteoblastlike cells on porous beta-tricalcium phosphate scaffolds. J Biomed Mater Res A, 2008. 86(3): p. 796-803. 30. Tortelli, F., Tasso, R., Loiacono, F., and Cancedda, R., The development of tissue-engineered bone of different origin through endochondral and intramembranous ossification following the implantation of mesenchymal stem cells and osteoblasts in a murine model. Biomaterials, 2010. 31(2): p. 242-9. 37 2 Chapter 2 31. Nicodemus, G.D. and Bryant, S.J., Cell encapsulation in biodegradable hydrogels for tissue engineering applications. Tissue Eng Part B Rev, 2008. 14(2): p. 149-65. 32. Drury, J.L. and Mooney, D.J., Hydrogels for tissue engineering: scaffold design variables and applications. Biomaterials, 2003. 24(24): p. 4337-51. 33. Betz, M.W., Modi, P.C., Caccamese, J.F., Coletti, D.P., Sauk, J.J., and Fisher, J.P., Cyclic acetal hydrogel system for bone marrow stromal cell encapsulation and osteodifferentiation. J Biomed Mater Res A, 2008. 86(3): p. 662-70. 34. Burdick, J.A. and Anseth, K.S., Photoencapsulation of osteoblasts in injectable RGD-modified PEG hydrogels for bone tissue engineering. Biomaterials, 2002. 23(22): p. 4315-23. 35. Temenoff, J.S., Park, H., Jabbari, E., Sheffield, T.L., LeBaron, R.G., Ambrose, C.G., and Mikos, A.G., In vitro osteogenic differentiation of marrow stromal cells encapsulated in biodegradable hydrogels. J Biomed Mater Res A, 2004. 70(2): p. 235-44. 36. Smidsrod, O. and Skjak-Braek, G., Alginate as immobilization matrix for cells. Trends Biotechnol, 1990. 8(3): p. 71-8. 37. Abrahamsson, C.K., Yang, F., Park, H., Brunger, J.M., Valonen, P.K., Langer, R., Welter, J.F., Caplan, A.I., Guilak, F., and Freed, L.E., Chondrogenesis and mineralization during in vitro culture of human mesenchymal stem cells on three-dimensional woven scaffolds. Tissue Eng Part A, 2010. 16(12): p. 3709-18. 38. Chang, S.C., Tai, C.L., Chung, H.Y., Lin, T.M., and Jeng, L.B., Bone marrow mesenchymal stem cells form ectopic woven bone in vivo through endochondral bone formation. Artif Organs, 2009. 33(4): p. 301-8. 39. Ichinose, S., Yamagata, K., Sekiya, I., Muneta, T., and Tagami, M., Detailed examination of cartilage formation and endochondral ossification using human mesenchymal stem cells. Clin Exp Pharmacol Physiol, 2005. 32(7): p. 561-70. 40. Khor, E. and Lim, L.Y., Implantable applications of chitin and chitosan. Biomaterials, 2003. 24(13): p. 2339-49. 41. Guzman-Morales, J., El-Gabalawy, H., Pham, M.H., Tran-Khanh, N., McKee, M.D., Wu, W., Centola, M., and Hoemann, C.D., Effect of chitosan particles and dexamethasone on human bone marrow stromal cell osteogenesis and angiogenic factor secretion. Bone, 2009. 45(4): p. 617-26. 42. Oliveira, S.M., Mijares, D.Q., Turner, G., Amaral, I.F., Barbosa, M.A., and Teixeira, C.C., Engineering endochondral bone: in vivo studies. Tissue Eng Part A, 2009. 15(3): p. 635-43. 43. Lee, C.H., Singla, A., and Lee, Y., Biomedical applications of collagen. Int J Pharm, 2001. 221(1-2): p. 1-22. 44. Heinemann, C., Heinemann, S., Worch, H., and Hanke, T., Development of an osteoblast/ osteoclast co-culture derived by human bone marrow stromal cells and human monocytes for biomaterials testing. Eur Cell Mater, 2011. 21: p. 80-93. 45. Oliveira, S.M., Ringshia, R.A., Legeros, R.Z., Clark, E., Yost, M.J., Terracio, L., and Teixeira, C.C., An improved collagen scaffold for skeletal regeneration. J Biomed Mater Res A, 2010. 94(2): p. 371-9. 38 Chapter 3 Biological evaluation of porous aliphatic polyurethane/ hydroxyapatite composite scaffolds for bone tissue engineering Wanxun Yang, Sanne K. Both, Yi Zuo, Zeinab Tahmasebi Birgani, Pamela Habibovic, Yubao Li, John A. Jansen and Fang Yang Biological Evaluation of PU/HA scaffolds Summary of this chapter Biomaterial scaffolds meant to function as supporting structures to osteogenic cells play a pivotal role in bone tissue engineering. Recently, we synthesized an aliphatic polyurethane (PU) scaffold via a foaming method using non-toxic components. Through this procedure a uniform interconnected porous structure was created. Furthermore, hydroxyapatite (HA) particles were introduced into this process to increase the bioactivity of the PU matrix. To evaluate the biological performances of these PU based scaffolds, their influence on in vitro cellular behavior and in vivo bone forming capacity of the engineered cell-scaffold constructs was investigated in this chapter. A simulated body fluid (SBF) test demonstrated that the incorporation of 40wt% HA particles significantly promoted the biomineralization ability of the PU scaffolds. Enhanced in vitro proliferation and osteogenic differentiation of the seeded mesenchymal stem cells (MSCs) were also observed on the PU/HA composite. Next, the cell-scaffold constructs were implanted subcutaneously in a nude mice model. After 8 weeks, a considerable amount of vascularized bone tissue with initial marrow stroma development was generated in both PU and PU/ HA40 scaffold. In conclusion, the PU/HA composite is a potential scaffold for bone regeneration applications. 1. Introduction Since the concept of tissue engineering was introduced, researchers have been seeking a solution for bone repair and regeneration by combining the three key factors, i.e. a porous scaffold, exogenous stem cells and biological cues.[1] Essentially, the choice of a biomaterial as scaffold is critical to the success of bone tissue engineering.[2] It is well accepted that an optimal scaffold should possess a porous and biologically compatible framework onto which the bone-forming cells can attach, function, and eventually form new bone tissue.[3] Recently, polyurethanes (PUs) gained popularity and have been investigated as scaffold material. A typical PU is a block copolymer consisting of a hard segment contributed by isocyanates and a soft segment formed by polyether or polyester polyols. For that reason, the material properties of PU, such as mechanical strength, biodegradability and cytocompatibility, can be easily modified by adjusting the components of the hard and soft segments during PU synthesis.[4, 5] For instance, a PU synthesized from aliphatic diisocyanate, e.g. isophorone diisocyanate (IPDI), non-toxic castor oil and 1,4-butanediol (BDO) gives rise to non-toxic degradation products, which is more desirable to be used as a scaffold material than the conventional PUs made from aromatic diisocyanates.[6, 7] However, like most 43 3 Chapter 3 synthetic polymers, PUs lack bioactive groups to facilitate biomineralization. The most common strategy to counteract the poor bioactivity is by introducing bioactive ceramics, e.g. hydroxyapatite (HA) particles, into the PU matrix during the polymerization process.[8, 9] Due to the presence of the polar groups in the molecular chain, PUs have relatively high affinity to HA.[10] Another attractive feature of PUs is that they have a preference to spontaneously foam during the copolymerization process, which facilitates a one-step process of making porous scaffolds. For instance, water, which is either unavoidably present in the reaction mixture or added intentionally during the copolymerization process, reacts with the isocyanate causing the release of CO2 gas. The released CO2 gas creates bubbles and eventually leads to foaming. Our previous study has demonstrated that a mild foaming process could be achieved to allow the formation of a uniform porous PU structure by carefully choosing a low reaction rate isocyanate.[11] We also demonstrated that nanosized HA particles could be added into this process, resulting the formation of a porous PU/HA composite scaffold with good dispersion and high occupancy of HA particles.[11, 12] In this study, we aimed to further test the biological performance of such PU/ HA scaffolds in comparison to HA free PU scaffolds, including the scaffold effect on cellular behavior and in vivo bone forming capacity of the engineered cellscaffold constructs. Firstly, the scaffolds were immersed in simulated body fluid (SBF) to test biomineralization ability. Furthermore, rat bone marrow derived mesenchymal stem cells (MSCs) were seeded onto the scaffolds, and the cell viability and osteogenic capacity were tested in vitro. Finally, the efficacy of bone formation on the cell-scaffold constructs was evaluated in vivo after subcutaneous implantation in nude mice. 2. Materials and methods 2.1 Scaffold fabrication and morphology characterization All the reagents were of analytic grade and purchased from Kelong Co. Ltd. (Chengdu, China), except that castor oil and isophorone diisocyanate (IPDI) were obtained from Aladdin Co. Ltd. (Shanghai, China). Both castor oil and 1,4-butanediol (BDO) were dehydrated under decompression with a vacuum of 1330 Pa at 120°C. IPDI was preserved in a refrigerator before use. Nano-sized HA slurry was prepared by wet synthesis as previously described,[13] then spray-dried to particles of approximately 5-15 μm in diameter for further experiments. The PU and PU/HA scaffolds were prepared by copolymerization and simultaneous foaming following our previously reported method [11]. The reaction was performed 44 Biological Evaluation of PU/HA scaffolds in a 250 ml three-necked round-bottom flask under a dry nitrogen atmosphere. Firstly, 38 g castor oil, and a certain amount of HA powders, were put into the flask and stirred uniformly. Subsequently, 11.2 g IPDI was added drop-wise into the HA/castor oil mixture and the reaction was kept for 4 h to form the prepolymer. Then 4.5 g BDO was used as a chain extender to cross-link the prepolymer and 0.9 g deionized water was added in the cross-linked prepolymer by stirring for 5 min. The mixture was placed in an oven at 120°C for 4 h accompanied with simultaneous foaming, after which the three-dimensional porous bulk scaffold was obtained. Three types of scaffolds were prepared based on the amount of HA. They were named as PU, PU/HA20 and PU/HA40 with the PU:HA weight ratios of 100:0, 80:20 and 60:40, respectively. For all three types of scaffolds, the overall porosity was 78-81%.[11] Disk-shaped samples with a diameter of 6 mm and a thickness of 2.5 mm were punched out of the bulk scaffolds. In order to achieve further exposure of the HA particles on the surfaces, all types of scaffolds were immersed in 20M NaOH solution for 5 days under gentle agitation, as NaOH has the erosive effect on PU.[14] After that all scaffolds were cleaned thoroughly in deionized water and freezedried before being used in the further experiments. The scaffold morphology was examined by scanning electron microscopy (SEM; JEOL6340F, Tokyo, Japan) after being sputter-coated with gold. 2.2 Simulated body fluid (SBF) immersion test SBF immersion test was used to evaluate the biomineralization behavior of the scaffolds. The recipe of SBF was adopted from Kokubo [15] and the pH value of the solution was adjusted to 7.4 after complete preparation. Immersion studies were performed by incubating each sample in 10 ml SBF solution in a 15 ml tube (n=3). All tubes were placed in a water bath at 37 °C under continuous shaking. After immersion periods of 1 and 4 weeks, the scaffolds were gently washed with deionized water and freeze-dried. SEM and elemental analysis of the deposits was carried out using a Philips XL30 scanning electron microscope (Eindhoven, the Netherlands) equipped with an energy dispersive spectrometer (EDS, AMETEK Materials Analysis Division, Mahwah, NJ, USA) after the samples were sputtercoated with gold (Cressington 108A, Watford, UK). The accelerating voltage was 10 KeV and the working distance was 10 mm. EDS analysis provided information on the distribution of elements of interest (Ca and P) in the analyzed area. 2.3 Cell isolation and seeding Rat MSCs were isolated from 6-week-old male Wistar rats after the approval from Radboud University Nijmegen Animal Ethics Committee. Briefly, two femora of 45 3 Chapter 3 each rat were extracted and the epiphyses were cut off. MSCs were flushed out of the remaining diaphyses using the primary cell culture medium, consisting of alpha Minimal Essential Medium (aMEM; Gibco, Grand Island, USA) supplemented with 10% fetal bovine serum (FBS; Sigma, St. Louis, USA), 100 U/ml penicillin and 100 μg/ml streptomycin (Gibco). The flush-out was cultured for two days in a humidified incubator (37 °C, 5% CO2), after which the medium was refreshed to remove non-adherent cells. Prior to detaching the cells, they were cultured for an additional three days. Then, the cells were detached using trypsin/EDTA (0.25% w/v trypsin, 0.02% w/v EDTA; Sigma) and counted. The PU scaffolds for cell culture experiments were sterilized by autoclave and prewetted in primary cell culture medium overnight. For cell seeding, 5× 104 cells were suspended in 50 µl medium and statically seeded onto each scaffold. All scaffolds were placed in 24-well non-adherent culture plates (1 scaffold per well) and incubated for 3 h in a humidified incubator (37 °C, 5% CO2) allowing the initial cell attachment to the scaffolds. Subsequently, osteogenic medium was added, which contained 50 μg/ml ascorbic acid (Sigma), 10nM dexamethasone (Sigma) and 10 mM sodium β-glycerophosphate (Sigma) on the basis of the primary culture medium. Scaffolds were maintained in culture for 24 days, with the cell culture medium being refreshed twice per week. Cell viability, DNA amount, alkaline phosphatase (ALP) activity and osteogenic gene expression were tested. 2.4 Cell viability LIVE/DEAD® Assay (Invitrogen) was used to assess cell viability after 24 h of cell seeding. The cell-scaffold constructs were washed in PBS and exposed to calcein AM / ethidium homodimer-1 / PBS working solution for 30 min at 37 °C according to the manufacturer’s instructions. Dye uptake was detected by using an automated fluorescent microscope (Axio Imager Microscope Z1; Carl Zeiss Micro Imaging GmbH) with a wave length of 488 nm for visualizing the live cells (green) and 568 nm for the dead cells (red). 2.5 DNA content and ALP activity DNA content (n=3) was quantified by PicoGreen assay (Quant-iT PicoGreen dsDNA, Invitrogen) after 4, 8, 16 and 24 days of osteogenic culture condition. After the culture medium was removed, scaffolds were washed twice with PBS. Cells were harvested by placing the cell-scaffold constructs in a 1.5-ml tube. One ml of deionized water was added to each sample, after which 2 freeze-thaw cycles and ultrasonication were performed. One-hundred μl cell lysate or DNA standard from the PicoGreen assay kit was added to 100 μl freshly made working solution in the wells. The results were read using a fluorescence microplate reader (Bio-Tek 46 Biological Evaluation of PU/HA scaffolds Instruments, Abcoude, The Netherlands) with an excitation wavelength at 485nm and an emission wavelength at 530nm. The ALP activity (n=3) was measured as a marker for early osteogenic differentiation using the same samples as for the DNA assay. Eighty μl of cell lysate or standard (serial dilutions of 4-nitrophenol at the concentrations of 0-25 nM) and 20μl of buffer solution (5 mM MgCl2, 0.5 M 2-amino-2-methyl-1-propanol) were added into a 96-well plate. Then 100μl of substrate solution (5 mM paranitrophenylphosphate) was added into all wells. Subsequently, the plate was incubated for 1 h at 37 °C before the reaction was stopped by adding 100 μl of 0.3M NaOH. The plate was read in an ELISA reader (Bio-Tek Instruments) at 405 nm. ALP activity results were normalized by the amount of DNA. 2.6 RNA extraction and real-time PCR analysis To analyze the expression of osteogenic-related genes of the cells seeded on each scaffold, a real-time PCR was performed. Total RNA was extracted using Trizol® method (Invitrogen) after 4 and 8 days of culturing. Briefly, scaffolds with cells were washed with PBS before 1 ml of Trizol® solution was added. The cell extract was then collected, mixed with chloroform and centrifuged. Only the upper aqueous phase was collected and mixed with equal amount of isopropanol. After 10min of incubation at room temperature, the extract was centrifuged and washed twice with 75% alcohol. The RNA pellet was dissolved in RNA free water and the total RNA concentration was measured with spectrophotometer (NanoDrop 2000, Thermal scientific, Wilmington, USA). Table 1. Overview of the primer sequences Forward (5’ → 3’) Reverse (5’ → 3’) Oc CGGCCCTGAGTCTGACAAA GCCGGAGTCTGTTCACTACCTT Bsp TCCTCCTCTGAAACGGTTTCC GGAACTATCGCCGTCTCCATT Runx-2 GAGCACAAACATGGCTGAGA TGGAGATGTTGCTCTGTTCG Gapdh CTTCACCACCATGGAGAAGGC GGCATGGACTGTGGTCATGAG First strand cDNA was reverse transcribed from RNA using the SuperScript FirstStrand Synthesis System kit (Invitrogen). Afterwards cDNA was further amplified and the expression of specific genes was quantified using IQ SYBR Green Supermix PCR kit (BioRad, Hemel Hempstead, United Kingdom) in a real-time PCR (BioRaD, CFX96™ real-time system). Osteogenic markers expressed on RNA level were evaluated, including osteocalcin (Oc), bone sialoprotein (Bsp) and runt-related transcription factor 2 (Runx2) (Table 1). The expression levels were analyzed and 47 3 Chapter 3 compared to the housekeeping gene glyceraldehyde-3-phosphate dehydrogenase (Gapdh). The specificity of the primers was tested separately before the real-time PCR reaction. The expression of the tested genes was calculated via the 2-∆∆Ct method [15] and the PU scaffolds were used as the reference group. 2.7 In vivo implantation Based on the in vitro results, two types of scaffolds, PU and PU/HA40 with distinct biological properties were sterilized by autoclave and selected for the in vivo study. The protocol was approved by the Animal Ethical Committee of the Radboud University Nijmegen Medical Centre (Approval no: RU-DEC 2010-254) and the national guidelines for the care and use of laboratory animals were applied. Before in vivo implantation, 250,000 cells were seeded on each scaffold and cultured for 5 days in vitro. Afterwards, the cell-scaffold constructs were implanted subcutaneously in three male nude mice (HsdCpb:NMRI-nu, Harlan). Surgery was performed under general inhalation anesthesia with a combination of isoflurane, nitrous oxide, and oxygen. All mice received analgesic before and after the surgery. The back of the mice was shaved and disinfected with povidoneiodine. Subsequently, four small longitudinal incisions were made. Lateral to each incision, a subcutaneous pocket was created. A cell-scaffold construct and a cell-free scaffold from both PU and PU/HA40 groups were implanted in each mouse. Afterwards the skin was closed using staples. After 8 weeks, the mice were euthanized by CO2-suffocation for sample collection. 2.8 Histological analysis All the in vivo samples were fixed in 10% phosphate buffered formalin for 30 h and dehydrated through graded ethanol before embedding in methylmetacrylate (MMA, L.T.I., Bilthoven, the Netherlands). Embedded samples were sectioned using a microtome equipped with diamond blade (Leica SP1600, Leica microsystem, Wetzlar, Germany) and stained with methylene blue and basic fuchsin (both reagents from Merck). The staining was imaged using a light microscope (Zeiss Imager Z1, Carl Zeiss AG Microscopy, Germany) equipped with AxioCam MRc5 camera. 2.9 Statistic analysis Statistical significance in this study was evaluated using ANOVA analysis followed by Tukey’s Multiple Comparison Test (GraphPad Software Inc., San Diego, USA). Results are reported as mean values and standard deviation. Differences were considered statistically significant at p < 0.05. 48 Biological Evaluation of PU/HA scaffolds 3. Results 3.1 Scaffold morphology As shown in the SEM micrographs (Figure 1), the PU and PU/HA scaffolds displayed a porous structure, containing both macropores ranging from 300 to 1000 μm and micropores from 50 to 300 μm which were mainly located on the surface of macropores (Figure 1a-c). Compared to PU scaffolds, PU/HA20 and PU/HA40 scaffolds exhibited a less regular shape of the macropores and also a less number of the micropores. In addition, due to the presence of the HA particles, the two PU/HA composites showed an unevenness of the pore wall topography compared to the PU scaffolds (Figure 1d-f). Figure 1. SEM micrographs of the PU(a and d), PU/HA20 (b and e) and PU/HA40 (c and f) scaffolds. The PU and PU/HA scaffolds displayed a porous structure, containing both macropores and micropores. Compared to PU scaffolds, PU/HA20 and PU/HA40 scaffolds exhibited a less regular shape of the macropores and also a less number of the micropores (a-c). In the magnified figures (d-f), an increased unevenness on pore surfaces of PU/HA20 and PU/HA40 was observed due to the presence of HA particles. (White arrows indicate the micropores.) 3.2 Biomineralization test After 1-week immersion in SBF solution, no obvious calcium phosphate (CaP) precipitation was observed on either PU or PU/HA20 scaffolds by SEM, whereas EDS detected marginal quantities of calcium and phosphorus (Figure 2a-d). In comparison, a significant extent of coverage by flake-like crystal structure was 49 3 Chapter 3 Figure 2. Biomineralization of the scaffolds after 1-week immersion in SBF. EDS detected marginal quantities of calcium and phosphorus on PU (a) and PU/HA20 (c) scaffolds, while no obvious CaP precipitation was observed on these scaffolds by SEM (b and d). In comparison, significant coverage by flake-like crystal structure was observed on the PU/HA40 scaffold (e), from which a considerable amount of calcium and phosphorus were detected by EDS (f). observed on the surfaces of the PU/HA40 scaffold (Figure 2e). The EDS detected considerable amount of calcium and phosphorus from these deposits, which confirmed the formation of CaP (Figure 2f). These flake-like CaP depositions formed a dense multi-layer crystal structure when the immersion period reached 4 weeks (Figure 3f). EDS also confirmed a more pronounced signal of calcium and phosphorus on the PU/HA40 scaffolds (Figure 3e). By contrast, only limited CaP crystal deposition was observed to scatter on the surface of PU and PU/HA20 scaffolds after 4 weeks (Figure 3a-d). 50 Biological Evaluation of PU/HA scaffolds 3 Figure 3. Biomineralization of the scaffolds after 4-week immersion in SBF. A limited CaP crystal deposit was shown on the surfaces of PU (a and b) and PU/HA20 (c and d) scaffolds, while a dense multi-layer crystal structure with pronounced calcium and phosphorus composition was detected on the PU/HA40 scaffolds (e and f). 3.3 Cell viability and proliferation As assessed by the Live/Dead assay, more than 95% of the cells were viable on all experimental scaffolds after 24 h of cell seeding (Figure 4). Furthermore, the DNA contents (Figure 5a), associated with the cell numbers, increased gradually until day 8 for all groups. On day 16, a decreased DNA content was detected on all groups compared to that on day 8, while the amount maintained stable thereafter. At all time points, the PU/HA20 and PU/HA40 groups displayed a significantly enhanced DNA content compared to the PU group. From day 16 onwards, a higher DNA content was also found on group PU/HA40 compared to PU/HA20. 51 Chapter 3 Figure 4. Cell viability on PU(a), PU/HA20 (b) and PU/HA40 (c) scaffolds. More than 95% of the cells were viable on all experimental scaffolds after 24 h of cell seeding. (Live cells are stained green and dead cells are stained red.) Figure 5. DNA contents (a) and ALP activity (b) on PU, PU/HA20 and PU/HA40 scaffolds. *: p<0.05, **: p<0.01; error bars represent standard deviation (n=3). 52 Biological Evaluation of PU/HA scaffolds 3.4 Osteogenic differentiation The expression of ALP activity was measured as a marker of osteogenic differentiation of the cells (Figure 5b). Higher ALP activity was found on the PU scaffolds compared to PU/HA composite in the early stage of cell culture until day 8. However, the PU/HA40 scaffolds exhibited considerably higher ALP activity compared to that on PU scaffolds at day 16 and day 24. On RNA level, the PCR results also revealed that the PU/HA composites, compared to the PU scaffold, promoted the expression of Bsp and Runx2 at day 4 (Figure 6a), and the expression of Oc at day 8 (Figure 6b). Figure 6. Expression of osteogenic related genes. Compared to PU, the PU/HA scaffolds promoted the expression of Bsp and Runx2 at day 4 (a), and Oc at day 8 (b). *: p<0.05, **: p<0.01; error bars represent standard deviation (n=4). 3.5 In vivo evaluation All animals remained in good health and no signs of wound complications were observed postoperatively. After 8 weeks, all implanted scaffolds were retrieved. Neither macroscopic signs of inflammation nor adverse tissue responses were discerned. Histological observation showed that all scaffolds (with or without cell seeded) were encapsulated by a thin fibrous layer (3-6 layers of fibroblasts) without significant inflammatory cells infiltration (Figure 7). Inside of the fibrous capsule, new bone formation (stained red) was observed in all cell-scaffold constructs for both PU and PU/HA40 groups, which was present not only at the periphery of the scaffolds (Figure 7a and b), but also penetrated in the core areas (Figure 8). By contrast, the implanted cell-free PU/HA scaffolds only exhibited fibrous capsulation without any bone formation (Figure 7c). Due to the deformation of the scaffolds during histological processing and the transparency of the PU scaffolds after being 53 3 Chapter 3 Figure 7. Capsulation and bone formation on the periphery of the scaffolds after in vivo implantation. All implanted scaffolds (with or without cell seeded) were encapsulated by a thin fibrous layer without significant inflammatory infiltration. Inside of the fibrous capsule, new bone formation (stained red) was observed in cellscaffold constructs for both PU/HA40 (a) and PU (b) groups, but not in the cell-free PU/HA scaffolds (c). (S: scaffolds; black arrow: fibrous capsule) embedded in MMA, quantification of the bone volume on the scaffolds could not be performed to determine which group had higher amount of bone formation. Nevertheless, similar bone quality was observed on the PU and PU/HA40 scaffolds from the histological evaluation. As shown in Figure 8a and b, lamellar like bone tissue was observed inside of the pores and mainly aligned along the pore surface. The central area displayed a bone marrow structure, which was filled with a great number of hematopoietic cells (stained dark blue) and adipocytes (with bubblelike morphology) (Figure 8c-d). Meanwhile, an immature woven bone like structure was also found in some pores, which showed the onset of bony matrix deposition with randomly arranged osteocytes. Osteoblasts were observed to align on the periphery of the newly formed bone matrix (Figure 8d). 4. Discussion In the current study, aliphatic PU scaffolds incorporated with different amounts of HA particles were synthesized by in situ polymerization and simultaneous foaming method, subsequently their biological performance for bone tissue engineering applications was evaluated. Biomineralization of the PU and PU/HA scaffolds, which is related to bone-bonding capacity, was evaluated by examining the apatite formation ability on the scaffold surfaces by incubating the scaffolds in SBF with ion concentrations equal to human blood plasma [15]. Our results indicated that, the incorporation of 40% HA particles significantly enhanced the biomineralization ability of the PU scaffolds. This might be attributed to the partial dissolution of HA and the subsequent release of calcium ions from PU/HA40 scaffold, which was more sufficient to favor CaP deposition compared to that from the PU or PU/HA20 scaffolds. Additionally, the exposure 54 Biological Evaluation of PU/HA scaffolds 3 Figure 8. Bone in-growth to the scaffolds after in vivo implantation. On both PU/ HA40 (a and c) and PU (b and d) constructs with cell seeded, lamellar like bone tissue (stained red) was observed in the core area of the scaffolds, mainly aligning along the pore surface. Surrounded by the bone matrix, a large number of hematopoietic cells (stained dark blue) and adipocytes (with bubble-like morphology) were observed, which displayed a bone marrow structure. (B: bone; BM: bone marrow; S: scaffold) of HA particles on the surface of PU/HA40 provided more nucleation sites for CaP formation and growth.[16] Biocompatibility is a primary feature of scaffold materials for tissue engineering applications. Recently, Williams has re-evaluated the definition of biocompatibility and pointed out that, biocompatibility of a scaffold should not be solely dependent on not eliciting any undesirable effects on the cells or the host as an insertable material [17]. More importantly, the scaffolds should also support the appropriate cellular activity to optimize tissue regeneration.[17] Our results showed that both PU and PU/HA scaffolds maintained the viability and supported progressive growth of the seeded cells in vitro. Notably, loading HA particles into the PU matrix improved the proliferation of the cells. The enhanced cell growth on the PU/HA composite might be closely related to the HA particles which increased the initial anchoring and spreading of serum proteins on the polymer surfaces.[18] 55 Chapter 3 Moreover, the addition of HA could increase surface oxygen,[19] which has been shown to improve the attachment and the proliferation of osteoblast-like cells on biomaterials.[20] An enhanced osteogenic differentiation of the seeded cells has also been demonstrated with the PU/HA scaffolds by the increase of ALP activity and the upregulation of the osteogenic related genes. The up-regulated genes were Bsp and Runx2, which are important for orienting osteoprogenitors toward osteo-lineage and regulating the initial stages of crystal growth,[21] and Oc, which is closely related to the late mineralization process. Addition of HA promoting osteogenic differentiation of the cells is probably attributed to two reasons: 1) the increased roughness of the PU/HA scaffolds, and 2) the release of calcium and phosphate from the PU/HA scaffolds into the cell culture medium or the microenvironment of the seeded cells. Previous studies have shown that an increase roughness in three-dimensional scaffolds resulted in an enhanced osteogenic differentiation. [22] However, such topography is expected to have a limited impact when a highly porous and interconnected scaffold is applied, as it will be only effective for the cells directly attached on the rough surface but not those present in the pores.[23] Thus we speculate that the calcium and phosphate release from HA represents a more effective differentiation signal. Our data is also consistent with other studies, which showed that HA provides a favorable environment for osteoblast-like cell differentiation.[24, 25] The engineered cell-scaffold constructs were further implanted subcutaneously to gain insight in their role to support osteogenic differentiation of the cells after transplantation. The reason to choose an ectopic model instead of an orthotopic model for this study was to eliminate or reduce the number of variables involved in bone formation present in a bony environment, for instance bone stimulating cytokines, endogenous progenitor/stem cells and potentially bone-stimulating mechanotransduction.[26] Our result indicated that the PU and PU/HA40 scaffolds supported the cellular growth in vivo. Furthermore, the cell-scaffold constructs worked as an osteoinductive complex and were capable of generating new bone tissue. After 8 weeks of implantation, the bone was still in an active forming and remodeling phase. On the other hand, no obvious degradation of these two scaffolds was observed after 8 weeks of in vivo implantation. In theory, the PU matrix can undergo degradation through the hydrolysis of ester linkages which were introduced by castor oil.[4] The breakdown of these ester bonds yields hydroxyl and carboxylic groups.[7] The acidic carboxyl group accelerates further hydrolysis and the degradation becomes autocatalytic.[6] Previous results demonstrated an 56 Biological Evaluation of PU/HA scaffolds approximate 10% weight loss of aliphatic PU/HA40 scaffolds after being soaked in PBS for 8 weeks in vitro.[12] In our study, however, no significant degradation was observed after 8 weeks of implantation. The possible reason is, due to the in vivo fluid exchange and transportation, carboxyl products were easily removed from the local environment of the scaffolds, thereby impeding the autocatalyzation process and slowing down the progression of degradation. Considering the bone was still in an active forming and remodeling stage, such degradation rate ensured the PU/HA scaffolds providing sufficient mechanical support for the neotissue before it can fulfill the structural role.[27] Still, further investigations are necessary to evaluate the degradation of the PU/HA scaffolds in a longer term and monitor their replacement by the tissue in-growth. Although the PU/HA40 scaffolds showed higher biomineralization ability and significant enhancement of osteoblastic differentiation of the seeded cells in vitro compared to the PU scaffolds, similar quality of the ectopic bone formation was revealed on these two scaffolds after 8 weeks in vivo. Previous studies have also shown that there was a lack of correlation and predictability of in vitro osteogenic marker expression on subsequent in vivo ectopic bone formation.[23, 28] One possible reason might be that, the release of calcium and phosphate from PU/HA40 was not sufficient to influence the cellular behaviors and local ionic concentration to trigger more bone formation or faster bone maturation once implanted. On the other hand, it should be noted that, the ectopic bone formation occurred in an intradermal environment, which lacks of naturally bone-forming stem cells and stimulators [26]. Such an environment also differs from the condition used for in vitro cell culture where the osteogenic growth factors and supplements are sufficiently supplied. Therefore, it is expected that the PU/HA40 scaffolds will probably show superior behavior in promoting the osteoblastic differentiation of the osteoprogenitors and/or stem cells when applied in the orthotopic locations. Our study warrants further investigation towards a clinical implementation of the PU/ HA scaffolds, including 1) scaling up of the implanted constructs and 2) orthotopic implantation in an immunocompetent animal model in which endogenous stem cells, mechanical load and biochemical/inflammatory factors are closely involved in the bone forming process. 5. Conclusions Porous aliphatic PU and PU/HA composite scaffolds were synthesized by a foaming method and their biological performances were evaluated in this study. The incorporation of 40wt% HA nanoparticles into PU significantly promoted the biomineralization ability of the scaffolds and enhanced the in vitro proliferation 57 3 Chapter 3 and osteogenic differentiation of the seeded MSCs. In vivo implantation revealed that a considerable amount of vascularized bone tissue with marrow stroma development was generated on both PU and PU/HA40 scaffold after 8 weeks. Overall, with improved mechanical strength and efficacy of supporting osteogenesis, the PU/HA composite is a potential scaffold for bone regeneration applications. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 58 Langer, R. and Vacanti, J.P., Tissue engineering. Science, 1993. 260(5110): p. 920-6. Lee, J., Cuddihy, M.J., and Kotov, N.A., Three-dimensional cell culture matrices: state of the art. Tissue Eng Part B Rev, 2008. 14(1): p. 61-86. Hutmacher, D.W., Schantz, J.T., Lam, C.X., Tan, K.C., and Lim, T.C., State of the art and future directions of scaffold-based bone engineering from a biomaterials perspective. J Tissue Eng Regen Med, 2007. 1(4): p. 245-60. Guelcher, S.A., Biodegradable polyurethanes: synthesis and applications in regenerative medicine. Tissue Eng Part B Rev, 2008. 14(1): p. 3-17. Santerre, J.P., Woodhouse, K., Laroche, G., and Labow, R.S., Understanding the biodegradation of polyurethanes: from classical implants to tissue engineering materials. Biomaterials, 2005. 26(35): p. 7457-70. Gorna, K. and Gogolewski, S., Preparation, degradation, and calcification of biodegradable polyurethane foams for bone graft substitutes. J Biomed Mater Res A, 2003. 67(3): p. 813-27. Kavlock, K.D., Pechar, T.W., Hollinger, J.O., Guelcher, S.A., and Goldstein, A.S., Synthesis and characterization of segmented poly(esterurethane urea) elastomers for bone tissue engineering. Acta Biomater, 2007. 3(4): p. 475-84. Rezwan, K., Chen, Q.Z., Blaker, J.J., and Boccaccini, A.R., Biodegradable and bioactive porous polymer/inorganic composite scaffolds for bone tissue engineering. Biomaterials, 2006. 27(18): p. 3413-31. Huang, M., Wang, Y., and Luo, Y., Biodegradable and bioactive porous polyurethanes scaffolds for bone tissue engineering. J Biomedical Science and Engineering, 2009. 2: p. 36-40. Wei, J. and Li, Y., Tissue engineering scaffold material of nano-apatite crystals and polyamide composite. European Polymer Journal, 2004. 40(3): p. 509-515. Wang, L., Li, Y., Zuo, Y., Zhang, L., Zou, Q., Cheng, L., and Jiang, H., Porous bioactive scaffold of aliphatic polyurethane and hydroxyapatite for tissue regeneration. Biomed Mater, 2009. 4(2): p. 025003. Liu, H., Zhang, L., Zuo, Y., Wang, L., Huang, D., Shen, J., Shi, P., and Li, Y., Preparation and characterization of aliphatic polyurethane and hydroxyapatite composite scaffold. Journal of Applied Polymer Science, 2009. 112(Compendex): p. 2968-2975. Correia, R.N., Magalhaes, M.C.F., Marques, P.A.A.P., and Senos, A.M.R., Wet synthesis and characterization of modified hydroxyapatite powders. Journal of Materials Science: Materials in Medicine, 1996. 7(Compendex): p. 501-505. Schweitzer, P.A., Corrosion Resistance Tables: Metals, Nonmetals, Coatings, Mortars, Plastics, Elastomers and Linings, and Fabrics. CRC Press; 2004. 3496P. Kokubo, T. and Takadama, H., How useful is SBF in predicting in vivo bone bioactivity? Biomaterials, 2006. 27(15): p. 2907-15. Yang, F., Both, S.K., Yang, X., Walboomers, X.F., and Jansen, J.A., Development of an electrospun nano-apatite/PCL composite membrane for GTR/GBR application. Acta Biomater, 2009. 5(9): p. 3295-304. Williams, D.F., On the mechanisms of biocompatibility. Biomaterials, 2008. 29(20): p. 2941-53. Bonzani, I.C., Adhikari, R., Houshyar, S., Mayadunne, R., Gunatillake, P., and Stevens, M.M., Synthesis of two-component injectable polyurethanes for bone tissue engineering. Biomaterials, 2007. 28(3): p. 423-33. Biological Evaluation of PU/HA scaffolds 19. Laschke, M.W., Strohe, A., Menger, M.D., Alini, M., and Eglin, D., In vitro and in vivo evaluation of a novel nanosize hydroxyapatite particles/poly(ester-urethane) composite scaffold for bone tissue engineering. Acta Biomater, 2010. 6(6): p. 2020-7. 20. Poulsson, A.H., Mitchell, S.A., Davidson, M.R., Johnstone, A.J., Emmison, N., and Bradley, R.H., Attachment of human primary osteoblast cells to modified polyethylene surfaces. Langmuir, 2009. 25(6): p. 3718-27. 21. Seyedjafari, E., Soleimani, M., Ghaemi, N., and Shabani, I., Nanohydroxyapatite-Coated Electrospun Poly(l-lactide) Nanofibers Enhance Osteogenic Differentiation of Stem Cells and Induce Ectopic Bone Formation. Biomacromolecules, 2010. 22. Vlacic-Zischke, J., Hamlet, S.M., Friis, T., Tonetti, M.S., and Ivanovski, S., The influence of surface microroughness and hydrophilicity of titanium on the up-regulation of TGFbeta/BMP signalling in osteoblasts. Biomaterials, 2011. 32(3): p. 665-71. 23. Vaquette, C., Ivanovski, S., Hamlet, S.M., and Hutmacher, D.W., Effect of culture conditions and calcium phosphate coating on ectopic bone formation. Biomaterials, 2013. 34(22): p. 5538-51. 24. Kim, K., Dean, D., Lu, A., Mikos, A.G., and Fisher, J.P., Early osteogenic signal expression of rat bone marrow stromal cells is influenced by both hydroxyapatite nanoparticle content and initial cell seeding density in biodegradable nanocomposite scaffolds. Acta Biomater, 2011. 7(3): p. 1249-64. 25. Ozawa, S. and Kasugai, S., Evaluation of implant materials (hydroxyapatite, glass-ceramics, titanium) in rat bone marrow stromal cell culture. Biomaterials, 1996. 17(1): p. 23-9. 26. Scott, M.A., Levi, B., Askarinam, A., Nguyen, A., Rackohn, T., Ting, K., Soo, C., and James, A.W., Brief review of models of ectopic bone formation. Stem Cells Dev, 2012. 21(5): p. 655-67. 27. Hutmacher, D.W., Scaffolds in tissue engineering bone and cartilage. Biomaterials, 2000. 21(24): p. 2529-43. 28. Chai, Y.C., Roberts, S.J., Desmet, E., Kerckhofs, G., van Gastel, N., Geris, L., Carmeliet, G., Schrooten, J., and Luyten, F.P., Mechanisms of ectopic bone formation by human osteoprogenitor cells on CaP biomaterial carriers. Biomaterials, 2012. 33(11): p. 3127-42. 59 3 Chapter 4 In vivo bone generation via the endochondral pathway on three-dimensional electrospun fibers Wanxun Yang, Fang Yang, Yining Wang, Sanne K. Both and John A. Jansen In Vivo Endochondral Bone Generation on Wet-electrospun Fibers Summary of this chapter A new concept of generating bone tissue via the endochondral route might have superiority over the standard intramembranous ossification approach. To implement the endochondral approach, suitable scaffolds are required to provide a three-dimensional (3D) substrate for cell population, differentiation and eventually for the generation of osteochondral tissue. Therefore, a novel wet-electrospinning system, using ethanol as collecting medium, was exploited in this chapter to fabricate a cotton-like poly(lactic-co-glycolic acid) (PLGA)/poly(εcaprolactone) (PCL) scaffold that consisted of an accumulation of fibers in a very loose and uncompressed manner. On these scaffolds rat bone marrow cells were seeded and chondrogenically differentiated in vitro for 4 weeks followed by subcutaneous implantation in vivo for 8 weeks. Cell pellets were used as a control. The glycosaminoglycan (GAG) assay and Safranin O staining showed that the cells infiltrated throughout the scaffolds and deposited an abundant cartilage matrix after in vitro chondrogenic priming. Histological analysis of the in vivo samples revealed extensive new bone formation through the remodeling of the cartilage template. In conclusion, using the wet-electrospinning method, we are able to create a 3D scaffold in which bone tissue can be formed via the endochondral pathway. This system can be easily processed for various assays and histological analysis. Consequently, it is more efficient than the traditional cell pellets as a tool to study endochondral bone formation for tissue engineering purposes. 1. Introduction Bone lesions and defects present a significant problem in clinics due to the limited self-healing capacity of the bone tissue. Therefore, the use of a biodegradable scaffold provided with multipotent cells and biological cues offers a promising approach for bone repair and regeneration [1]. Since the introduction of this concept, experiments to engineer bone tissue have been primarily focused on a process resembling intramembranous ossification, i.e. direct osteoblastic differentiation [2] . However, the success of this approach is hindered by poor vascularization inside the entire cell-based construct [3]. In line with skeleton development and bone fracture healing, a new tactic has been reported recently, which is based on mimicking endochondral bone formation. Thereby, the bone is generated after a cartilage intermediate stage instead of direct osteoblastic differentiation [4-7] . The rationale behind this approach is that chondrocytes are able to survive with limited nutrition and oxygen. Secondly, they can secrete vascular endothelial growth factor (VEGF) in the hypertrophic stage, which is beneficial for blood vessel in-growth [8]. As a consequence, this route has the potential to circumvent the 63 4 Chapter 4 issue faced by the conventional bone tissue engineering, that is upon implantation the constructs will be cut off from the supply of nutrition and oxygen. To study this endochondral pathway, the cell pellet culture system is so far the most frequently used technique [9]. The cell pellet system provides a threedimensional (3D) environment and allows cell-cell interactions, which are similar to pre-cartilage condensation during embryonic development. However, this model is not suitable for the regeneration of bone tissue because of limitations in the size, quantity, and lack of mechanical stability [10]. As a consequence, for final clinical application of this endochondral pathway a scaffold material has to be used, which can serve as a carrier for the cells to provide a 3D environment as well as mechanical support. In this way, a robust cell proliferation and differentiation can be achieved, which leads to the formation of proper extracellular matrix (ECM) and eventually functional tissue. Tremendous efforts have been made in the development of bone supporting scaffolds with different compositions and 3D configurations using a wide variety of techniques, among which electrospinning has recently gained popularity. Electrospinning, also named electric spinning, is known as a simple and versatile method to produce fibrous polymeric meshes [11]. With the range of tens of nanometers to a few microns in diameter and the form of non-woven structure, electrospun fibers are morphologically similar to natural ECM, which makes them attractive for cells to populate on and to function effectively [12, 13]. However, a scaffold made by the conventional electrospinning process is mostly composed of a compact fiber network; thus it is very difficult for the cells to penetrate into the internal part of the scaffold [14, 15]. Such a scaffold cannot be considered a 3D structure as the cells only populate on the superficial surface. To address the aforementioned issues, the aim of this study was to fabricate a 3D scaffold using a modified electrospinning technique, and to further test its efficacy in endochondral bone formation. To this end, the newly developed scaffolds were seeded with rat bone marrow cells (RBMCs), followed by in vitro chondrogenic differentiation and in vivo subcutaneous implantation. It was hypothesized that the RBMCs can generate bone in vivo on the 3D electrospun scaffolds via the endochondral pathway. 2. Materials and methods 2.1. Fabrication of electrospun scaffolds Poly(lactic-co-glycolic acid) (PLGA, Purasorb® PDLG 8531) and poly(ε-caprolactone) (PCL, LACTEL® Absorbable Polymers, inherent viscosity range: 1.0 – 1.3 dl/g) were 64 In Vivo Endochondral Bone Generation on Wet-electrospun Fibers purchased from Purac Biomaterials BV (Gorinchem, The Netherlands) and DURECT Corporation (Pelham, USA), respectively, and used in the electrospinning process. Organic solvent 2,2,2-trifluoroethanol (TFE) (purity≥99.8%) was obtained from Acros (Geel, Belgium). The electrospinning solution was prepared by dissolving PLGA/PCL (weight ratio 3:1) in TFE at a concentration of 0.12 g/ml. The 3D scaffolds were fabricated using a so-called wet-electrospinning technique as depicted in Fig. 1 in a commercial available electrospinning set-up (Esprayer ES-2000S, Fuence Co., Ltd, Tokyo, Japan). The optimal processing parameters for stable formation of electrospun fibers were selected based on an earlier pilot study. Briefly, the prepared polymer solution was fed into a syringe and delivered to an 18G nozzle at a feeding rate of 50 μl/min. A high voltage of 20 kV was applied at the nozzle to generate a stable polymer jet by overcoming the surface tension of the polymer solution [16]. Different from the conventional electrospinning setup, Figure 1. Illustration of 3D scaffold generation by the wet-electrospinning method. A polymer solution was fed into a syringe and a high voltage was applied at the nozzle to generate the polymer jet. A grounded bath filled with 100% ethanol was used as collector. As ethanol is a wetting agent for both PLGA and PCL, the resulting fibers formed a loose cotton-like mesh in the bath. 65 4 Chapter 4 a grounded bath filled with 100% ethanol was used to collect the fibers, located 15 cm under the nozzle. To control the size of resulting fiber meshes, the process was stopped every 10 minutes for fiber mesh collection. Subsequently, the wetelectrospun scaffolds were washed thoroughly in MilliQ water and freeze-dried for 3 days. For comparison, electrospun fibers were also collected by the conventional electrospinning method on top of flat aluminum foil, using the same processing parameters and collecting time. 2.2. Characterization of the scaffolds The morphology of the wet- and conventional electrospun scaffolds was observed by scanning electron microscopy (SEM; JEOL6340F, Tokyo, Japan) after being sputter-coated with gold-platinum. The fiber diameters were measured from the SEM micrographs as obtained at random locations (n=30) using Image J software (National Institutes of Health, Bethesda, USA). The porosity of scaffolds was evaluated by a gravimetric measurement [17]. In brief, the electrospun scaffolds (n = 4) were punched into disc shape forms (a diameter of 6 mm for wet-electrospun scaffolds and a diameter of 15 mm for conventional scaffolds) and their dimensions were measured to calculate the volume of the scaffolds. The weight of the scaffolds was also measured to determine the apparent density of the scaffolds (ρap). The porosity was then calculated according to the following equation: Porosity = (1 - ρap/ ρm) × 100% where ρm is the density of the blend PLGA/PCL and was calculated as 1.215 g/cm3 based on the weight ratio and respective densities of PLGA (ρ = 1.24 g/cm3, Purac MSDS database) and PCL (ρ = 1.145 g/cm3, Sigma-Aldrich MSDS database). 2.3. Cell isolation, seeding and culture RBMCs were isolated from 7-week-old male Fischer rats after the approval from Radboud University Nijmegen Animal Ethics Committee (Approval no: RU-DEC 2011-142). Briefly, two femora of each rat were removed and the epiphyses were cut off. RBMCs were flushed out of the remaining diaphyses using the proliferation medium through an 18G needle (BD Microlance, Drogheda, Ireland). The proliferation medium consisted of alpha Minimal Essential Medium (aMEM; Gibco, Grand Island, USA) supplemented with 10% fetal bovine serum (FBS; Sigma, St. Louis, USA), 100 U/ml penicillin, 100μg/ml streptomycin (Gibco), 50 μg/ml l-ascorbic acid (Sigma) and 10nM dexamethasone (Sigma). The flush-out was cultured for 2 days in a humidified incubator (37 °C, 5% CO2), after which the medium was refreshed to remove non-adherent cells. Prior to detaching the cells, 66 In Vivo Endochondral Bone Generation on Wet-electrospun Fibers they were cultured for an additional 3 days then the cells were detached using trypsin/EDTA (0.25% w/v trypsin, 0.02% w/v EDTA; Sigma) and counted. Disk-shaped scaffolds with a diameter of 6 mm and a thickness of about 2.5 mm were punched out (biopsy punch; Kai medical, Gifu, Japan) from each wetelectrospun mesh that was collected for 10 minutes. Afterwards they were sterilized in 70% ethanol for 2 hours and soaked in the proliferation medium overnight after washed with PBS. During cell loading, the scaffolds were incubated in the cell suspension with a concentration of 1 × 106 cells/ml (4 scaffolds per 1 ml of cell suspension) and gently rotated for 3 hours. Successively, the scaffolds were placed in non-adherent tissue culture plates and the unattached cells from the suspension were collected, centrifuged and reseeded onto the scaffolds to ensure a high cell loading efficiency. The cell-scaffold constructs were then cultured for 1 week in proliferation medium. Thereafter, the medium was changed to chondrogenic medium, consisting of high-glucose DMEM (Gibco), 1% FBS, 100 U/ ml penicillin, 100 μg/ml streptomycin, 50 μg/ml l-ascorbic acid, 100 mM of sodium pyruvate (Gibco), 1:100 insulin-transferrin-selenium (ITS; BD Biosciences, Bedford, USA), 100nM dexamethasone, 10 ng/ml of transforming growth factor beta-2 (TGF-β2; R&D systems, Minneapolis, USA) and 100 ng/ml bone morphogenetic protein 2 (BMP-2; BD biosciences). The constructs were cultured for 4 weeks and the medium was refreshed twice a week. As a control, also a pellet culture was made to induce chondrogenic differentiation. To ensure a similar amount of cell doublings compared to the constructs, the cells were re-seeded after the first passage in tissue culture flasks and kept in proliferation medium for 1 week. Subsequently, cell pellets (250,000 cells/pellet) were made by centrifugation for 8 minutes at 200g and cultured in the chondrogenic medium for 4 weeks. Both the cell-scaffold constructs and the pellets were then used for in vitro and in vivo tests. 2.4. In vitro studies After 4 weeks of chondrogenic differentiation, both the cell-scaffold constructs and the cell pellets were tested for their glycosaminoglycan (GAG) and DNA content. A Safranin O staining was performed to analyze the in vitro formed cartilage. The amount of GAG in each sample was detected by a sulfated glycosaminoglycan (sGAG) assay (n=3). After being washed twice in PBS, all samples were digested in proteinase K solution (1 mg/ml proteinase K in 50 mM Tris with 1 mM EDTA, 1 mM iodoacetamide, and 10 μg/ml pepstatin A) (all reagents from Sigma) for 16 hours at 56°C. The GAG content was determined by using dimethylmethylene blue (Polysciences Inc., Warrington, USA), with chondroitin sulfate (Sigma) as standard control. All results were measured with an ELISA reader (Bio-Tek Instruments Inc., 67 4 Chapter 4 Winooski, USA) at 530 nm and 590 nm. For each sample, the GAG content was normalized by the amount of DNA as described below. DNA content (n=3) was quantified by PicoGreen assay (Quant-iT PicoGreen dsDNA, Invitrogen, Oregon, USA) using the same cell extract solution as used for the GAG assay. A DNA standard curve was made and the results were obtained using a fluorescence microplate reader (Bio-Tek Instruments Inc.) with an excitation wavelength at 485nm and an emission wavelength at 530nm. For the Safranin O staining, the in vitro scaffolds (n=4) were fixed in 10% phosphate buffered formalin for 1 hour and embedded in paraffin. The microtome sections with 6 μm of thickness were stained with Safranin O (Merck, Darmstadt, Germany) after deparaffinization in xylene and rehydration through graded ethanol. 2.5. In vivo implantation After 4 weeks of in vitro culture, the cell-scaffold constructs and the cell pellets were implanted subcutaneously in ten 8-week-old male nude rats (Crl:NIH-Foxn1mu, Charles River). The protocol was approved by the Animal Ethical Committee of the Radboud University Nijmegen Medical Centre (Approval no: RU-DEC 2011-142) and the national guidelines for the care and use of laboratory animals were applied. Surgery was performed under general inhalation anesthesia with a combination of isoflurane, nitrous oxide, and oxygen. The back of the rats was shaved and disinfected with povidone-iodine. Subsequently, two small longitudinal incisions were made. Lateral to each incision, a subcutaneous pocket was created. After placing the samples, the skin was closed using staples. One cell-scaffold construct or three cell pellets were inserted in each pocket. After 8 weeks, the rats were euthanized by CO2-suffocation for sample collection. 2.6. Histological staining All in vivo samples were fixed in 10% phosphate buffered formalin for 36 hours and dehydrated through graded ethanol before embedding in either methylmetacrylate (MMA, L.T.I., Bilthoven, the Netherlands) or paraffin. MMA embedded samples (n = 5) were sectioned using a sawing microtome technique (Leica SP1600, Leica microsystem, Wetzlar, Germany) and stained with methylene blue and basic fuchsin (both reagents from Merck). Prior to paraffin embedding, the fixed samples (n = 5) were decalcified by 10% EDTA and the decalcification process was monitored via X-ray. Six μm-thick sections were made from each sample and used for hematoxylin/eosin (HE), elastic Van Gieson (EVG) and Safranin O staining. 68 In Vivo Endochondral Bone Generation on Wet-electrospun Fibers 2.7. Back-scattering scanning electron microscopy (BS-SEM) Thick MMA sections (n=3) were polished by silicon carbide sandpapers up to grit 4000 and then sputter-coated with gold–platinum. The micrographs were acquired by high resolution SEM in the backscattering mode at 25kV (JEOL6340F, Tokyo, Japan). 2.8. Statistic analysis All data in this study was reported as means ± SD. Statistical significance was evaluated using an unpaired t-test (Instat 3.05, GraphPad Software Inc., San Diego, USA). Differences were considered statistically significant at p < 0.05. 3. Results 4 3.1. Scaffold characterization The wet-electrospinning technique developed in this study displayed obvious differences in the collected fiber structures compared with conventional electrospinning. After 10 minutes of collection, the wet-electrospun PLGA/PCL mesh showed a loose and fluffy structure with a diameter of 3-4 cm and a thickness of 2-3 mm (Fig. 2a, I). In contrast, the fibers collected via the conventional way formed a thin and compressed membrane with a diameter of around 12 cm and a thickness of 40 μm (Fig. 2a, II). The SEM micrographs show that the fibers were oriented in a random and dispersive manner, forming a porous structure for both types of scaffolds (Fig. 2b and 2c). The average fiber diameters measured from the SEM micrographs are 1.98 ± 0.51 µm for the wet-electrospun scaffolds and 2.01 ± 0.56 µm for the conventional electrospun scaffolds, with no statistical significant difference. However, there exists a significant difference between the measured porosities, showing 99.0 ± 0.2% for the wet-electrospun scaffolds and 79.4 ± 2.8% for the conventional scaffolds, respectively. 3.2. DNA and GAG content RBMCs were seeded onto the scaffolds to test their chondrogenic potential compared to the cell pellet system. Prior to in vivo implantation the amount of DNA and deposited GAG in the pellets and scaffolds were measured. As shown in Fig. 3a, a significantly higher amount of DNA could be found in the PLGA/PCL scaffolds compared to the cell pellets after 4 weeks of chondrogenic differentiation in vitro. Moreover, the amount of GAG formed on the PLGA/PCL scaffolds, after normalization by the DNA amount, was five times higher compared to that of the cell pellets (Fig. 3b). 69 Chapter 4 Figure 2. Comparison of different PLGA/PCL scaffolds. The cotton-like scaffold made via the wet-electrospinning method (a, I) showed an uncompressed structure and increased thickness compared to the membrane made via the conventional method (a, II). Dispersive and randomly oriented fiber arrangement was observed by SEM for both wet-electrospun scaffolds (b) and conventional electrospun scaffolds (c), showing a porous structure. 3.3. Histological analysis of in vitro cartilage formation To visualize the cartilage matrix deposition in the cell pellets and the scaffolds, sections were made and stained with Safranin O after 4 weeks of chondrogenic culture. As shown in Fig. 3e, the cell-scaffold constructs displayed an abundant GAG content stained strongly in orange-red color, representing the cartilage matrix. The PLGA/PCL fibers were embedded within the cartilage matrix and a homogeneous distribution of cells could be observed inside of the fibrous structure. The cell morphology differed due to the different stages of cell differentiation. A proportion of cells displayed fibroblast-like morphology, while a great number of cells had a typical chondrocyte morphology embedded in large lacunae and formed isogenous groups (Fig. 3f). In contrast, the cell pellets only displayed chondrogenic differentiation in the outer layer. There was no cartilage matrix found in the central region and a number of cells displayed a necrotic phenotype, leaving empty spaces inside of the pellets (Fig. 3c and d). 70 In Vivo Endochondral Bone Generation on Wet-electrospun Fibers 3.4. Histological analysis of in vivo bone formation After 8 weeks of implantation, an increased hardness of the scaffolds could be sensed during sample retrieval. Unfortunately, the implanted cell pellets could not be traced. To evaluate the bone forming quality and quantity, different histological staining methods were used for either MMA or paraffin embedded samples. 4 Figure 3. DNA content and chondrogenic differentiation. After 28 days of chondrogenic differentiation in vitro, a significantly higher amount of DNA (a) and GAG (b, normalized for DNA), was found in the PLGA/PCL scaffolds compared to the cell pellets. In the cell pellets, only a limited region in outer layer underwent chondrogenic differentiation (c and d). In PLGA/PCL scaffolds, abundant GAG (orange-red color) was secreted, which was visualized by a Safranin O staining (e). Cells with different morphologies could be observed; chondrocyte-like cells were embedded in lacunae and formed isogenous groups (f, black arrows), while a proportion of cells kept their original fibroblast-like morphology (f, white arrows). ** p<0.01; error bars represent standard error of the mean (n=3). 71 Chapter 4 Four out of five MMA embedded scaffolds showed bone formation. The bone forming quality and quantity were similar in the four samples. As shown in Fig. 4a, extensive bone formation can be observed inside the scaffolds. Islets of bone tissue are scattered amongst the remnants of scaffold material, indicating the onset of the newly formed bone (Fig. 4b). Concentrically organized bone structures which indicate mature bone could hardly be seen in the bone forming area. Figure 4. In vivo bone formation in PLGA/PCL electrospun scaffolds. Sections of samples after 8 weeks in vivo were stained with methylene blue and basic fuchsin. Bone formation (red) occurred inside of the scaffold (a). The osteocytes embedded in the bone matrix showed a random arrangement (b, white arrows). The onset of the newly formed bone tissue (b, black arrows) was scattered amongst the remnants of scaffold material. To better discriminate tissue structures within the scaffolds, HE staining was applied to the paraffin embedded samples. All five scaffolds demonstrated bone formation with similar quality and quantity. As shown in Fig. 5a, woven bone like structure can be observed throughout the scaffold. A large amount of immature osteocytes with round shapes were embedded and arranged randomly, while chondrocytes were hardly seen at this time point. The newly formed bone resembled a mixture of cartilage and calcified cartilage tissues (bone). The vascular invasion could be observed among the material remnants and the fibrous tissue (Fig. 5d). To further investigate the state of bone maturation, an EVG staining was performed. Thereby the bone area and the surrounding fibrous tissue, as well as the different maturities of the new bone, were distinguished by the clear contrast between the red, pink and light brown colors, representing bone, uncalcified cartilage and fibrous tissue, respectively (Fig. 5c). 72 In Vivo Endochondral Bone Generation on Wet-electrospun Fibers 4 Figure 5. Histological stainings of PLGA/PCL electrospun scaffolds after in vivo implantation. After 8 weeks of implantation, HE staining (a and d) showed an extensive distribution of woven bone like structure and blood vessel infiltration (d, black arrows). Safranin O staining (b and e) confirmed the existence of remaining GAG (red color) (e, black arrow) among calcified cartilage (green color) (e, asterisk). Bone (f, asterisk), uncalcified cartilage (f, black arrow) and the surrounding fibrous tissue were distinguished by the contrast between red, pink and light brown color in EVG staining (c and f). Furthermore, a Safranin O staining (Fig. 5b and e) confirmed the existence of the remaining GAG (red color) among calcified cartilage (green color). Compared to the cell-scaffold constructs before implantation (Fig. 3e), the in vivo samples consisted extracellular matrix with considerably reduced GAG levels. Although a few sporadic cells with strong red color remained, most regions of the cartilage matrix were remodeled and calcified into bone tissue. 3.5. BS-SEM analysis The grey levels in the BS-SEM images showed that there is a clear difference between the electrospun fibers and newly formed bone (Fig. 6a). Bone tissue, with the highest density, displayed by a grey opaque in the black background of the MMA block. A great number of electrospun fibers were distributed dispersively inside the bone area, visualized by black round dots or needle-like shapes, depending on the sectioning angles (Fig. 6b). Moreover, the differences in bone density could also be addressed according to the grey scales. Immature bone was found in most of the bone forming areas displayed by a relatively darker color. 73 Chapter 4 4. Discussion In this study, we successfully developed a simple wet-electrospinning system by using ethanol as the collecting media to fabricate a highly porous 3D electrospun scaffold for bone tissue engineering applications. Successively, the capacity of RBMCs to generate de novo bone tissue in vivo through the endochondral pathway was investigated using this 3D scaffold. After 4 weeks of in vitro chondrogenic priming, a cell infiltration throughout the scaffolds and abundant cartilage matrix formation were observed. Upon subcutaneous implantation, the bone tissue was generated ectopically by extensive calcification of the cartilage templates with significant blood vessel infiltration. Figure 6. Back scattering SEM images of PLGA/PCL electrospun scaffolds after in vivo implantation. In the black background of the MMA (M), bone forming area (B) demonstrated grey opaque with a great number of electrospun fibers (black color) penetrating. In the magnified image of the red squared box (b), bone tissue with different densities showed in different grey scales; brighter colors indicated the regions of more mature bone. Essentially, the lack of a real 3D structure in conventional electrospun scaffolds hampers their use in tissue regeneration. To overcome this drawback, various methods have been recently explored. A salt leaching/gas forming technique was used to introduce micro-voids into the scaffold after electrospinning [18]. However, the fiber morphology of the scaffolds could not be well preserved after the postspinning processes. Ultrasonication of the electrospun membrane in an aqueous solution was also applied to make the densely packed fibers become loose, while the increase in pore size and the thickness of the membrane was limited [19]. Pham et al. designed bilayered constructs consisting of microfibers with varying thicknesses of nanofibers on top to improve the cell infiltration [17]. However, 74 In Vivo Endochondral Bone Generation on Wet-electrospun Fibers this method implied an increased complexity of the technique. To modify the collecting apparatus, Blakeney et al. embedded a metal array in a semi-spherical dish which allowed the fibers to accumulate in an open space and formed a focused and uncompressed 3D scaffold, resembling a cotton ball [20]. It was shown that such a scaffold allowed better cell penetration and proliferation compared to the conventional electrospun scaffolds. In this study, a similar cotton ball like structure was made by wet-electrospinning technique, simply using ethanol as the collecting media. The general technique of electrospinning into a coagulation bath was first introduced by Yokoyama et al. in 2009 [21]. Using ethanol as collecting media was first reported by Fang et al. and used for studying the evolution of fiber morphology [22]. As ethanol is a wet agent for synthetic polymers, e.g. PLGA and PCL, it prevents the electrospun fibers from densely packing on top of each other in contrary to the conventional flat collector. Afterwards, ethanol was replaced by water and the sample was freeze-dried. During this process, the loose structure could be well preserved and characterized. Moreover, the use of ethanol did not alter the morphology and size of the electrospun fiber. To further prove the efficacy of wet-electrospinning, fabricated 3D PLGA/PCL scaffolds were combined with RBMCs to test their bone forming ability in vivo. Loosened electrospun scaffolds have been used before for bone tissue engineering purposes [23, 24]. However, these were implanted in rodent calvaria using different methods for cell differentiation. Our study is the first instance of endochondral bone formation subcutaneously using an electrospun scaffold as a tool to improve efficacy. To achieve in vivo bone generation, we chose the pathway that resembles endochondral ossification by first forming a cartilage template in vitro and subsequently allowing the cartilage to be replaced by bone in vivo. This process relies on the specialized functions of hypertrophic chondrocytes, from which a type X collagen-rich avascular cartilaginous matrix can be produced in vitro. After in vivo implantation, it is expected that the hypertrophic chondrocytes at the periphery of the cartilage template instruct the surrounding mesenchymal cells to differentiate into osteoblasts. In parallel, the chondrocytes in the central regions lead mineralization of the hypertrophic cartilage by initiating remodeling via the production of specific matrix metalloproteinases (MMPs) and attract blood vessels by releasing VEGF [25]. Using this approach, a high number of well-organized blood vessels were present within the scaffolds, which showed the feasibility and superiority of endochondral route for engineering a vascularized bone construct. As demonstrated by Chan at al [26], endochondral ossification is required for the haematopoietic stem cell niche formation with a subpopulation of fetal progenitor cells giving rise to bone with a marrow cavity. This happens only if they would undergo endochondral ossification as opposed to intramembranous ossification. 75 4 Chapter 4 In this study, extensive and robust bone formation was demonstrated. It is known that osteogenesis and angiogenesis are two closely correlated processes during bone growth, remodeling and repair [27]. In the endochondral bone formation, the released VEGF regulates blood vessel invasion into the hypertrophic cartilage [28]. Afterwards, undifferentiated mesenchymal cells are brought to the mineralization front through the invading blood vessels thereby participating in the osteogenesis [29] . Thus the turnover from cartilage to bone is synergized when a good vascularization is established inside the cell-scaffold constructs. In this study, cell pellets were used as control because they provide a 3D environment and high cell density for cell-cell interactions which mimic the precartilage condensation during embryonic development [30, 31]. This system has been applied as a study model for engineering chondrogenically differentiated constructs in vitro and also for the creation of vascularized bone in vivo [32]. The in vitro results in our study showed greater cartilage production in PLGA/PCL scaffolds, while the cell pellets exhibited limited chondrogenic differentiation in the outer layer and necrosis in the core area. This could be related to the fact that the scaffolds provide a solid surface for cell attachment while allowing better penetration of fluid flows carrying the oxygen and nutrient for the developing matrix in vitro. Moreover, the form of cell-scaffold system might not only mimic the condensation phase directly but also provide some physical stimulus and biomechanical cues related to the cellular environments in vivo [33]. During the in vivo experiment, the implanted cell pellets could unfortunately not be retrieved. In a previous study the pellets were retrievable when a nude mouse model was used [4]. This time we used nude rats and the small pellets might have fused with the thick subcutaneous adipose tissue thereby hampering sample retrieval. It is known that the pellets occasionally get lost in an in vivo model due to their small size. Obviously, the scaffold system provided better handling abilities. Besides the aforementioned aspects, various assays and histological processing could also be implemented using this PLGA/PCL scaffold, which makes it a superior model for studying endochondral bone regeneration. It is known that the progression of bone formation in vivo varies when the cartilaginous templates are generated in different developmental stages in vitro before implantation [30]. However, no conclusion has been reached on the length of in vitro chondrogenic priming time because different cell resources or scaffold types were applied in previous studies [4, 34-36]. In this study 4-weeks of in vitro chondrogenic differentiation were applied. Histological staining showed extensively cartilage matrix deposition within the scaffolds and a number of chondrocytes presented hypertrophic morphology. In vivo results demonstrated that 8 weeks of implantation was not sufficient to ossify the complete construct, 76 In Vivo Endochondral Bone Generation on Wet-electrospun Fibers while a comprehensive remodeling of the cartilage has been revealed. The new bone (woven bone) mainly consisted of immature characteristic of bone and uncalcified cartilage. A prolonged in vivo implantation time may result in the complete maturation and eventually the functional bone tissue formation with marrow stroma. On the other hand, more research work is still needed to determine the optimal in vitro priming time to obtain satisfactory bone formation in vivo. From material aspects, it should also be noted that PLGA and PCL, as synthetic polymers, are in general not osteoconductive. In most cases, polymers are combined or coated with osteoconductive components, for instance calcium phosphates and hydroxyapatite (HA), to improve their osteo-compatibility [37]. Previous studies verified that a composite electrospun scaffold based on HA and PCL could stimulate bone formation by promoting cell proliferation and osteoblastic differentiation [38] . Therefore, the addition of osteoconductive components may offer some extra biological cues to influence the bone forming process during endochondral route. Additionally, the modulation of material properties of PLGA/PCL, such as degradation time [39], may also influence the in vivo outcome. It has been found that the incorporation of nano sized HA could slow down the degradation of PLGA/ PCL materials in a HA amount-dependent manner and significantly improved the tissue response during 4-week of subcutaneous implantation [39]. For that reason, it is necessary to address the relationship between the supporting material and endochondral bone formation in future research, so that this PLGA/PCL constructs can be eventually applied for maxillofacial or dental applications, e.g. bone reconstruction for dental implantation and bone defects in non-load bearing sites. 5. Conclusions In this study the new concept of engineering bone tissue via the endochondral route was successfully combined with a 3D PLGA/PCL blend scaffold fabricated using a wet-electrospinning system. After 4 weeks of in vitro chondrogenic priming, a cell infiltration and abundant cartilage matrix deposition by RBMCs could be observed throughout the scaffolds. Eight weeks of in vivo implantation revealed extensive new bone formation through the remodeling of cartilage template. This system can be easily processed for various assays and histological analysis. Therefore, it is more efficient than the cell pellets as a tool to study endochondral bone formation for tissue engineering purposes. 77 4 Chapter 4 References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 78 Langer, R. and Vacanti, J.P., Tissue engineering. Science, 1993. 260(5110): p. 920-6. Caplan, A.I., Review: mesenchymal stem cells: cell-based reconstructive therapy in orthopedics. Tissue Eng, 2005. 11(7-8): p. 1198-211. Meijer, G.J., de Bruijn, J.D., Koole, R., and van Blitterswijk, C.A., Cell-based bone tissue engineering. PLoS Med, 2007. 4(2): p. e9. Farrell, E., Both, S.K., Odorfer, K.I., Koevoet, W., Kops, N., O’Brien, F.J., Baatenburg de Jong, R.J., Verhaar, J.A., Cuijpers, V., Jansen, J., Erben, R.G., and van Osch, G.J., In-vivo generation of bone via endochondral ossification by in-vitro chondrogenic priming of adult human and rat mesenchymal stem cells. BMC Musculoskelet Disord, 2011. 12: p. 31. Farrell, E., van der Jagt, O.P., Koevoet, W., Kops, N., van Manen, C.J., Hellingman, C.A., Jahr, H., O’Brien, F.J., Verhaar, J.A., Weinans, H., and van Osch, G.J., Chondrogenic priming of human bone marrow stromal cells: a better route to bone repair? Tissue Eng Part C Methods, 2009. 15(2): p. 285-95. Mackie, E.J., Ahmed, Y.A., Tatarczuch, L., Chen, K.S., and Mirams, M., Endochondral ossification: how cartilage is converted into bone in the developing skeleton. Int J Biochem Cell Biol, 2008. 40(1): p. 46-62. Jukes, J.M., Both, S.K., Leusink, A., Sterk, L.M., van Blitterswijk, C.A., and de Boer, J., Endochondral bone tissue engineering using embryonic stem cells. Proc Natl Acad Sci U S A, 2008. 105(19): p. 6840-5. Dai, J. and Rabie, A.B., VEGF: an essential mediator of both angiogenesis and endochondral ossification. J Dent Res, 2007. 86(10): p. 937-50. Zhang, L., Su, P., Xu, C., Yang, J., Yu, W., and Huang, D., Chondrogenic differentiation of human mesenchymal stem cells: a comparison between micromass and pellet culture systems. Biotechnol Lett, 2010. 32(9): p. 1339-46. Tare, R.S., Howard, D., Pound, J.C., Roach, H.I., and Oreffo, R.O., Tissue engineering strategies for cartilage generation--micromass and three dimensional cultures using human chondrocytes and a continuous cell line. Biochem Biophys Res Commun, 2005. 333(2): p. 609-21. Pham, Q.P., Sharma, U., and Mikos, A.G., Electrospinning of polymeric nanofibers for tissue engineering applications: a review. Tissue Eng, 2006. 12(5): p. 1197-211. Flemming, R.G., Murphy, C.J., Abrams, G.A., Goodman, S.L., and Nealey, P.F., Effects of synthetic micro- and nano-structured surfaces on cell behavior. Biomaterials, 1999. 20(6): p. 573-88. Sharma, B. and Elisseeff, J.H., Engineering structurally organized cartilage and bone tissues. Ann Biomed Eng, 2004. 32(1): p. 148-59. Sill, T.J. and von Recum, H.A., Electrospinning: applications in drug delivery and tissue engineering. Biomaterials, 2008. 29(13): p. 1989-2006. Yang, X., Yang, F., Walboomers, X.F., Bian, Z., Fan, M., and Jansen, J.A., The performance of dental pulp stem cells on nanofibrous PCL/gelatin/nHA scaffolds. J Biomed Mater Res A, 2010. 93(1): p. 247-57. Spivak, A.F., Dzenis, Y.A., and Reneker, D.H., A model of steady state jet in the electrospinning process. Mechanics Research Communications, 2000. 27: p. 37-42. Pham, Q.P., Sharma, U., and Mikos, A.G., Electrospun poly(epsilon-caprolactone) microfiber and multilayer nanofiber/microfiber scaffolds: characterization of scaffolds and measurement of cellular infiltration. Biomacromolecules, 2006. 7(10): p. 2796-805. Lee, Y.H., Lee, J.H., An, I.G., Kim, C., Lee, D.S., Lee, Y.K., and Nam, J.D., Electrospun dual-porosity structure and biodegradation morphology of Montmorillonite reinforced PLLA nanocomposite scaffolds. Biomaterials, 2005. 26(16): p. 3165-72. Lee, J.B., Jeong, S.I., Bae, M.S., Yang, D.H., Heo, D.N., Kim, C.H., Alsberg, E., and Kwon, I.K., Highly porous electrospun nanofibers enhanced by ultrasonication for improved cellular infiltration. Tissue Eng Part A, 2011. 17(21-22): p. 2695-702. Blakeney, B.A., Tambralli, A., Anderson, J.M., Andukuri, A., Lim, D.J., Dean, D.R., and Jun, H.W., Cell infiltration and growth in a low density, uncompressed three-dimensional electrospun nanofibrous scaffold. Biomaterials, 2011. 32(6): p. 1583-90. In Vivo Endochondral Bone Generation on Wet-electrospun Fibers 21. Yokoyama, Y., Hattori, S., Yoshikawa, C., Yasuda, Y., Koyama, H., Takato, T., and Kobayashi, H., Novel wet electrospinning system for fabrication of spongiform nanofiber 3-dimensional fabric. Materials Letters, 2009. 63(9-10): p. 754-756. 22. Fang, J., Wang, H., Niu, H., Lin, T., and Wang, X., Evolution of fiber morphology during electrospinning. J Appl Polym Sci, 2010. 118: p. 2553-2561. 23. Shim, I.K., Jung, M.R., Kim, K.H., Seol, Y.J., Park, Y.J., Park, W.H., and Lee, S.J., Novel threedimensional scaffolds of poly(L-lactic acid) microfibers using electrospinning and mechanical expansion: Fabrication and bone regeneration. J Biomed Mater Res B Appl Biomater, 2010. 95(1): p. 150-60. 24. Schneider, O.D., Weber, F., Brunner, T.J., Loher, S., Ehrbar, M., Schmidlin, P.R., and Stark, W.J., In vivo and in vitro evaluation of flexible, cottonwool-like nanocomposites as bone substitute material for complex defects. Acta Biomater, 2009. 5(5): p. 1775-84. 25. Kronenberg, H.M., Developmental regulation of the growth plate. Nature, 2003. 423(6937): p. 332-6. 26. Chan, C.K., Chen, C.C., Luppen, C.A., Kim, J.B., DeBoer, A.T., Wei, K., Helms, J.A., Kuo, C.J., Kraft, D.L., and Weissman, I.L., Endochondral ossification is required for haematopoietic stem-cell niche formation. Nature, 2009. 457(7228): p. 490-4. 27. Das, A. and Botchwey, E., Evaluation of angiogenesis and osteogenesis. Tissue Eng Part B Rev, 2011. 17(6): p. 403-14. 28. Kanczler, J.M. and Oreffo, R.O., Osteogenesis and angiogenesis: the potential for engineering bone. Eur Cell Mater, 2008. 15: p. 100-14. 29. Yang, Y.Q., Tan, Y.Y., Wong, R., Wenden, A., Zhang, L.K., and Rabie, A.B., The role of vascular endothelial growth factor in ossification. Int J Oral Sci, 2012. 4. 30. Scotti, C., Tonnarelli, B., Papadimitropoulos, A., Scherberich, A., Schaeren, S., Schauerte, A., Lopez-Rios, J., Zeller, R., Barbero, A., and Martin, I., Recapitulation of endochondral bone formation using human adult mesenchymal stem cells as a paradigm for developmental engineering. Proc Natl Acad Sci U S A, 2010. 107(16): p. 7251-6. 31. Lin, Z., Willers, C., Xu, J., and Zheng, M.H., The chondrocyte: biology and clinical application. Tissue Eng, 2006. 12(7): p. 1971-84. 32. Gawlitta, D., Farrell, E., Malda, J., Creemers, L.B., Alblas, J., and Dhert, W.J., Modulating endochondral ossification of multipotent stromal cells for bone regeneration. Tissue Eng Part B Rev. 16(4): p. 385-95. 33. Mahmoudifar, N. and Doran, P.M., Chondrogenesis and cartilage tissue engineering: the longer road to technology development. Trends Biotechnol, 2012. 30(3): p. 166-76. 34. Pelttari, K., Winter, A., Steck, E., Goetzke, K., Hennig, T., Ochs, B.G., Aigner, T., and Richter, W., Premature induction of hypertrophy during in vitro chondrogenesis of human mesenchymal stem cells correlates with calcification and vascular invasion after ectopic transplantation in SCID mice. Arthritis Rheum, 2006. 54(10): p. 3254-66. 35. Tortelli, F., Tasso, R., Loiacono, F., and Cancedda, R., The development of tissue-engineered bone of different origin through endochondral and intramembranous ossification following the implantation of mesenchymal stem cells and osteoblasts in a murine model. Biomaterials, 2010. 31(2): p. 242-9. 36. Farrell, E., O’Brien, F.J., Doyle, P., Fischer, J., Yannas, I., Harley, B.A., O’Connell, B., Prendergast, P.J., and Campbell, V.A., A collagen-glycosaminoglycan scaffold supports adult rat mesenchymal stem cell differentiation along osteogenic and chondrogenic routes. Tissue Eng, 2006. 12(3): p. 459-68. 37. Kretlow, J.D. and Mikos, A.G., Review: mineralization of synthetic polymer scaffolds for bone tissue engineering. Tissue Eng, 2007. 13(5): p. 927-38. 38. Yang, F., Both, S.K., Yang, X., Walboomers, X.F., and Jansen, J.A., Development of an electrospun nano-apatite/PCL composite membrane for GTR/GBR application. Acta Biomater, 2009. 5(9): p. 3295-304. 39. Ji, W., Yang, F., Seyednejad, H., Chen, Z., Hennink, W.E., Anderson, J.M., van den Beucken, J.J., and Jansen, J.A., Biocompatibility and degradation characteristics of PLGA-based electrospun nanofibrous scaffolds with nanoapatite incorporation. Biomaterials, 2012. 33(28): p. 6604-14. 79 4 Chapter 5 Performance of different three-dimensional scaffolds for in vivo endochondral bone generation Wanxun Yang, Sanne K. Both , Gerjo J. V. M.van Osch, Yining Wang, John A. Jansen and Fang Yang Performance of Different 3D Scaffolds for Endochondral Bone Formation Summary of this chapter In the context of skeletal tissue development and repair, endochondral ossification has inspired a new approach to regenerate bone tissue in vivo via a cartilage intermediate as an osteoinductive template. The aim of this chapter was to investigate the behavior of mesenchymal stem cells (MSCs) in regard to in vitro cartilage formation and in vivo bone regeneration when combined with different three-dimensional (3D) scaffold materials, i.e. hydroxyapatite/tricalcium phosphate (HA/TCP) composite block, polyurethane (PU) foam, poly(lactic-co-glycolic acid)/ poly(ε-caprolactone) electrospun fibers (PLGA/PCL) and collagen I gel. To this end, rat MSCs were seeded on these scaffolds and chondrogenically differentiated in vitro for 4 weeks followed by in vivo subcutaneous implantation for 8 weeks. After in vitro chondrogenic priming, comparable cell amounts and cartilage formation were observed in all types of scaffolds. Nonetheless, the quality and maturity of in vivo ectopic bone formation appeared to be scaffold/material-dependent. Eight weeks of implantation was not sufficient to ossify the entire PLGA/PCL constructs, albeit a comprehensive remodeling of the cartilage had occurred. For HA/TCP, PU and collagen I scaffolds, more mature bone formation with rich vascularity and marrow stroma development could be observed. These data suggest that chondrogenic priming of MSCs in the presence of different scaffold materials allows the establishment of reliable templates for generating functional endochondral bone tissue in vivo without using osteoinductive growth factors. The morphology and maturity of bone formation can be dictated by the scaffold properties. 1. Introduction Since the introduction of tissue engineering concept, cell-based bone regeneration strategies have received much attention [1, 2]. The conventional approach in this field is to combine mesenchymal stem cells (MSCs) with biomaterials/scaffolds followed by directing MSCs towards the bone lineage in vitro, which resembles intramembranous ossification [3]. However, the clinical translation of this approach so far has been unsatisfactory [2]. Lack of sufficient vascular supply, resulting in immediate cell death after implantation, is generally thought to be the cause of failure of the cell-scaffold constructs in patients [4]. Recently, in the context of skeletal tissue development and repair, endochondral ossification has inspired a new approach to regenerate bones [5-7]. In this approach, MSCs seeded on scaffolds are first primed towards the chondrogenic lineage in vitro before implantation. These cartilage intermediates work as osteoinductive templates and transform into bone tissue in vivo [4, 8, 9]. The recent findings suggest 83 5 Chapter 5 that this endochondral ossification approach may be more successful compared to the intramembranous route. It is known that cartilage is well adapted to limited nutrition supply and hypoxic conditions [10]. In addition, during the hypertrophic stage of the endochondral process, the chondrocytes produce substantial amount of vascular endothelial growth factor (VEGF), which plays a critical role in the establishment of vessel in-growth and extracellular matrix remodeling [11]. The morphogenetic signals from the chondrocytes also promote the osteogenic differentiation of local MSCs [12, 13]. Hence, the endochondral bone tissue engineering route may allow for a longer temporal window for vascularization and has potential for engineering larger bone constructs . For the final clinical application of such a strategy, an appropriate scaffold is an indispensable element, which provides a three-dimensional (3D) environment as well as the mechanical support for the cells [14]. Regarding endochondral bone formation, it is a unique transition process from chondrogenesis to osteogenesis and involves several instances of extracellular matrix (ECM) degradation and remodeling. Thus, the scaffolds for endochondral bone tissue engineering should be able to support cells in differentiating into chondrogenic lineage in vitro and meanwhile, also promote the calcification of a cartilage template into bone after in vivo implantation. Cartilage matrix consists of highly hydrated proteoglycan embedded into a type II collagen network. Based on the previous findings, hydrogels made from natural or synthetic polymers have shown to support the phenotype of chondrocyte [15]. On the other hand, bone tissue is more rigid than cartilage and impregnated mainly with hydroxyapatite (HA) and type I collagen. Among the candidate scaffold materials, porous bioceramics and degradable polymers are widely studied because of their mechanical competence and structural feasibility for fluid perfusion and vessel in-growth. Apart from the aforementioned, nanofibrous scaffolds made by electrospinning has also gained research interests in both cartilage and bone tissue engineering due to their morphological resemblance to the natural extracellular matrix [16, 17]. In this study, we aimed at testing the performance of different 3D scaffolds for endochondral bone tissue engineering. To that end, four representative scaffolds (ceramic- or polymer-based) with distinct structural features were selected from the aforementioned materials, including porous hydroxyapatite/tricalcium phosphate (HA/TCP), porous polyurethane (PU), fibrous poly(lactic-co-glycolic acid) (PLGA) electrospun scaffolds and collagen I hydrogel. Rat primary MSCs were seeded on these scaffolds and chondrogenically primed for 4 weeks in vitro. The formation of cartilage templates was characterized by matrix quantification and gene analysis. After 4 weeks of in vitro culture, the chondrogenically primed cellscaffold constructs were subcutaneously implanted in rats for 8 weeks and then 84 Performance of Different 3D Scaffolds for Endochondral Bone Formation the spatial pattern of bone formation and cartilage remains were assessed in each construct in terms of quality and quantity via immunohistological stainings and polarized microscopy. 2. Materials and Methods 2.1 Scaffold preparation Four different types of scaffold materials were used in this study: Porous sintered HA/TCP composite (CAM bioceramics B.V., Leiden, the Nethelands) was obtained in a cubic shape of 3×3×3 mm3 and consisted of HA and TCP with a 75/25 ratio. The volumetric porosity was 75% (provided by the manufacturer). HA/TCP (Fig. 1a) showed a macro-porous structure with pore sizes ranging from 300-800 mm. Porous PU scaffold was fabricated by in situ polymerization and simultaneous foaming as previously described [18, 19]. The scaffold displayed a similar porous structure as HA/TCP (Fig. 1b) and the porosity is 80% [18]. Figure 1. SEM micrographs of HA/TCP (a), PU (b) and PLGA/PCL (c) scaffolds. (Scale bar in a and b: 100µm; c: 10µm) (Micrograph of collagen I scaffold is not shown because the structure could not be preserved after freeze-drying.) Fibrous PLGA (Purasorb® PDLG 8531, Purac Biomaterials BV, Gorinchem, the Netherlands) combined with poly(ε-caprolactone) (PCL, LACTEL® Absorbable Polymers, DURECT Corporation , Pelham, USA) electrospun scaffold was made by a wet-electrospinning technique (Esprayer ES-2000S, Fuence Co., Ltd, Tokyo, Japan) as previously described [20]. The obtained scaffold (Fig. 1c) displayed an uncompressed structure with an average fiber diameter of 1.98 ± 0.51mm and a porosity of 99% [20]. 85 5 Chapter 5 Collagen hydrogel was prepared from type I collagen (rat tail) (BD Biosciences, Bedford, USA) with the final concentration at 6 mg/ml according to the manufacturer’s instructions. Both PLGA/PCL and PU scaffolds were punched into a disk shape for the following experiments with a diameter of 6 mm and a thickness of 2 mm. Collagen I gel was shaped to the same size during the cell seeding procedure. The morphology of the scaffolds was observed by scanning electron microscopy (SEM; JEOL6340F, Tokyo, Japan) after being sputter-coated with gold-platinum. 2.2 Cell isolation, seeding and culture MSCs were isolated from 6-week-old male Fischer rats (12 rats in total) (Charles River) after the approval from Radboud University Nijmegen Animal Ethics Committee (Approval no: RU-DEC 2011-142). Briefly, two femora of each rat were removed and the epiphyses were cut off. Rat MSCs were flushed out of the remaining diaphyses using the proliferation medium through an 18G needle (BD Microlance, Drogheda, Ireland). The proliferation medium consisted of alpha Minimal Essential Medium (αMEM; Gibco, Grand Island, USA) supplemented with 10% fetal bovine serum (FBS; Sigma, St. Louis, USA), 100 U/ml penicillin, 100μg/ml streptomycin (Gibco), 50 μg/ml l-ascorbic acid (Sigma) and 10nM dexamethasone (Sigma). The flush-out was cultured for 2 days in a humidified incubator (37 °C, 5% CO2), after which the medium was refreshed to remove non-adherent cells. The cells were cultured for an additional 3 days followed by detachment using trypsin/ EDTA (0.25% w/v trypsin, 0.02% w/v EDTA; Sigma) and counted. Before cell seeding, the scaffolds were sterilized either by autoclave (PU and HA/TCP) or incubation in 70% ethanol for 2 hours (PLGA/PCL) then washed with PBS. Subsequently, all scaffolds were soaked in the proliferation medium overnight. During cell loading, the scaffolds were incubated in the cell suspension with a concentration of 1 × 106 cells/ml (4 scaffolds per 1 ml of cell suspension) and gently rotated for 3 hours. Successively, the scaffolds were placed in nonadherent tissue culture plates and the unattached cells from the suspension were collected, centrifuged and reseeded onto the scaffolds to ensure a high cell loading efficiency. For collagen I scaffolds, the gel was made aseptically according to the manufacturer’s instructions, mixed with cells, and set in cell culture inserts (ThinCertTM for 24-well plate, Greiner Bio-One, Germany). A concentration of 250,000 cells per insert (with 120µl gel) was used to obtain a disc shaped cell/ collagen gel construct with a diameter of 6 mm and a thickness of 2 mm. All cell-scaffold constructs were cultured for 1 week in the proliferation medium. Thereafter, the medium was changed to the chondrogenic medium, consisting of 86 Performance of Different 3D Scaffolds for Endochondral Bone Formation high-glucose DMEM (Gibco), 1% FBS, 100 U/ml penicillin, 100 μg/ml streptomycin, 50 μg/ml l-ascorbic acid, 100 mM of sodium pyruvate (Gibco), 40μg/ml l-proline (Sigma), 1:100 insulin-transferrin-selenium (ITS; BD Biosciences, Bedford, USA), 100nM dexamethasone, 10 ng/ml of transforming growth factor beta-2 (TGF-β2; R&D systems, Minneapolis, USA) and 100 ng/ml bone morphogenetic protein 2 (BMP-2; BD biosciences). The constructs were cultured for 4 weeks and the medium was refreshed twice a week. Afterwards the samples were divided into two groups, one for in vitro characterization and the other for in vivo implantation. 2.3 In vitro tests 2.3.1 GAG content The amount of GAG in each sample was detected by a sulfated glycosaminoglycan (sGAG) assay (n=3). After being washed twice in PBS, all samples were digested in proteinase K solution (1 mg/ml proteinase K in 50 mM Tris with 1 mM EDTA, 1 mM iodoacetamide, and 10 μg/ml pepstatin A) (all reagents from Sigma) for 20 hours at 56°C. The GAG content was determined by using dimethylmethylene blue (Polysciences Inc., Warrington, USA), with chondroitin sulfate (Sigma) as standard control. Measurements of absorption were performed at 530 nm and 570 nm using an ELISA reader (Bio-Tek Instruments Inc., Winooski, USA). The results were calculated according to the previously described method [21]. For each sample, the GAG content was normalized by the amount of DNA as described in 2.3.2. 2.3.2 DNA content To evaluate the seeding efficiency on each type of scaffolds and the cell number before the implantation, DNA content of the seeded cells was measured after 24 hours of seeding and 28 days of chondrogenic differentiation, respectively. PicoGreen assay (Quant-iTTM PicoGreen® dsDNA kit, Invitrogen, Oregon, USA) was performed using the same cell extract solution as treated for the GAG assay. A DNA standard curve was used to quantify the amount of DNA in each sample and the results were measured using a fluorescence microplate reader (BioTek Instruments Inc.) with an excitation wavelength at 485nm and an emission wavelength at 530nm. 2.3.3 RNA extraction and qPCR analysis Total RNA was extracted using Trizol method after 4 weeks of chondrogenic culture (n=4). Briefly, scaffolds with cells were washed with PBS and cut into small pieces before adding 1 ml of Trizol (Invitrogen). The cell extract was collected, mixed with chloroform and centrifuged. Only the upper aqueous phase was collected and mixed with equal amount of isopropanol. After 10 min of incubation at room 87 5 Chapter 5 temperature, the mixture was centrifuged and washed twice with 75% alcohol. Thereafter, the obtained RNA pellet was dissolved in RNase free water and the RNA concentration was measured with spectrophotometer (NanoDrop 2000, Thermal scientific, Wilmington, USA). Table 1. Primer sequences used for qPCR Forward (5’ → 3’) Reverse (5’ → 3’) Col2α1 GACGCCACGCTCAAGTCGCT CGCTGGGTTGGGGTAGACGC Col10α1 GGGCCCTATTGGACCACCAGGTA CCGGCATGCCTGTTACCCCC Acan CATTCGCACGGGAGCAGCCA TGGGGTCCGTGGGCTCACAA Vegfa ACTCATCAGCCAGGGAGTCT GGGAGTGAAGGAGCAACCTC Gapdh CGATGCTGGCGCTGAGTAC CGTTCAGCTCAGGGATGACC First strand cDNA was reverse transcribed from RNA using the SuperScript® FirstStrand Synthesis System kit (Invitrogen). Afterwards cDNA was further amplified and the expression of specific genes was quantified using IQ SYBR Green Supermix PCR kit (BioRad, Hemel Hempstead, United Kingdom) in a real-time PCR (BioRaD, CFX96™ real-time system) (n=4). Chondrogenic differentiation related gene markers were evaluated, including collagen type II (Col2α1), collagen type X (Col10α1), aggrecan (Acan), and VEGF (Vegfa) (Table 1). The expression levels were analyzed and compared to the housekeeping gene glyceraldehydes 3-phosphate dehydrogenase (Gapdh). The specificity of the primers was confirmed separately before the real-time PCR reaction. The expression of the tested genes was calculated via the 2-∆∆Ct method [22] using HA/TCP as the reference group. 2.3.4 Histological analysis All the samples were fixed in 10% phosphate buffered formalin for 1 hour before embedding. The in vitro PLGA/PCL and collagen I scaffolds were embedded in paraffin (n=3). The microtome sections with 6 μm of thickness were stained with 0.4% thionin (Sigma) after deparaffinization in xylene and rehydration through graded ethanol. Simultaneously, HA/TCP and PU scaffolds were embedded in methylmetacrylate (MMA, L.T.I., Bilthoven, the Netherlands) because of their rigid character. The samples were sectioned to 15μm using a microtome equipped with diamond blade (Leica SP1600, Leica microsystem, Wetzlar, Germany) and then stained with 0.4% thionin solution. 88 Performance of Different 3D Scaffolds for Endochondral Bone Formation 2.4 In vivo implantation After 4 weeks of in vitro culture, the cell-scaffold constructs were implanted subcutaneously in ten 8-week-old male nude rats (Crl:NIH-Foxn1mu, Charles River). The protocol was approved by the Animal Ethical Committee of the Radboud University Nijmegen Medical Centre (Approval no: RU-DEC 2011-142) and the national guidelines for the care and use of laboratory animals were applied. All rats received analgesic before and after the surgery. The surgical operations were performed under general inhalation anesthesia with a combination of isoflurane, nitrous oxide, and oxygen. The back of the rats was shaved and disinfected with povidone-iodine. Subsequently, four small longitudinal incisions were made. Lateral to each incision, a subcutaneous pocket was created. Each rat received a cell-scaffold construct from each of the four materials and the constructs were randomized. After placing the samples, the skin was closed using staples. After 8 weeks, the rats were euthanized by CO2-suffocation for sample collection. 2.5 Histological analysis of the in vivo samples All in vivo samples were fixed in 10% phosphate buffered formalin for 30 hours and dehydrated through graded ethanol before embedding either in MMA or paraffin. MMA embedded samples (n = 5) were sectioned to 15μm using a microtome with diamond blade (Leica SP1600, Leica microsystem, Wetzlar, Germany) and then stained with methylene blue and basic fuchsin (both reagents from Merck). Prior to paraffin embedding, the fixed samples (n = 5) were decalcified by 10% EDTA and the decalcification process was monitored via X-ray. Five μm-thick sections were made from each sample and used for hematoxylin/eosin (HE), collagen II and tartrate-resistant acid phosphatase (TRAP) staining. For collagen II staining, sections were treated with 0.1% pronase and 1% hyaluronidase for antigen retrieval (both from Sigma). Afterwards the sections were incubated for 1 h at room temperature with mouse monoclonal antibody against collagen type II (II-II6B3 antibody, 1:100; Developmental Studies Hybridoma Bank, Iowa City, USA) linked with a biotin-labeled secondary antibody (1:500, Jackson Immuno Research, Newmarket, England). Then alkaline phosphatase-conjugated streptavidin (1:50, Biogenex HK-321-UK) was used as the third antibody and the activity was visualized by using a new fuchsin substrate (Sigma). An isotype immuno-globulin G1 monoclonal antibody was used as negative control. Counterstaining was performed with Gill’s hematoxylin (Sigma). For TRAP staining, sections were first incubated in TRIS/MgCl2-buffer for 1 hour. Afterwards they were stained with acid phosphatase substrate solution at 37 °C for 30 min, which contained Naphthol AS-BI phosphate, sodium nitrite, N, N- dimethylformamide, potassium sodium tartrate and pararosaniline (all reagents from sigma). 89 5 Chapter 5 2.6 Polarized light microscopy The in vivo MMA cross-sections were observed under polarized light using a light microscope (DM RBE Leica, Bensheim, Germany) in which the polarizer and analyzer were fixed perpendicularly to each other (cross-polarized light). 2.7 Statistic analysis The data were analyzed using one-way ANOVA followed by Tukey’s Multiple Comparison Test (GraphPad Prism 5, GraphPad Software Inc., San Diego, USA). Results are reported as mean values and standard deviation. Differences were considered statistically significant at p < 0.05. 3. Results 3.1 Cell seeding efficiency After 24 hours of cell seeding, DNA content of the seeded cells on each group was measured to assess the seeding efficiency (Fig. 2a). Highest cell number was found on collagen I group, as all of the cells were encapsulated in the gel during cell seeding. Comparable amount of cells was detected on PU and PLGA/PCL scaffolds, which was around 65% of the total cell number seeded onto the scaffolds. In comparison, lowest cell number was shown on HA/TCP scaffolds, with the seeding efficiency of around 25%. 3.2 DNA and GAG content pre-implantation After 1 week of proliferation and 4 weeks of chondrogenic differentiation in vitro, comparable amounts of DNA could be found in HA/TCP, PLGA/PCL and collagen Figure 2. DNA (a and b) and GAG (c) content on different scaffolds. DNA content after 24 hours of seeding (a) was measured to evaluate the seeding efficiency. GAG content (c) was measured after 4 weeks of chondrogenic priming to assess the cartilage formation and normalized by the amount of DNA (b). ** indicates statistically significant difference at p<0.01; error bars represent standard deviation (n=3). 90 Performance of Different 3D Scaffolds for Endochondral Bone Formation I scaffolds, which was significant higher than that on the PU scaffolds as shown in Fig. 2b. Likewise, the normalized GAG content on the aforementioned three scaffolds was also comparable, but the PLGA/PCL constructs showed significant higher GAG/DNA ratios than the PU samples (Fig. 2c). 5 Figure 3. Gene expression of aggrecan (a), collagen II (b), collagen X (c) and VEGF (d) after in vitro chondrogenic priming. The gene expression level of the cells seeded on HA/TCP scaffold was used for normalization. * indicates statistically significant difference at p<0.05; ** indicates statistically significant difference at p<0.01; error bars represent standard deviation (n=4). The red dash line indicates the gene expression level of the reference group (HA/TCP). 3.3 Expression of chondrogenic genes pre-implantation After 4 weeks of chondrogenic differentiation in vitro, a 2-fold higher level of aggrecan expression was observed for HA/TCP group compared to collagen I group, while there was no difference detected among HA/TCP, PU and PLGA/PCL groups (Fig. 3a). Similar result was also found for VEGF expression that HA/TCP group showed significantly higher level of VEGF compared to collagen I group (Fig. 3d). For collagen II gene expression, all groups showed a similar level (Fig. 3b). In 91 Chapter 5 contrast, a significant higher expression of collagen X was detected for PLGA/PCL groups (Fig. 3c). Figure 4. In vitro chondrogenic differatiation. The cartilage matrix deposition (stain purple) in the scaffolds was visualized by thionin staining after 4 weeks of in vitro culture. On HA/TCP (a and e) and PU (b and f), the cartilage matrix was formed in the pores of the scaffolds. The cells exhibited typical chondrocyte like morphology and hypertrophy could be observed, whereby the lacunae of the cells were enlarged and the matrix was reduced (e and f). The shrinkage of the PLGA/PCL (c and g) and collagen I (d and h) scaffolds during the histological processing compromised the morphology of the cells. However, obvious cartilage matrix deposition was displayed in both scaffolds and a homogeneous distribution of cells could be observed. (S: scaffolds) 92 Performance of Different 3D Scaffolds for Endochondral Bone Formation 3.4 Histological evaluation pre-implantation To visualize the cartilage matrix deposition in the scaffolds, histological sections were made and stained with thionin after 4 weeks of chondrogenic culture. For both HA/TCP and PU, as shown in Fig. 4a and b, the cartilage matrix was formed in the pores of the scaffolds and stained purple. The cells had a rounded or polygonal shape and formed isogenous groups, exhibiting typical chondrocyte like morphology (Fig. 4e and f). Furthermore, the majority of the cells had reached the hypertrophic stage whereby the lacunae were enlarged and the matrix was reduced. The shrinkage of the fibrous scaffold during the histological processing compromised the morphology of the cells on PLGA/PCL. Nevertheless, a substantial GAG content was displayed in the scaffolds stained strongly in purple, representing the cartilage matrix (Fig. 4c). The PLGA/PCL fibers were embedded within the matrix and a homogeneous distribution of cells could be observed throughout the fibrous structure (Fig. 4g). Similarly, the collagen I scaffolds also showed cartilage matrix deposition although here the shrinkage was less, while the total amount appeared to be lower than PLGA/PCL scaffolds. (Fig. 4d and h). 3.5 Histological analysis of in vivo bone formation All animals recovered uneventfully from the surgical procedure and remained in good health. No signs of wound complications were observed post-operatively. After 8 weeks of implantation, all samples were retrieved and an obvious increase in hardness of PLGA/PCL, collagen I and PU scaffolds could be sensed by hand. Macroscopic signs of inflammation or adverse tissue responses were absent for all retrieved samples. To evaluate the bone forming quality and quantity, different histological staining methods were used for either MMA (Fig. 5) or paraffin (Fig.6) embedded samples. Bone formation was found in all HA/TCP, PU and collagen I constructs, and nine (out of ten) PLGA/PCL scaffolds. For the bone-formed samples within each group, the bone forming quality was similar as observed from the histological stainings. Due to the deformation of PU and PLGA/PCL scaffolds during the tissue processing procedures, the original material area could not be defined to quantify the ratio of bone formation within the scaffolds. 3.5.1 HA/TCP scaffolds As shown in Fig. 5a, substantial bone formation was present at both the periphery and inside of the scaffolds. Within the macro-pores, the structure of a marrow cavity filled with hematopoietic cells and adipocytes could be observed, surrounded by concentric deposition of bone matrix. Moreover, plenty of blood vessels 93 5 Chapter 5 penetrated randomly in these bone forming areas (Fig. 6a and b). Osteocytes were embedded in the bone matrix, showing relatively round shapes. Collagen II staining confirmed that only a small amount of cartilage (stained pink) remained, which was occasionally found in the small pores of a few samples where cells kept the chondrocyte like shape without breaking down the surrounding matrix (Fig. 6c). Figure 5. Overview of in vivo bone formation. MMA sections of HA/TCP (a), PU (b), PLGA/PCL (c) and Collagen I (d) samples after 8 weeks of implantation were stained with methylene blue and basic fuchsin. Bone formation (stained red) displays in different spatial arrangements on each type of scaffold. 3.5.2 PU scaffolds PU scaffold underwent deformation during the histological processing. Bone formation was observed in the scaffolds, especially in the pores located in the peripheral area (Fig. 5b). Surrounded by the bone tissue, a marrow-like structure or a loosely organized fibrous tissue-like structure was detected in these pores together with blood vessel in-growth (Fig. 6d and e). The amount of bone formation in the central pores was relatively low compared to HA/TCP scaffolds. Chondrocytes or cartilage like structure could not be seen at this time point. Collagen II staining revealed a slight amount of tissue stained light pink, indicating the replacement of in vitro cartilage with bone tissue was nearly completed (Fig. 6f). 94 Performance of Different 3D Scaffolds for Endochondral Bone Formation 3.5.3 PLGA/PCL scaffolds Extensive cartilage remodeling could be observed inside of PLGA/PCL scaffolds. Islets of bone tissue were scattered amongst the remnants of scaffold material, 5 Figure 6. Histological analysis of in vivo bone formation on scaffolds. After 8 weeks of implantation, HE (left and middle column) staining showed that the bone formation on HA/TCP (a), PU (d) and Collagen I (j) scaffolds was associated with bone marrow development while on PLGA/PCL scaffold (g) the bone tissue displayed a woven bone like structure without concentrically organized structure and marrow formation. In the magnified images of the black squared box (middle column), osteocytes were embedded in the bone matrix and blood vessels penetrated in the bone forming areas on all scaffolds. Collagen II staining (right column) confirmed significant remaining of cartilage matrix (red color) on PLGA/PCL (i) while the remodeling of in vitro cartilage into bone tissue was nearly completed on HA/TCP (c), PU (f) and Collagen I (l) scaffolds. The black arrow indicates a blood vessel and the arrow head in the white magnified box indicates embedded electrospun fibers. (B: bone; BM: bone marrow; S: scaffold; C: cartilage) 95 Chapter 5 indicating the onset of the newly formed bone. Concentrically organized bone tissue with marrow cavity structures could not be observed (Fig. 5c). HE staining Figure 7. Histological analysis of in vivo osteoclast activity. After 8 weeks of in vivo implantation, the activity of osteoclasts (stain red) was visualized by TRAP staining. For HA/TCP, PU and collagen I scaffolds, osteoclast activity was positive (Fig. 7a, b and d) which scattered along the edge of bone matrix (Fig. 7e, f and h). In comparison, an extensive level of osteoclast activity was observed throughout the PLGA/PCL scaffolds (Fig. 7c), which was not only located along the edge of new bone formation, but also present in the middle area of the bone/cartilage mixed matrix (Fig. 7g). 96 Performance of Different 3D Scaffolds for Endochondral Bone Formation revealed that the bone tissue displayed a woven bone like structure with vascular invasion. A large number of immature osteocytes were embedded in the matrix and randomly arranged (Fig. 6g and h). Interestingly, the remained PLGA/PCL fibers of the scaffolds were also embedded in the matrix as indicated in Fig. 6h. Furthermore, the collagen II staining (Fig. 6i) confirmed the existence of the remaining cartilage matrix (red color). Compared to the PLGA/PCL constructs before implantation (Fig. 3c), the in vivo samples contained considerably reduced amount of cartilage matrix. However, the newly formed bone still resembled a mixture of cartilage and calcified cartilage tissue or bone. 3.5.4 Collagen scaffolds Bone formation with marrow development was observed in collagen I gel. An intact bone layer was displayed along the fibrous capsule. Trabecular bone-like structure, together with a great number of adipocytes and hematopoietic cells, was distributed in the middle part (Fig. 5d). Blood vessels penetrated in the stroma and were also present inside the bone tissue (Fig. 6j and k). There was no remaining cartilage matrix visualized by a collagen II staining which indicated the complete replacement of in vitro cartilage by new bone tissue (Fig. 6l). 3.6 Histological analysis of osteoclast activity The activity of osteoclasts was visualized by TRAP staining. After 8 weeks of in vivo implantation, osteoclast activity was positive on HA/TCP, PU and collagen I scaffolds (Fig. 7), which scattered along the edges of bone matrix. In comparison, an extensive level of osteoclast activity was observed throughout the PLGA/PCL scaffolds, which was not only located along the edges of new bone formation, but also present in the middle area of the bone/cartilage mixed matrix. 3.7 Polarized light observation The collagen bundle arrangement of the newly formed bone tissue on each scaffold was examined under the polarized light on the basis of its birefringent character. As shown in Fig. 8, the collagen bundles within the bone forming areas which were oriented transversely to the direction of the light propagation were visualized in bright pink under the polarized light. In HA/TCP (Fig. 8a and e), PU (Fig. 8b and f) and collagen I scaffolds (Fig. 8d and h), a parallel pattern of the collagen fibril arrangement could be observed partially in the bone forming regions, indicating the initial organization of the lamellar bone. However, in the PLGA/PCL scaffolds (Fig. 8c and g), the bone matrix displayed a random spatial organization, as shown by an isotropic and un-polarized structure. 97 5 Chapter 5 Figure 8. Observation of in vivo bone formation under polarized light. Newly formed bone tissue in each scaffold (embedded in MMA) was stained with methylene blue / basic fuchsin (a-d) and examined under the polarized light (e-h). The bone tissue was stained red (a-d) and the collagen bundles which were oriented transversely to the direction of the light propagation were distinguished in bright pink under the polarized light (e-h). In HA/TCP (a and e), PU (b and f) and collagen I (d and h) scaffolds, a concentric and parallel pattern of the collagen fibril arrangement could be observed partially in the bone forming regions. For PLGA/PCL scaffolds (c and g), the bone matrix displayed an isotropic organization. (Scale bar: 50µm) 98 Performance of Different 3D Scaffolds for Endochondral Bone Formation 4. Discussion An appropriate scaffold material is an essential component in bone tissue engineering as it regulates the cellular behaviors, such as cell proliferation and differentiation, on both chondrogenic and osteogenic pathways [23, 24]. In this study, we aimed to evaluate and compare the efficacy of four scaffolds for bone tissue regeneration via the endochondral ossification, which were porous HA/TCP composites, porous PU foam, PLGA/PCL electrospun fibers and collagen I gel. Our results showed different cell seeding efficiencies on the tested scaffolds due to their varied structural features, among which the HA/TCP scaffolds had the lowest initial cell number. Nevertheless, 1-week culture in proliferation medium ensured a comparable amount of cells present on these scaffolds prior to the implantation, as the low-serum chondrogenic culture condition would principally not stimulate the cell proliferation. The results also indicated that, all of the selected scaffolds supported cell growth and proliferation. Similarly, comparable cartilage formation in the scaffolds was also shown before the in vivo implantation. However, cells cultured on the PLGA/PCL scaffolds expressed a significantly higher level of collagen X compared to those on the other types of scaffolds. In vivo results revealed that, 8 weeks of implantation was sufficient to ossify nearly the complete constructs of HA/TCP, PU and collagen constructs and form a hematopoietic bone marrow, while the bone formation on PLGA/PCL scaffolds displayed a immature structure with a random spatial organization of collagen bundle and obvious cartilage remaining. Furthermore, an extensive level of osteoclast activity present throughout the PLGA/PCL scaffold also demonstrated a highly active cartilage and bone remodeling in PLGA/PCL compared to the other three scaffolds. These results indicated that, although the cells on PLGA/PCL might have an earlier hypertrophic stage in vitro, the cartilage remodeling and vessel invading in PLGA/PCL was less efficient in vivo compared to the other scaffolds. Hypertrophic cartilage remodeling is an critical process to initialize the successive endochondral bone formation and such progression is closely related to the reciprocal interaction between the cells and natural ECM [25]. Cellular products, e.g. proteinases and VEGF, modify the type X collagen-rich ECM and promote vessel formation; ECM also releases sequestered growth factors and cytokines to regulate the endochondral commitment of the cells [26]. Hypertrophic cartilage remodeling usually couples the expression of VEGF [27]. Surprisingly, a significantly higher level of collagen X expression on PLGA/PCL scaffolds did not result in an elevated VEGF expression of the cells. This may consequently lead to the formation of less mature bone in PLGA/PCL scaffolds compared to the other scaffolds. Moreover, a comparable endochondral bone formation was observed on HA/TCP and collagen 99 5 Chapter 5 I gel, although significant difference of VEGF expression was shown for these two scaffolds prior to the implantation. Therefore, apart from the natural ECM, the scaffolds, acting as a synthetic ECM, also played an essential role in directing the cells to form endochondral bone in vivo. Noticeably, our data demonstrated that the structure and maturity of the endochondral bone is material/scaffold dependent. Among various scaffold properties, geometry and composition played important roles. In vivo, higher percentage of porosity allows for better cell recruitment and vascularization, which leads to improved osteogenesis [28]. On top of that, scaffold pore size can dictate the structure and the mode of the newly formed bone (intramembranous or endochondral) [24]. In an in vivo study of BMP-2 induced osteogenesis and chondrogenesis via honey-comb-shaped HA scaffolds, small diameter ‘tunnels’ (90120µm) favored chondrogenesis followed by osteogenesis, while large diameter ‘tunnels’ (350µm) favored intramembranous bone formation [24, 29]. Regarding the bone structure, increased trabeculae formation appeared to be supported when the scaffolds have smaller pores [30]. Similarly, in our study, more lamellar-like bone structure with bone marrow formation was generated on HA/TCP and PU scaffolds compared to PLGA/PCL. This might be related to the larger pore size of HA/TCP and PU scaffolds, which can facilitate more oxygen and nutrient diffusion and allow easier vessel in-growth during cartilage/bone remodeling. In comparison, the small pore size of PLGA/PCL scaffolds might have created a hypoxic environment which favored the cartilage maintenance [28]. Furthermore, the confined porous structure of HA/TCP and PU can concentrate collagen fibers within the pores and triggered their self-assembly, thereby forming the lamellar like bone [31]. Conversely, in fibrous PLGA/PCL scaffolds, new collagen fibers formed by the cells were distributed in a random manner, resulting mostly in woven bone [31]. Thus, the specific design of the scaffold can drive the arrangement of collagen fibers and consequently target bone formation. Bone in-growth to the central part of the HA/TCP scaffold had a relatively higher volume compared to the PU scaffolds, even though HA/TCP scaffolds were a little thicker. The main reason for the difference of bone in-growth might be that, unlike PU, HA and TCP are osteoconductive [32, 33]. This osteoconductive property can promote the attachment of bone-forming cells from the host as well as cell migration and vessel formation [14]. Both PU and HA/TCP are slowly degrading materials which were unlikely to degrade after 8 weeks of implantation [19, 34] . However, compared to PU scaffolds, HA/TCP could actively release calcium and phosphate ions which can influence the local ionic concentration to trigger and provide the source for the deposition of calcium in the cartilage matrix for subsequent remodeling [34]. With respect to PLGA/PCL scaffolds and collagen I gel, 100 Performance of Different 3D Scaffolds for Endochondral Bone Formation the naturally derived collagen I component resembles the organic composition of natural bone and is considered more osteoconductive than the synthetic polymers, i.e. PLGA and PCL [35], which have also contributed to the differences of the bone formation between these two scaffolds. For in vivo bone generation, the origin of bone cells and the ossification type are closely related to the nature and commitment of the seeded cells. Previous results have suggested a tendency of MSCs to undergo endochondral ossification when implanted in vivo. Hartman et al. seeded rat MSCs onto porous HA/TCP scaffold to investigate ectopic bone formation via the intramembranous approach, while the onset of bone tissue was always surrounded by hypertrophic cartilage-like cells [36]. Tortelli et al. combined murine MSCs or mature osteoblasts with porous ceramic scaffolds and subsequently implanted them ectopically in mice [37]. New vascularized bone was formed through the activation of the endochondral ossification process in the MSC-scaffold complex. Conversely, osteoblasts directly formed bone via an intramembranous ossification, without any signs of vascularization [37]. Aforementioned results indicated that, activating MSCs toward an endochondral ossification route is presumably in line with their natural developmental pattern and provides significant advantages in generating vascularized bone tissue. Our results demonstrated the tendency of in vitro chondrogenically induced MSC to ossify in a non-chondrogenic environment in vivo, which is also consistent with some previous studies [15, 38, 39]. When combined with various scaffolds, scaled-up bone formation with the structure and functionality comparable to that of native bones could be achieved in vivo without the use of any osteoinductive growth factors, e.g. BMP-2, at the implant site. Among the tested scaffold materials, HA/ TCP block and collagen I gel supported most mature bone formation within the entire constructs. Regarding the clinical applications, Collagen I gel is appealing for irregularly shaped defects and its fast degradation allows rapid turnover of the new bone tissue in the implantation site. HA/TCP scaffold is more suitable for the implantation in the load-bearing areas, which can provide initial structural stability. Nevertheless, considering the ceramic scaffolds are brittle, the mechanical property of HA/TCP needs to be further modified, for instance by incorporating polymeric components. Some limitations of our study have to be addressed: i) the evaluation was only performed at one time point of 8 weeks after implantation, ii) the scaffolds were in a relatively small size which might favor cell invasion and vascularization, iii) the implantation was only performed in a subcutaneous environment where the mechanical stimuli are not involved, and iv) due to the differences in composition and structure of the selected scaffolds, the specific parameters of a material scaffold that dominate the endochondral bone forming process could not be 101 5 Chapter 5 concluded from the current study. More completed understanding will be obtained if all structural parameters of the scaffolds are kept constant and meanwhile only the composition is varied, or vice versa. Further investigations are warranted to dynamically monitor the endochondral bone forming process in an orthotopic implantation site. Furthermore, systematical investigations are also needed to address how this process is driven by the parameters of the supporting materials, so that more appropriate scaffolds can be selected and tailored to target largescale bone regeneration. 5. Conclusion In this study, chondrogenic priming rat MSCs in the presence of different 3D scaffold materials followed by subcutaneous implantation allowed us to establish a reliable scaffold-based heterotopic model for endochondral bone formation. The bone quality and maturity was scaffold/material-dependent. Eight weeks of implantation was not sufficient to ossify the entire PLGA/PCL constructs, while a comprehensive remodeling of the cartilage has occurred. HA/TCP, PU and collagen I scaffolds supported full bone formation with rich vascularity and marrow stroma development. This study revealed that bone could be generated heterotopically via the endochondral pathway in absence of osteoinductive growth factors and scaffold properties could dictate the morphology and maturity of endochondral bone formation. Continued research in this direction would open the attractive possibility to target large-scale bone regeneration. References 1. 2. 3. 4. 5. 6. 102 Langer, R. and Vacanti, J.P., Tissue engineering. Science, 1993. 260(5110): p. 920-6. Meijer, G.J., de Bruijn, J.D., Koole, R., and van Blitterswijk, C.A., Cell-based bone tissue engineering. PLoS Med, 2007. 4(2): p. e9. Caplan, A.I., Review: mesenchymal stem cells: cell-based reconstructive therapy in orthopedics. Tissue Eng, 2005. 11(7-8): p. 1198-211. Farrell, E., van der Jagt, O.P., Koevoet, W., Kops, N., van Manen, C.J., Hellingman, C.A., Jahr, H., O’Brien, F.J., Verhaar, J.A., Weinans, H., and van Osch, G.J., Chondrogenic priming of human bone marrow stromal cells: a better route to bone repair? Tissue Eng Part C Methods, 2009. 15(2): p. 285-95. Farrell, E., Both, S.K., Odorfer, K.I., Koevoet, W., Kops, N., O’Brien, F.J., Baatenburg de Jong, R.J., Verhaar, J.A., Cuijpers, V., Jansen, J., Erben, R.G., and van Osch, G.J., In-vivo generation of bone via endochondral ossification by in-vitro chondrogenic priming of adult human and rat mesenchymal stem cells. BMC Musculoskelet Disord, 2011. 12: p. 31. Farrell, E., O’Brien, F.J., Doyle, P., Fischer, J., Yannas, I., Harley, B.A., O’Connell, B., Prendergast, P.J., and Campbell, V.A., A collagen-glycosaminoglycan scaffold supports adult rat mesenchymal stem cell differentiation along osteogenic and chondrogenic routes. Tissue Eng, 2006. 12(3): p. 459-68. Performance of Different 3D Scaffolds for Endochondral Bone Formation 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. Janicki, P., Kasten, P., Kleinschmidt, K., Luginbuehl, R., and Richter, W., Chondrogenic preinduction of human mesenchymal stem cells on beta-TCP: enhanced bone quality by endochondral heterotopic bone formation. Acta Biomater, 2010. 6(8): p. 3292-301. Jukes, J.M., Both, S.K., Leusink, A., Sterk, L.M., van Blitterswijk, C.A., and de Boer, J., Endochondral bone tissue engineering using embryonic stem cells. Proc Natl Acad Sci U S A, 2008. 105(19): p. 6840-5. Scotti, C., Piccinini, E., Takizawa, H., Todorov, A., Bourgine, P., Papadimitropoulos, A., Barbero, A., Manz, M.G., and Martin, I., Engineering of a functional bone organ through endochondral ossification. Proc Natl Acad Sci U S A, 2013. 110(10): p. 3997-4002. Mackie, E.J., Ahmed, Y.A., Tatarczuch, L., Chen, K.S., and Mirams, M., Endochondral ossification: how cartilage is converted into bone in the developing skeleton. Int J Biochem Cell Biol, 2008. 40(1): p. 46-62. Dai, J. and Rabie, A.B., VEGF: an essential mediator of both angiogenesis and endochondral ossification. J Dent Res, 2007. 86(10): p. 937-50. Gerstenfeld, L.C., Cruceta, J., Shea, C.M., Sampath, K., Barnes, G.L., and Einhorn, T.A., Chondrocytes provide morphogenic signals that selectively induce osteogenic differentiation of mesenchymal stem cells. J Bone Miner Res, 2002. 17(2): p. 221-30. Hwang, N.S., Varghese, S., Puleo, C., Zhang, Z., and Elisseeff, J., Morphogenetic signals from chondrocytes promote chondrogenic and osteogenic differentiation of mesenchymal stem cells. J Cell Physiol, 2007. 212(2): p. 281-4. Hutmacher, D.W., Schantz, J.T., Lam, C.X., Tan, K.C., and Lim, T.C., State of the art and future directions of scaffold-based bone engineering from a biomaterials perspective. J Tissue Eng Regen Med, 2007. 1(4): p. 245-60. Spiller, K.L., Maher, S.A., and Lowman, A.M., Hydrogels for the repair of articular cartilage defects. Tissue Eng Part B Rev, 2011. 17(4): p. 281-99. Sill, T.J. and von Recum, H.A., Electrospinning: applications in drug delivery and tissue engineering. Biomaterials, 2008. 29(13): p. 1989-2006. Pham, Q.P., Sharma, U., and Mikos, A.G., Electrospinning of polymeric nanofibers for tissue engineering applications: a review. Tissue Eng, 2006. 12(5): p. 1197-211. Wang, L., Li, Y., Zuo, Y., Zhang, L., Zou, Q., Cheng, L., and Jiang, H., Porous bioactive scaffold of aliphatic polyurethane and hydroxyapatite for tissue regeneration. Biomed Mater, 2009. 4(2): p. 025003. Liu, H., Zhang, L., Zuo, Y., Wang, L., Huang, D., Shen, J., Shi, P., and Li, Y., Preparation and characterization of aliphatic polyurethane and hydroxyapatite composite scaffold. Journal of Applied Polymer Science, 2009. 112(Compendex): p. 2968-2975. Yang, W., Yang, F., Wang, Y., Both, S.K., and Jansen, J.A., In vivo bone generation via the endochondral pathway on three-dimensional electrospun fibers. Acta Biomater, 2013. 9(1): p. 4505-12. Farndale, R.W., Buttle, D.J., and Barrett, A.J., Improved quantitation and discrimination of sulphated glycosaminoglycans by use of dimethylmethylene blue. Biochim Biophys Acta, 1986. 883(2): p. 173-7. Livak, K.J. and Schmittgen, T.D., Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods, 2001. 25(4): p. 402-8. Mahmoudifar, N. and Doran, P.M., Chondrogenesis and cartilage tissue engineering: the longer road to technology development. Trends Biotechnol, 2012. 30(3): p. 166-76. Kuboki, Y., Jin, Q., and Takita, H., Geometry of carriers controlling phenotypic expression in BMP-induced osteogenesis and chondrogenesis. J Bone Joint Surg Am, 2001. 83-A Suppl 1(Pt 2): p. S105-15. Benders, K.E., van Weeren, P.R., Badylak, S.F., Saris, D.B., Dhert, W.J., and Malda, J., Extracellular matrix scaffolds for cartilage and bone regeneration. Trends Biotechnol, 2013. 31(3): p. 169-76. Kronenberg, H.M., Developmental regulation of the growth plate. Nature, 2003. 423(6937): p. 332-6. Gerber, H.P., Vu, T.H., Ryan, A.M., Kowalski, J., Werb, Z., and Ferrara, N., VEGF couples hypertrophic cartilage remodeling, ossification and angiogenesis during endochondral bone formation. Nat Med, 1999. 5(6): p. 623-8. 103 5 Chapter 5 28. Karageorgiou, V. and Kaplan, D., Porosity of 3D biomaterial scaffolds and osteogenesis. Biomaterials, 2005. 26(27): p. 5474-91. 29. Jin, Q.M., Takita, H., Kohgo, T., Atsumi, K., Itoh, H., and Kuboki, Y., Effects of geometry of hydroxyapatite as a cell substratum in BMP-induced ectopic bone formation. J Biomed Mater Res, 2000. 51(3): p. 491-9. 30. Sundelacruz, S. and Kaplan, D.L., Stem cell- and scaffold-based tissue engineering approaches to osteochondral regenerative medicine. Semin Cell Dev Biol, 2009. 20(6): p. 646-55. 31. Scaglione, S., Giannoni, P., Bianchini, P., Sandri, M., Marotta, R., Firpo, G., Valbusa, U., Tampieri, A., Diaspro, A., Bianco, P., and Quarto, R., Order versus Disorder: in vivo bone formation within osteoconductive scaffolds. Sci Rep, 2012. 2: p. 274. 32. Nirmala, R., Nam, K.T., Navamathavan, R., Park, S.J., and Kim, H.Y., Hydroxyapatite Mineralization on the Calcium Chloride Blended Polyurethane Nanofiber via Biomimetic Method. Nanoscale Res Lett, 2010. 6(1): p. 2. 33. Chai, Y.C., Carlier, A., Bolander, J., Roberts, S.J., Geris, L., Schrooten, J., Van Oosterwyck, H., and Luyten, F.P., Current views on calcium phosphate osteogenicity and the translation into effective bone regeneration strategies. Acta Biomater, 2012. 8(11): p. 3876-87. 34. LeGeros, R.Z., Calcium phosphate-based osteoinductive materials. Chem Rev, 2008. 108(11): p. 4742-53. 35. Oliveira, S.M., Ringshia, R.A., Legeros, R.Z., Clark, E., Yost, M.J., Terracio, L., and Teixeira, C.C., An improved collagen scaffold for skeletal regeneration. J Biomed Mater Res A, 2010. 94(2): p. 371-9. 36. Hartman, E.H., Vehof, J.W., Spauwen, P.H., and Jansen, J.A., Ectopic bone formation in rats: the importance of the carrier. Biomaterials, 2005. 26(14): p. 1829-35. 37. Tortelli, F., Tasso, R., Loiacono, F., and Cancedda, R., The development of tissue-engineered bone of different origin through endochondral and intramembranous ossification following the implantation of mesenchymal stem cells and osteoblasts in a murine model. Biomaterials, 2010. 31(2): p. 242-9. 38. Pelttari, K., Winter, A., Steck, E., Goetzke, K., Hennig, T., Ochs, B.G., Aigner, T., and Richter, W., Premature induction of hypertrophy during in vitro chondrogenesis of human mesenchymal stem cells correlates with calcification and vascular invasion after ectopic transplantation in SCID mice. Arthritis Rheum, 2006. 54(10): p. 3254-66. 39. Abrahamsson, C.K., Yang, F., Park, H., Brunger, J.M., Valonen, P.K., Langer, R., Welter, J.F., Caplan, A.I., Guilak, F., and Freed, L.E., Chondrogenesis and mineralization during in vitro culture of human mesenchymal stem cells on three-dimensional woven scaffolds. Tissue Eng Part A, 2010. 16(12): p. 3709-18. 104 Chapter 6 Effects of in vitro chondrogenic priming time of bone marrow derived mesenchymal stem cells on in vivo endochondral bone formation Wanxun Yang, Sanne K. Both, Gerjo J. V. M.van Osch, Yining Wang, John A. Jansen and Fang Yang Effects of In Vitro Priming Time on Endochondral Bone Formation Summary of this chapter Recapitulation of endochondral ossification leads to a new concept of bone tissue engineering via a cartilage intermediate as an osteoinductive template. To enable an effective translation of this endochondral approach to clinic, it is imperative to possibly shorten the in vitro procedure for the creation of cartilage template and meanwhile ensure a sufficient bone formation. Therefore, in the chapter, we aimed to investigate the influence of in vitro chondrogenic priming time on the in vivo endochondral bone formation both qualitatively and quantitatively. To this end, rat bone marrow derived mesenchymal stem cells were seeded on two scaffolds with distinguished features, i.e. a fibrous poly(lactic-coglycolic acid)/poly(ε-caprolactone) electrospun scaffold (PLGA/PCL) or a porous hydroxyapatite/tricalcium phosphate composite (HA/TCP). Next the constructs were chondrogenically differentiated for 2, 3 and 4 weeks in vitro followed by subcutaneous implantation for up to 8 weeks. A longer chondrogenic priming time resulted in a significantly increased amount and homogeneous deposition of the cartilage matrix on both PLGA/PCL and HA/TCP scaffolds in vitro. In vivo, all implanted constructs gave rise to endochondral bone formation, whereas the bone volume was not affected by a longer priming time. An unpolarized woven bone like structure, with significant amounts of cartilage remaining, was generated in fibrous PLGA/PCL scaffolds, while porous HA/TCP scaffolds supported progressive lamellar-like bone formation with mature bone marrow development. These data suggest that, by utilizing a chondrogenically differentiated MSC-scaffold construct as cartilage template, the essential in vitro priming time could be shortened to 2 weeks to generate a substantial amount of vascularized endochondral bone in vivo. The structure of the bone was controlled by the chemical and structural cues provided by the scaffold design. 1. Introduction The bone forming process during development and regeneration is governed by intramembranous or endochondral mechanism, through which bone is generated by direct osteoblastic differentiation or remodelling from a cartilage intermediate, respectively [1-3]. Recapitulation of endochondral ossification leads to a new route to engineer bone tissue, which generally involves two steps: 1) creating a cartilage intermediate in vitro and 2) implanting the cartilage intermediate as an osteoinductive template to transform into bone [4, 5]. Recent studies have shown the superiority of generating vascularized bone tissue via the endochondral approach, which has the potential to overcome the issue of insufficient vascularization faced 109 6 Chapter 6 by the conventional bone tissue engineering based on the intramembranous ossification [6-9]. In general, to create the cartilage intermediate, mesenchymal stem cells (MSCs) loaded in a three dimensional scaffold or in a cell pellet form are cultured in chondrogenic conditions for a certain time period before in vivo implantation. The obtained cartilage serves as a morphological template enabling blood vessel invasion and endogenous MSCs recruitment for successive bone formation [10]. It is clear that such a cartilage construct contains essential “biological instructions” to initiate and regulate the endochondral bone forming process, and this largely relies on the stage of the chondrogenic differentiation, as chondrocytes possess specific metabolic features and secrete diverse biochemical cues when they are in different developmental stages [9, 11]. A fine coordination between the progression of the cartilage construct and endochondral transformation needs to be launched by choosing an optimal time period for in vitro chondrogenic priming. In literature, the protocol applied for chondrogenically differentiating MSCs was typically aimed to reach the stage with a certain degree of hypertrophy before implantation, which usually lasts for 4 to 5 weeks [4, 6, 9]. The hypothesis behind this approach was that hypertrophy could ensure the progression of endochondral differentiation and secretion of substantial amount of vascular endothelial growth factor (VEGF), which might enable timely establishment of vascularization into the implants [12]. However, this assumption was made mainly based on the use of cell pellets, and it has not yet been confirmed [13, 14]. For the final application, it is more relevant to elucidate the optimal in vitro priming time for the creation of cartilage template when a scaffold material is involved. From a clinical point of view, a long in vitro culture time associates with considerably high treatment cost and regulatory burden, which hampers the effective translation of this tissue engineering strategy to the clinics. Therefore, it is imperative to seek the possibility of shortening this procedure and meanwhile ensure a sufficient endochondral bone formation. In order to gain more insight on the endochondral approach for future clinical applications, we aimed to investigate the influence of chondrogenic priming time on the in vivo endochondral bone formation both qualitatively and quantitatively. Furthermore, to examine if such influence is universal or also depends on the properties of the applied scaffolds, two representative scaffolds with distinguished characteristics were selected as cell carrier, i.e. a polymer-based fibrous poly(lactic-co-glycolic acid)/poly(εcaprolactone) electrospun scaffold (PLGA/PCL) and a ceramic-based porous hydroxyapatite/tricalcium phosphate composite (HA/TCP). Rat bone marrow derived MSCs were first seeded on these two scaffolds and cultured in chondrogenic 110 Effects of In Vitro Priming Time on Endochondral Bone Formation condition for 2, 3 and 4 weeks. Afterwards, these cell-scaffold constructs were implanted subcutaneously for up to 8 weeks in nude rats to assess bone formation and cartilage remains by histological stainings and histomorphometric analysis. 2. Materials and Methods 2.1 Cell isolation, seeding and culture MSCs were isolated from 6-week-old male Fischer rats (Charles River) after the approval from Radboud University Animal Ethics Committee (Approval no: RUDEC 2012-195). Briefly, two femora of each rat were removed and the epiphyses were cut off. Rat MSCs were flushed out from the remaining diaphyses using the proliferation medium through a needle. The proliferation medium consisted of alpha Minimal Essential Medium (aMEM; Gibco, Grand Island, USA) supplemented with 10% fetal bovine serum (FBS; Sigma, St. Louis, USA), 100 U/ml penicillin, 100μg/ml streptomycin (Gibco), 50 μg/ml l-ascorbic acid (Sigma) and 10nM dexamethasone (Sigma). The flush-out was cultured for 1 day in a humidified incubator (37 °C, 5% CO2), after which the medium was refreshed to remove non-adherent cells. The cells were cultured for an additional 4 days then they were detached using trypsin/ EDTA (0.25% w/v trypsin, 0.02% w/v EDTA; Sigma) and counted. The cells were seeded in two types of scaffold: 1) Fibrous poly(lactic-co-glycolic acid) (PLGA, Purasorb® PDLG 8531, Purac Biomaterials BV, Gorinchem, the Netherlands) combined with poly(ε-caprolactone) (PCL, LACTEL® Absorbable Polymers, DURECT Corporation, Pelham, USA) electrospun scaffold made by a wetelectrospinning technique (Esprayer ES-2000S, Fuence Co., Ltd, Tokyo, Japan) as previously described [7]. This PLGA/PCL scaffold was in a disk shape with 6 mm of diameter and 2 mm of thickness. 2) Porous sintered hydroxyapatite (HA)/tricalcium phosphate (TCP) composite (CAM biceramics, Leiden, the Nethelands) in a cubic shape of 3×3×3 mm. This HA/TCP scaffold consisted of HA and TCP with a 75/25 weight ratio and the volumetric porosity was 75% (provided by the manufacturer). Before cell seeding, the scaffolds were sterilized by either autoclave (HA/TCP) or in 70% ethanol for 2 hours (PLGA/PCL) and subsequently soaked in the proliferation medium overnight after being washed with PBS. During cell loading, the scaffolds were incubated in the cell suspension with a concentration of 1 × 106 cells/ml (4 scaffolds per 1 ml of cell suspension) and gently rotated for 3 hours. Afterwards the unattached cells from the suspension were collected, centrifuged and reseeded onto the scaffolds to ensure a high cell loading efficiency. All cell-scaffold constructs were cultured for 1 week in proliferation medium. Thereafter, the medium was changed to chondrogenic medium, consisting of 111 6 Chapter 6 high-glucose DMEM (Gibco), 1% FBS, 100 U/ml penicillin, 100 μg/ml streptomycin, 50 μg/ml l-ascorbic acid, 100 mM of sodium pyruvate (Gibco), 40μg/ml l-proline (Sigma), 1:100 insulin-transferrin-selenium (ITS; BD Biosciences, Bedford, USA), 100nM dexamethasone, 10 ng/ml of transforming growth factor beta-2 (TGF-β2; R&D systems, Minneapolis, USA) and 100 ng/ml bone morphogenetic protein 2 (BMP-2; BD biosciences). The constructs were cultured for 2, 3 or 4 weeks in chondrogenic medium which was refreshed twice a week. At each time point, the cell-scaffold constructs were collected and divided into two groups. One group of samples were used for in vitro characterization and the other for subcutaneous implantation. 2.2 In vitro tests 2.2.1 GAG content The amount of GAG in each sample was detected by a sulfated glycosaminoglycan (sGAG) assay (n=3). After being washed twice in PBS, all samples were digested in proteinase K solution (1 mg/ml proteinase K in 50 mM Tris with 1 mM EDTA, 1 mM iodoacetamide, and 10 μg/ml pepstatin A) (all reagents from Sigma) for 20 hours at 56°C. The GAG content was determined by using dimethylmethylene blue (Polysciences Inc., Warrington, USA), with chondroitin sulfate (Sigma) as standard control. Measurements of absorption were performed at 530 nm and 570 nm using an ELISA reader (Bio-Tek Instruments Inc., Winooski, USA). The results were calculated according to the previously described method [15]. 2.2.2 DNA content DNA content (n=3) was quantified by QuantiFluorTM DNA system (Promega Corporation, Madison, USA) using the same cell extract solution as used for the GAG assay. Followed the manufacturer’s instruction, a DNA standard curve was used to quantify the amount of DNA in each sample and the results were measured using a fluorescence microplate reader (Bio-Tek Instruments Inc.) with an excitation wavelength at 485nm and an emission wavelength at 530nm. 2.2.3 RNA extraction and qPCR analysis After 2, 3 and 4 weeks of chondrogenic culture, RNA of the seeded cells was extracted using TRIzol® reagent according to the manufacturer’s instructions (Life Technologies, Carlsbad, USA). The obtained RNA was dissolved in RNase free water and the concentration was measured with spectrophotometer (NanoDrop 2000, Thermal scientific, Wilmington, USA). Subsequently, first strand cDNA was reverse transcribed from RNA using iScript™ cDNA Synthesis Kit (Bio-Rad Laboratories, Inc., USA). Afterwards cDNA was further amplified and the expression of specific genes 112 Effects of In Vitro Priming Time on Endochondral Bone Formation was quantified using IQ SYBR Green Supermix PCR kit (BioRad, Hemel Hempstead, United Kingdom) in a real-time PCR (BioRad, CFX96™ real-time system) (n=4). Chondrogenic-related gene markers were evaluated, including collagen type II (Col2α1), aggrecan (Acan), collagen type X (Col10α1) and VEGF (Vegfa) (Table 1). The specificity of the primers was confirmed separately before the real-time PCR reaction. The expression levels were compared to the housekeeping gene β-Actin and calculated via the 2-∆∆Ct method [16]. Table 1. Primer sequences used for qPCR Forward (5’ → 3’) Reverse (5’ → 3’) Col2α1 GACGCCACGCTCAAGTCGCT CGCTGGGTTGGGGTAGACGC Col10α1 GGGCCCTATTGGACCACCAGGTA CCGGCATGCCTGTTACCCCC Acan CATTCGCACGGGAGCAGCCA TGGGGTCCGTGGGCTCACAA Vegfa ACTCATCAGCCAGGGAGTCT GGGAGTGAAGGAGCAACCTC β-Actin TTCAACACCCCAGCCATGT TGTGGTACGACCAGAGGCATAC 2.2.4 Histological analysis The in vitro HA/TCP scaffolds were fixed in 10% phosphate buffered formalin for 1 hour and embedded in methylmetacrylate (MMA, L.T.I., Bilthoven, the Netherlands) (n=3). Sections of to 15μm were made using a microtome equipped with diamond blade (Leica SP1600, Leica microsystem, Wetzlar, Germany) and then stained with 0.04% thionin solution (Sigma). The in vitro PLGA/PCL scaffolds were fixed in 10% phosphate buffered formalin for 1 hour and embedded in paraffin (n=3). The microtome sections with 5 μm of thickness were also stained with 0.04% thionin solution after deparaffinization in xylene and rehydration through graded ethanol. 2.3 In vivo implantation After 2, 3 and 4 weeks of in vitro culture, the cell-scaffold constructs were implanted subcutaneously in 8-week-old male nude rats (Crl:NIH-Foxn1mu, Charles River). The protocol was approved by the Animal Ethical Committee of the Radboud University Medical Centre (Approval no: RU-DEC 2012-195) and the national guidelines for the care and use of laboratory animals were applied. For each time condition, 3 rats were used for the implantation. All rats received analgesic before and after the surgery. Surgery was performed under general inhalation anesthesia with a combination of isoflurane, nitrous oxide, and oxygen. The back of the rats was shaved and disinfected with povidone-iodine. Subsequently, four small longitudinal incisions were made and a subcutaneous pocket was created lateral to each incision. Two HA/TCP and two PLGA/PCL constructs were implanted to 113 6 Chapter 6 each rat (one construct per pocket). After placing the samples, the skin was closed using staples. After 4, 6 and 8 weeks, rats were euthanized by CO2-suffocation for sample collection. 2.4 Histological staining of the in vivo samples All in vivo samples were fixed in 10% phosphate buffered formalin for 30 hours and dehydrated through graded ethanol. For each rat, half of the samples (one HA/TCP and one PLGA/PCL construct) were embedded in MMA, and the other half samples were embedded in paraffin. MMA embedded samples (n = 3 for each time condition) were sectioned using a microtome equipped with diamond saw (Leica SP1600, Leica microsystem, Wetzlar, Germany) and then stained with methylene blue and basic fuchsin (both reagents from Merck). Prior to paraffin embedding, the fixed samples (n = 3 for each time condition) were decalcified by 10% EDTA and the decalcification process was monitored via X-ray. Five μm-thick sections were made from each sample and used for hematoxylin/ eosin (HE), Elastic Van Gieson (EVG) and collagen II staining. 2.5 Histological and histomorphometrical analysis Digital images of all histological sections were made using a light microscope equipped with a digital camera (Axio Imager Z1, Carl Zeiss AG Light Microscopy, Göttingen, Germany). Histological analysis consisted of a concise morphological and histological evaluation of all sections. Histomorphometrical analysis was performed with computer-based quantitative image analysis techniques based on color recognition with manual correction (Leica QWin Pro®, Leica Microsystems AG, Wetzlar, Germany). For both HA/TCP and PLGA/PCL scaffolds, quantification of bone and cartilage area was performed on EVG staining images and collagen II staining images, respectively. The region of interest (ROI) in each image was defined from the total area of the cross section (including the scaffold region) which was manually set using Leica QWin software. Bone and cartilage tissue were discriminated by the difference in staining intensity and cellular morphology. The parameters “bone area” and “cartilage area” were evaluated and defined as a percentage of bone or cartilage tissue within the ROI, respectively. Histomorphometrical data were collected based on three sections per specimen and averaged to one value per specimen. 114 Effects of In Vitro Priming Time on Endochondral Bone Formation 2.6 Statistic analysis The data was analyzed using two-way ANOVA followed by Bonferroni’s Multiple Comparison Test (GraphPad Prism 5, GraphPad Software Inc., San Diego, USA). Mean values and standard deviations (SD) are reported in each figure as well as relevant statistical relationships. Differences were considered statistically significant at p < 0.05. 3. Results 3.1 In vitro cartilage formation pre-implantation 3.1.1 GAG production and DNA content As shown in Fig. 1a, on both PLGA/PCL and HA/TCP scaffolds the total amount of GAG deposition accumulated with chondrogenic culture time and reached the highest amount in week 4. PLGA/PCL demonstrated significant higher amount 6 Figure 1. GAG (a) and DNA (b) content of PLGA/PCL and HA/TCP constructs after in vitro chondrogenic priming. * p<0.05, ** p<0.01; error bars represent standard deviation of the mean (n=3) 115 Chapter 6 of GAG compared to that on the HA/TCP scaffolds. Furthermore, a comparable DNA amount was detected on both PLGA/PCL and HA/TCP, which remained stable during the entire culture period (Fig. 1b). Figure 2. Gene expression of collagen II (a), aggrecan (b), collagen X (c) and VEGF (d) after in vitro chondrogenic priming. * p<0.05, ** p<0.01; error bars represent standard deviation of the mean (n=4) 3.1.2 Expression of chondrogenic genes For both PLGA/PCL and HA/TCP scaffolds, Col2α1 and Acan expression mantained stable during the in vitro chondrogenic priming process and comparable expression levels were detected on both scaffolds (Fig. 2a and b). On the PLGA/PCL constructs there was a peak for the Col10α1 expression at week 3. In comparison, for the HA/TCP constructs, the expression of Col10α1 remained constant till week 3 and then decreased dramatically at week 4. Moreover, significantly higher expression of Col10α1 was shown on the PLGA/PCL constructs compared to that on HA/TCP after 3 or 4 weeks of chondrogenic priming (Fig. 2c). Regarding the expression of Vegfa, HA/TCP constructs demonstrated a significantly higher level than that of PLGA/PCL. Overall, for both scaffolds, the level of Vegfa expression remained constant during the in vitro priming. 116 Effects of In Vitro Priming Time on Endochondral Bone Formation 6 Figure 3. In vitro chondrogenic differatiation in PLGA/PCL and HA/TCP scaffolds after 2 and 4 weeks of culture. Cartilage matrix formation (stained purple) and cell penetration inside of the PLGA/PCL scaffolds could be observed after 2 weeks of chondrogenic priming (a and c). Nevertheless, the cartilage matrix mainly located on one side of the scaffolds (a). When the culture time reached 4 weeks, the cartilage matrix spread throughout the entire fibrous structure (b) and a great number of cells with typical chondrocyte-like morphology could be observed (d). For HA/TCP scaffolds, the cartilage formation could be observed on the pore surfaces of 2-week primed constructs (e), while the cell penetration and matrix deposition were very limited in the center of the scaffolds (e). Chondrocyte-like cells were embedded in the matrix and mixed with undifferentiated cells (g). As the priming time increased, the cartilage matrix deposition was augmented and distributed more homogeneously in the scaffolds (f). After 4 weeks, the presented cells in the scaffolds mainly consisted of chondrocyte-like cells (h). 117 Chapter 6 3.1.3 Histological analysis To visualize the cartilage matrix deposition in the scaffolds, sections were made and stained with thionin. During the histological processing, PLGA/PCL scaffolds shrank and formed folds on the surfaces as shown in Fig. 4a. Yet cartilage matrix (GAG) formation and cell penetration inside of the scaffolds could be clearly observed after 2 weeks of chondrogenic priming (Fig. 3a and c). Nevertheless, the cartilage matrix mainly located on one side of the scaffolds. When the culture time reached 4 weeks, the cartilage matrix spread throughout the entire fibrous structure (Fig. 3b) and a large number of cells with typical chondrocyte-like morphology could be observed (Fig. 3d). For HA/TCP scaffolds, the cartilage formation could be observed on the pore surfaces after 2 weeks of chondrogenic priming, while the cell penetration and matrix deposition were limited in the center of the scaffolds (Fig. 3e). Typical chondrocytelike cells were embedded in the matrix and mixed with undifferentiated cells (Fig. 3g). As the chondrogenic priming time increased, the cartilage matrix deposition was augmented and distributed more homogeneously in the scaffolds (Fig. 3f). After 4 weeks, the cells present in the scaffolds mainly consisted of chondrocytes which were surrounded by abundant matrix and appeared in isogenous groups or with large lacunae (Fig. 3h). 3.2 In vivo bone formation 3.2.1 General observations From the total 27 animals available for surgery, one animal died during anesthesia. Due to this, 3-week in vitro priming group had one animal less than the other groups for evaluating the bone formation at the 8-week time point. The remaining 26 animals recovered uneventfully from the surgical procedure and remained in good health. No signs of wound complications were observed post-operatively. At the time point of 4, 6 or 8 weeks, all samples were retrieved and macroscopic signs of inflammation or adverse tissue responses were absent for all of the retrieved samples. 3.2.2 Descriptive histology Bone formation was found in all PLGA/PCL and HA/TCP scaffolds. The bone forming quality and quantity were similar in all samples within the same group. 3.2.2.1 PLGA/PCL scaffold With different in vitro chondrogenic priming time, extensive cartilage remodeling has already occurred within the PLGA/PCL scaffolds after 4 weeks of implantation (Fig. 4a and b). The bone/cartilage complex distributed in the scaffolds in a 118 Effects of In Vitro Priming Time on Endochondral Bone Formation dispersive manner and displayed a trabecula-like structure. The implanted scaffolds that received 4 weeks of chondrogenic priming in vitro showed higher amount of bone/cartilage matrix presence in vivo with less fibrous tissue occupation (Fig. 4b) compared to the constructs which were chondrogenically differentiated for 2 weeks (Fig. 4a). From the HE staining pictures, a mixture of bone and cartilage tissue could be observed in all PLGA/PCL scaffolds independent of in vitro priming and in vivo implantation time (Fig. 5). Bone matrix was frequently found deposited on the periphery of cartilage tissue or around the enlarged lacunae of chondrocytes. The newly formed bone tissue resembled a woven bone structure, in the surrounding of which the blood vessel formation could occasionally be observed. Concentrically organized bone tissue and marrow development could not be seen (Fig. 5). Furthermore, the non-degraded electrospun fibers of the scaffolds were also integrated with the bone/cartilage matrix and surrounding tissue. 6 Figure 4. Overview of in vivo bone formation. MMA sections of PLGA/PCL (a and b) and HA/TCP (c and d) constructs after 4 weeks of implantation were stained with methylene blue and basic fuchsin. In PLGA/PCL constructs, the bone/cartilage complex distributed in the scaffolds in a dispersive manner and displayed a trabeculalike structure. With 2 weeks of in vitro chondrogenic priming time, extensive cartilage remodeling has already occurred (a), while 4-week primed PLGA/PCL constructs showed higher amount of bone/cartilage matrix presence in vivo with less fibrous tissue occupation (b). For HA/TCP scaffolds, bone formation occurred throughout the whole implants with 2-week (c) or 4-week (d) in vitro priming, which was characterized by close apposition of matrix on the surface of the pores together with islet or bridge like bone tissue in the middle space of the pores. 119 Chapter 6 Figure 5. Morphology of in vivo bone formation in PLGA/PCL constructs (HE staining). A mixture of bone and cartilage tissue could be observed in all PLGA/PCL scaffolds independent of in vitro priming and in vivo implantation time. Bone matrix (stained red) was frequently found deposited on the periphery of cartilage tissue (stained purple) or around the enlarged lacunae of chondrocytes. The newly formed bone resembled a structure of the woven bone with vascular infiltration in the surrounding tissue. Concentrically organized bone tissue and marrow development could not be seen. (B: bone; C: cartilage.) 120 Effects of In Vitro Priming Time on Endochondral Bone Formation 6 Figure 6. Morphology of in vivo bone formation in HA/TCP constructs (HE staining). The length of the in vitro priming time did not show obvious influences on the structure of the newly formed bone (stained red) at each in vivo time point. After 4 weeks of in vivo implantation, the new bone displayed in an irregular shape and the osteoblasts could be recognized aligning at the periphery of the bone matrix (a and e). The cartilage remnants (stained purple) were observed to fill small pores or mix with the bone tissue (b and f). Marrow stroma development was initiated and filled with abundant hematopoietic cells. After 8 weeks in vivo, the matrix exhibited a concentric pattern and the collagen bundles arranged in a more parallel manner (c and g). The bone marrow cavity was mainly occupied by adipose cells, amongst which a great number of blood vessels were well represented. Cartilage remnant was hardly seen inside of the scaffolds (d and h). (BM: bone marrow; Black arrow: cartilage.) 121 Chapter 6 3.2.2.2 HA/TCP scaffold For HA/TCP scaffolds, bone formation occurred throughout the whole implants. Only small amount of cartilage remained after 4 weeks of in vivo implantation. The bone tissue was characterized by close apposition of matrix on the surface of the HA/TCP pores together with islet or bridge like bone tissue in the middle space of the pores (Fig. 4c and d). At each in vivo time point, the length of the in vitro priming time did not show obvious influences on the structure of the newly formed bone and cartilage remnant. The maturity of the bone tissue was observed to change over time in vivo. After 4 weeks of in vivo implantation, the new bone displayed in an irregular shape and the osteoblasts could be recognized aligning at the periphery of the bone matrix (Fig. 6a and e). The cartilage remnants were observed to fill small pores or mix with the bone tissue (Fig. 6b and f). Marrow stroma development was initiated and filled with abundant hematopoietic cells. After 8 weeks in vivo, the bone tissue further matured. As shown in Fig. 6c and g, the matrix exhibited a concentric pattern and the collagen bundles arranged in a more parallel manner compared to the early stage bone formation. Bone surface was partially covered by elongated, thin lining cells. Moreover, the bone marrow cavity was mainly occupied by adipose cells, amongst which a great number of blood vessels were well represented (Fig. 6d and h). Cartilage remnants were hardly detected inside of the scaffolds. 3.2.3 Histomorphometrical evaluation The EVG staining and Collagen II staining were used to visualize the presence of bone and cartilage tissue, respectively. As shown in the overview pictures of PLGA/ PCL scaffolds (Fig. 7a), longer in vitro priming time resulted in more homogeneous distribution of in vivo bone formation and less fibrous tissue occupation within the PLGA/PCL scaffolds. The collagen II staining confirmed the significant existence of the remaining cartilage matrix (Fig. 7b). For HA/TCP scaffolds, similar amount and location of bone formation was revealed at all in vivo time points with very limited presence of cartilage remnant (Fig. 8). Quantitative results of bone and cartilage amount are depicted in Fig. 9. The volume of newly formed bone in PLGA/PCL scaffolds did not alter significantly within the in vivo implantation time, with the average bone area fraction ranged from 12% to 18% (Fig. 9a). In contrast, for HA/TCP scaffolds, bone formation was increased as the in vivo time and the average area fraction value raised from 15% to 25% (Fig. 9b). For both scaffolds, the length of in vitro priming time did not show obvious influences on bone volume at each in vivo time point. However, significant higher amount of cartilage remnants was detected in PLGA/PCL scaffolds when the constructs received 4-week chondrogenic treatment in vitro (Fig. 9c). In HA/ 122 Effects of In Vitro Priming Time on Endochondral Bone Formation TCP scaffolds, marginal amount of cartilage was found after 4 weeks of in vivo implantation and maintained stable thereafter, of which the area fraction was less than 5% (Fig. 9b). 6 Figure 7. Overview of in vivo bone formation and cartilage remnant in PLGA/PCL constructs. The decalcified in vivo samples were stained with EVG (a) and collagen II staining (b) to visualize the presence of bone and cartilage tissue, respectively. Longer in vitro priming time resulted in more homogeneous distribution of in vivo bone formation (stained red) and less fibrous tissue occupation within the PLGA/PCL scaffolds (a). The collagen II staining confirmed the significant existence of the remaining cartilage matrix (stained red) (b). (scale bar: 500µm) 4. Discussion To implement endochondral bone formation route for the clinical applications, it is imperative to utilize an essentially shortened in vitro priming procedure for the creation of a cartilage template to induce bone formation in vivo. In the current study, we aimed to investigate the influence of chondrogenic priming time on the in vivo endochondral bone formation regarding its quality and quantity. Furthermore, to examine if such influence is universal or also depends on which 123 Chapter 6 Figure 8. Overview of in vivo bone formation and cartilage remnant in HA/TCP constructs with different in vitro priming time. The decalcified in vivo samples were stained with EVG (a) and collagen II staining (b) to visualize the presence of bone and cartilage tissue, respectively. Different in vitro priming time resulted in similar amounts and locations of bone formation (stained red) with marginal cartilage remnants at each in vivo time point (stained red). (scale bar: 500µm) 124 Effects of In Vitro Priming Time on Endochondral Bone Formation scaffold material is applied, two scaffolds with distinguished characteristics were selected as cell carrier, i.e. fibrous PLGA/PCL and porous HA/TCP scaffolds. By varying the chondrogenic priming time, we were able to establish cartilage constructs with distinct features for this study. With longer priming time, the amount and homogeneity of cartilage matrix deposition increased significantly over time, and meanwhile the number of undifferentiated MSCs decreased. The expression of Col10α1, associated with chondrocyte hypertrophy [17, 18], was detected after 2 weeks of chondrogenic priming for both PLGA/PCL and HA/TCP constructs. These data suggest that a part of the MSCs on these scaffolds may have proceeded toward the hypertrophic stage at this time point. For 4-week primed constructs, a significant decrease of Col10α1 expression was shown, indicating the late stage of cell hypertrophy. Notably, for both scaffolds, Vegfa expression of the seeded cells has already been detected after 2 weeks, and the level maintained stable thereafter. These data confirmed the capability of VEGF release for all implanted constructs irrespective of the in vitro priming time. After implantation, our data suggested that further differentiation of the MSCscaffold based cartilage constructs prior to implantation did not result in a faster or enhanced bone formation in vivo. Similar quantity and quality of new bone tissue was generated in vivo in HA/TCP scaffolds irrespective of in vitro culturing time. For PLGA/PCL scaffolds, although longer-primed cartilage templates gave rise to higher quantity of cartilage remains and more homogeneously distribution of the new bone formation in vivo, the bone amount and maturity was in fact comparable. These results are inconsistent with several previous studies. Scotti et al. applied an advanced in vitro chondrogenic priming of human MSC pellets for 5 weeks, which resulted in accelerated in vivo bone formation compared to short priming for 2 weeks [13]. In comparison, Freeman et al. chondrogenically differentiated human and murine MSCs in pellets with different time periods from 10 days to 4 weeks. The results demonstrated 2 or 3-week priming was optimal to achieve the highest in vitro mineralization capacity of the cell pellets rather than 4 weeks [14]. However, it has to be noted that, in these studies the in vivo endochondral bone formation was tested via a scaffold-free cell pellet form, where the microenvironment and cellular behaviors are different from the case when a scaffold is involved. For the first time, our results indicated that, by utilizing a chondrogenically differentiated MSC-scaffold construct as cartilage template, the essential in vitro priming time could be shortened to 2 weeks to generate a substantial amount of vascularized endochondral bone in vivo. This can be closely correlated to the profile of Vegfa expression, which has already been detected after 2 weeks. As previously shown by Farrell et al., a considerable amount of VEGF was released from the cell 125 6 Chapter 6 pellets after 2 weeks of in vitro chondrogenic induction [4]. After implantation, the distinctive roles of VEGF could be exerted in the cartilage remodeling and bone forming process, such as promoting angiogenesis [19], chemoattractant effects on MSCs [20] and mediating osteoblastic differentiation [21]. Moreover, it is likely that the undifferentiated MSCs also participated in the process of endochondral bone formation. An explanation is that, these cells committed to chondrogenic Figure 9. Quantification of bone and cartilage amount after implantation. The total ratio of newly formed bone in PLGA/PCL scaffolds remained constant within the in vivo implantation time (a), while for HA/TCP scaffolds, bone formation was increased over time (b). For both scaffolds, the length of in vitro priming time did not show obvious influences on bone volume at each in vivo time point (a and b). However, significant higher amount of cartilage remnant was detected in PLGA/PCL scaffolds when the constructs received 4-week chondrogenic treatment in vitro (c). In HA/TCP scaffolds, marginal amount of cartilage was detected (d). (* p<0.05, ** p<0.01; # represents significant difference compared to the bone ratio at week 4 in vivo, # p<0.05, ## p<0.01; error bars represent standard deviation of the mean, n=3) lineage after implantation and fulfill their role in hypertrophy together with other chondrocytes to facilitate the vascularization and the recruitment of osteoprogenitors from the host animal [2, 22]. Another explanation is, that the undifferentiated MSCs were instructed by the hypertrophic chondrocytes to become osteoblasts or hematopoietic cells, and directly participated in building the vascularized bone tissue [23]. In either case, it implies that a certain extent of in vitro chondrogenic priming of the MSCs, e.g. for 2 weeks, or possibly with 126 Effects of In Vitro Priming Time on Endochondral Bone Formation even shorter time, might be sufficient to ensure the progression of endochondral differentiation of the seeded cells and enable successful bone formation, even after been implanted in a non-chondrogenic or non-osteogenic environment. Instead of being influenced by the in vitro priming time, our data have suggested that the newly formed endochondral bone could be generated in different manners based on the chemical and structural cues provided by the scaffold design. Newly formed woven bone like structure and significant cartilage presence have been observed in PLGA/PCL scaffolds after 4 weeks in vivo. Surprisingly, such structures were maintained up to 8 weeks. In comparison, cartilage remnants were hardly seen in HA/TCP scaffolds after 4 weeks. Over time, the amount and maturity of endochondral bone formation significantly improved after 8 weeks. These findings indicated that, HA/TCP scaffold supported more rapidly progression of the endochondral bone generation and vascularization compared to PLGA/PCL, which might be closely related to the high VEGF expression of the cells stimulated by the HA/TCP scaffolds. Regarding the pattern of the bone matrix, HA/TCP scaffolds instructed the new bone tissue to deposit on the surface of the interconnected pores in a polarized fashion, whereas for PLGA/PCL constructs, a fully integrated bone/cartilage/electrospun fiber complex was built without any preferential arrangement of bone tissue. Similar results have also shown previously by Scaglione et al. and they addressed such different bone-forming patterns were attributed to the specific scaffold geometry [24]. The confined porous structure could concentrate collagen fibers and serve as a polarizing template within the pores to trigger the formation of lamellar like bone. On the contrary, fibrous scaffolds possessed a high level of geometrical complexity, thus the new collagen fibers formed by the cells were distributed in a isotropic manner, resulting mostly in woven bone [24]. Above all, our data confirmed that scaffolds played a crucial role in influencing cell behaviors and bone deposition kinetics. The control over endochondral bone formation could be exerted through rational design of the scaffolds. This study contributes to the view that chondrogenical priming of rat MSCs in the presence of biomaterials followed by in vivo implantation was able to generate endochondral bone in vivo. Moreover, we demonstrated that it was not necessary to differentiate MSCs till fully hypertrophy to accomplish vascularized bone formation. However, it has to be noted that this conclusion needs to be further evidenced in a more clinically relevant situation by using human MSCs and implanting in an orthotopic location, where the mechanical load and biochemical/ inflammatory factors are closely involved. Moreover, we could not conclude if 2-week is the essential time length for chondrogenic priming and if such process can be further shortened, because no other shorter period has been chosen for this study. Additionally of interest is the further understanding of 1) the function 127 6 Chapter 6 and the fate of the implanted cells, and 2) the specific responses of the seeded cells induced by particular scaffold properties and the underlying mechanisms, so that the endochondral route can be established to target efficient and large-scale bone regeneration. 5. Conclusion In vitro chondrogenic priming of rat MSCs for 2, 3 and 4 weeks leads to vascularized endochondral bone formation in vivo on two distinct scaffolds, i.e. fibrous PLGA/ PCL and porous HA/TCP scaffolds. A longer in vitro priming time can result in a more homogeneous distribution of bone formation, whereas this does not have to give rise to a higher amount of in vivo bone formation in the selected scaffolds. Evidently, in our study the quality of endochondral bone was influenced by the chemical-physical design of the scaffolds. An unpolarized woven bone like structure was generated in fibrous PLGA/PCL scaffolds, while porous HA/TCP scaffolds supported progressive lamellar-like bone formation with mature bone marrow development. Further investigations are warranted to investigate the essential in vitro priming time and the mechanism of cell-scaffold interaction in the endochondral bone forming process. References 1. 2. 3. 4. 5. 6. 7. 128 Tortelli, F., Tasso, R., Loiacono, F., and Cancedda, R., The development of tissue-engineered bone of different origin through endochondral and intramembranous ossification following the implantation of mesenchymal stem cells and osteoblasts in a murine model. Biomaterials, 2010. 31(2): p. 242-9. Mackie, E.J., Ahmed, Y.A., Tatarczuch, L., Chen, K.S., and Mirams, M., Endochondral ossification: how cartilage is converted into bone in the developing skeleton. Int J Biochem Cell Biol, 2008. 40(1): p. 46-62. Buckwalter, J.A., Glimcher, M.J., Cooper, R.R., and Recker, R., Bone biology. II: Formation, form, modeling, remodeling, and regulation of cell function. Instr Course Lect, 1996. 45: p. 387-99. Farrell, E., van der Jagt, O.P., Koevoet, W., Kops, N., van Manen, C.J., Hellingman, C.A., Jahr, H., O’Brien, F.J., Verhaar, J.A., Weinans, H., and van Osch, G.J., Chondrogenic priming of human bone marrow stromal cells: a better route to bone repair? Tissue Eng Part C Methods, 2009. 15(2): p. 285-95. Jukes, J.M., Both, S.K., Leusink, A., Sterk, L.M., van Blitterswijk, C.A., and de Boer, J., Endochondral bone tissue engineering using embryonic stem cells. Proc Natl Acad Sci U S A, 2008. 105(19): p. 6840-5. Scotti, C., Piccinini, E., Takizawa, H., Todorov, A., Bourgine, P., Papadimitropoulos, A., Barbero, A., Manz, M.G., and Martin, I., Engineering of a functional bone organ through endochondral ossification. Proc Natl Acad Sci U S A, 2013. 110(10): p. 3997-4002. Yang, W., Yang, F., Wang, Y., Both, S.K., and Jansen, J.A., In vivo bone generation via the endochondral pathway on three-dimensional electrospun fibers. Acta Biomater, 2013. 9(1): p. 4505-12. Effects of In Vitro Priming Time on Endochondral Bone Formation 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. Farrell, E., Both, S.K., Odorfer, K.I., Koevoet, W., Kops, N., O’Brien, F.J., Baatenburg de Jong, R.J., Verhaar, J.A., Cuijpers, V., Jansen, J., Erben, R.G., and van Osch, G.J., In-vivo generation of bone via endochondral ossification by in-vitro chondrogenic priming of adult human and rat mesenchymal stem cells. BMC Musculoskelet Disord, 2011. 12: p. 31. Gawlitta, D., Farrell, E., Malda, J., Creemers, L.B., Alblas, J., and Dhert, W.J., Modulating endochondral ossification of multipotent stromal cells for bone regeneration. Tissue Eng Part B Rev. 16(4): p. 385-95. Goldring, M.B., Tsuchimochi, K., and Ijiri, K., The control of chondrogenesis. J Cell Biochem, 2006. 97(1): p. 33-44. Allori, A.C., Sailon, A.M., and Warren, S.M., Biological basis of bone formation, remodeling, and repair-part I: biochemical signaling molecules. Tissue Eng Part B Rev, 2008. 14(3): p. 259-73. Dai, J. and Rabie, A.B., VEGF: an essential mediator of both angiogenesis and endochondral ossification. J Dent Res, 2007. 86(10): p. 937-50. Scotti, C., Tonnarelli, B., Papadimitropoulos, A., Scherberich, A., Schaeren, S., Schauerte, A., Lopez-Rios, J., Zeller, R., Barbero, A., and Martin, I., Recapitulation of endochondral bone formation using human adult mesenchymal stem cells as a paradigm for developmental engineering. Proc Natl Acad Sci U S A, 2010. 107(16): p. 7251-6. Freeman, F.E., Haugh, M.G., and McNamara, L.M., Investigation of the optimal timing for chondrogenic priming of MSCs to enhance osteogenic differentiation in vitro as a bone tissue engineering strategy. J Tissue Eng Regen Med, 2013. Farndale, R.W., Buttle, D.J., and Barrett, A.J., Improved quantitation and discrimination of sulphated glycosaminoglycans by use of dimethylmethylene blue. Biochim Biophys Acta, 1986. 883(2): p. 173-7. Livak, K.J. and Schmittgen, T.D., Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods, 2001. 25(4): p. 402-8. Mueller, M.B. and Tuan, R.S., Functional characterization of hypertrophy in chondrogenesis of human mesenchymal stem cells. Arthritis Rheum, 2008. 58(5): p. 1377-88. Barry, F., Boynton, R.E., Liu, B., and Murphy, J.M., Chondrogenic differentiation of mesenchymal stem cells from bone marrow: differentiation-dependent gene expression of matrix components. Exp Cell Res, 2001. 268(2): p. 189-200. Gerber, H.P., Vu, T.H., Ryan, A.M., Kowalski, J., Werb, Z., and Ferrara, N., VEGF couples hypertrophic cartilage remodeling, ossification and angiogenesis during endochondral bone formation. Nat Med, 1999. 5(6): p. 623-8. Fiedler, J., Leucht, F., Waltenberger, J., Dehio, C., and Brenner, R.E., VEGF-A and PlGF-1 stimulate chemotactic migration of human mesenchymal progenitor cells. Biochem Biophys Res Commun, 2005. 334(2): p. 561-8. Mayr-Wohlfart, U., Waltenberger, J., Hausser, H., Kessler, S., Gunther, K.P., Dehio, C., Puhl, W., and Brenner, R.E., Vascular endothelial growth factor stimulates chemotactic migration of primary human osteoblasts. Bone, 2002. 30(3): p. 472-7. Ortega, N., Behonick, D.J., and Werb, Z., Matrix remodeling during endochondral ossification. Trends Cell Biol, 2004. 14(2): p. 86-93. Kronenberg, H.M., Developmental regulation of the growth plate. Nature, 2003. 423(6937): p. 332-6. Scaglione, S., Giannoni, P., Bianchini, P., Sandri, M., Marotta, R., Firpo, G., Valbusa, U., Tampieri, A., Diaspro, A., Bianco, P., and Quarto, R., Order versus Disorder: in vivo bone formation within osteoconductive scaffolds. Sci Rep, 2012. 2: p. 274. 129 6 Chapter 7 Roles of cartilage templates in mediating mesenchymal stem cell behavior in endochondral bone tissue engineering Wanxun Yang, Sanne K. Both, John A. Jansen and Fang Yang Roles of Cartilage Templates in Mediating MSCs Behavior Summary of this chapter Costly and lengthy process for creating cartilage template hampers the translation of endochondral approach to clinic. A more comprehensive understanding of the crosstalk between implanted cartilage template and host cells is essentially beneficial to possibly simplify this procedure or develop substitutes for effective bone regeneration. Therefore, this chapter aimed to explore the role of cartilage template in mediating MSC behavior in the process of endochondral bone formation, as well as the possible signaling pathways involved in these biological events. To that end, rat bone marrow derived mesenchymal stem cells (MSCs) were seeded on a porous hydroxyapatite (HA) /tricalcium phosphate (TCP) composite and cultured in chondrogenic condition to create the cartilage templates. Afterwards, undifferentiated MSCs were co-cultured with these constructs in a transwell system (separating cartilage template and MSCs with a permeable membrane) to assess the migration and osteogenic differentiation capacity of the MSCs. Furthermore, the expression profile of chemokine signaling related genes was analyzed via PCR array. The results demonstrated that chondrogenically differentiated MSC-HA/TCP constructs (chondro-TCP) were able to induce MSCs migration by possibly activating the IL8/CXCR2 signaling pathway. Additionally, they could trigger a considerable elevation of signaling factors responsible for MSCs recruitment and osteoclast stimulation, including CCL3, CCL7, CCL19 and NFκB. The osteogenic differentiation of MSCs was not significantly influenced after short exposure to chondro-TCP templates. 1. Introduction During both embryogenesis and postnatal fracture repair, the majority of bone formation is governed by a mechanism known as endochondral ossification [1]. Such endochondral process has recently inspired a new concept of bone tissue engineering via a two-step approach, during which a cartilage intermediate is first created in vitro, and subsequently implanted as an osteoinductive template in vivo to transform into bone [2]. It has been revealed this endochondral approach has a superior efficacy for generating vascularized bone tissue compared to the traditional intramembranous ossification method [3-5]. Still, several issues hurdle the effective translation of this tissue engineering strategy to the clinics. For instance, the creation of cartilage templates is usually time-consuming (3-5 weeks), and requires a large cell number as well as high usage of growth factors, e.g, transforming growth factor beta (TGF-β) and bone morphogenetic proteins (BMPs). All of these associate not only the difficulties in obtaining sufficient cell source but also considerably high treatment costs and regulatory burdens. 133 7 Chapter 7 Therefore, it is necessary to simplify the in vitro procedure or develop alternative products/compounds with similar or better functions as cartilage templates. Essentially, this would not be possible without a complete understanding of the mechanism underlying this endochondral approach, especially the in vivo functions of the cartilage template. In practice, the engineered cartilage template used in endochondral approach is usually created by loading mesenchymal stem cells (MSCs) in a three-dimensional scaffold or in cell pellet form followed by in vitro chondrogenic culturing. After implantation, the transformation from cartilage to bone is a very complex progression and involves the coordination of a variety of cell populations. Previous studies have shown that, instead of the direct contribution of the seeded cells from the cartilage templates, the in vivo endochondral bone formation was predominately originated by host cells [6]. Based on these results, it is speculated that the cartilage template plays an essential role in recruiting the host cells and regulating their performance in the successive bone forming process. Yet, this hypothesis has not been tested. The exact functions of the cartilage template for inducing regenerative functions of the host cells, e.g. MSCs, and the involved signaling pathways still remain unknown. Therefore, this study intended to investigate the effects of cartilage template on MSCs in terms of their migration and osteogenic differentiation ability. Furthermore, we also aimed to discover the possible signaling pathways involved in the crosstalk between cartilage templates and host MSCs. It is known that the communications between the cell populations are realized by different ways, including cell-cell, cell-matrix interactions, release of the soluble factors, or a combination thereof. Our study represented the first step and mainly focused on the effects of soluble factors. To that end, rat bone marrow derived MSCs were first seeded on a porous hydroxyapatite/tricalcium phosphate composite (HA/TCP) and cultured in chondrogenic condition to create the cartilage templates based on our previous study [5]. Afterwards, undifferentiated MSCs were co-cultured with these constructs in an indirect co-culture system (separating cartilage template and MSCs with a permeable membrane) to assess the migration and osteogenic differentiation capacity of the MSCs. Furthermore, the expression profile of chemokine signaling related genes was analyzed via PCR array to investigate the mechanism underlying the biological response of MSCs elicited by the cartilage template. 134 Roles of Cartilage Templates in Mediating MSCs Behavior 2. Materials and methods 2.1 Cell isolation Four 6-week-old male Wistar rats (Charles River, Wilmington, USA.) were used for MSC isolation after the approval from Radboud University Animal Ethics Committee. Two femora of each rat were removed and the epiphyses were cut off. Afterwards, rat MSCs were flushed out of the remaining diaphyses using the primary culture medium through an 18G needle (BD Microlance, Drogheda, Ireland), which consisted of alpha Minimal Essential Medium (αMEM; Gibco, Grand Island, USA) supplemented with 10% fetal bovine serum (FBS; Sigma, St. Louis, USA), 100 U/ml penicillin and 100μg/ml streptomycin (Gibco). The flushout was cultured for 2 days in a humidified incubator (37 °C, 5% CO2), after which the medium was refreshed to remove non-adherent cells. The cells were cultured for an additional 3 days followed by detachment using trypsin/EDTA (0.25% w/v trypsin, 0.02% w/v EDTA; Sigma) and counted. Half of the obtained the cells were used for cell seeding on the scaffolds, the other half was frozen in the culture medium containing 10% DMSO in the liquid nitrogen condition for later use. 2.2 Preparation of cartilage template A porous HA/TCP composite (CAM bioceramics B.V., Leiden, the Netherlands) was used as cell carrier, which was obtained in a cubic shape of 3×3×3 mm3 and consisted of HA and TCP with a 75/25 ratio. The volumetric porosity was 75% (provided by the manufacturer). Prior to cell seeding, the scaffolds were autoclaved and soaked in the primary culture medium overnight. During cell seeding, the scaffolds were incubated in the cell suspension with a concentration of 1 × 106 cells/ml (4 scaffolds per 1 ml of cell suspension) and gently rotated for 3 hours. Afterwards, the scaffolds were placed in non-adherent tissue culture plates and the unattached cells from the suspension were collected, centrifuged and reseeded onto the scaffolds to ensure a high cell loading efficiency. All of the constructs were proliferated for 1 week. Thereafter, the medium was changed to chondrogenic medium, consisting of high-glucose DMEM (Gibco), 1% FBS, 100 U/ ml penicillin, 100 μg/ml streptomycin, 50 μg/ml l-ascorbic acid, 100 mM of sodium pyruvate (Gibco), 1:100 insulin-transferrin-selenium (ITS; BD Biosciences, Bedford, USA), 100nM dexamethasone, 10 ng/ml of transforming growth factor beta-2 (TGF-β2; R&D systems, Minneapolis, USA) and 100 ng/ml bone morphogenetic protein 2 (BMP-2; BD biosciences). The constructs were cultured for 4 weeks to obtain the cartilage templates (named chondro-TCP). Bare HA/TCP scaffolds without cell seeding were used as control (named bare-TCP). 135 7 Chapter 7 2.3 MSCs migration test A tranwell migration model was used to examine the migration of MSCs in response to the cartilage constructs (Fig. 1a). Primary rat MSCs were defrosted, expanded and then seeded onto ThinCertTM cell culture inserts (membrane pore size 8μm, Greiner Bio-One) in a 24-well plate at a density of 5,000 cells/insert. Then 200 μl and 600 μl DMEM supplemented with 1% FBS was added into the insert and lower chamber, respectively. Subsequently, two chondro-TCP or bare-TCP constructs were placed in the lower chambers. After 24 h culture in a humidified incubator (37 °C, 5% CO2), the inserts were removed and the upper side of the membrane was scraped to remove adherent cells. Further, they were fixed in 10% formalin for 30 min, and stained with toluidine blue (0.05%) for 30 min to visualize the migrated cells on the lower side of the membrane. All samples were then photographed with Zeiss Imager Z1 equiped with AxioCam MRc5 camera (Carl Zeiss Microimaging GmbH, Göttingen, Germany). The number of the migrated cells was quantified based on the obtained images using Image J software (National Institutes of Health, Bethesda, USA). Figure 1. Scheme of co-culture systems. 2.4 Osteogenic differentiation of MSCs To further investigate the influence of cartilage constructs on the osteogenic differentiation of MSCs, a second indirect co-culture system with smaller membrane pore size was established (Fig. 1b). MSCs were seeded in the 24-well plate at a density of 100,000 cells per well. Afterwards, two chondro-TCP or bareTCP constructs were added into the top inserts (ThinCertTM cell culture inserts, membrane pore size 0.4 μm, Greiner Bio-One) overlying the MSCs. Two-hundred μl and 800 μl DMEM supplemented with 1% FBS was added into the insert and the 136 Roles of Cartilage Templates in Mediating MSCs Behavior lower well, respectively. After 4 days, MSCs in the lower well were collected for the following tests. 2.4.1 PCR analysis of osteogenic-related genes RNA of MSCs was extracted using TRIzol® reagent according to the manufacturer’s instructions (Life Technologies, Carlsbad, USA). The obtained RNA (n=3) was dissolved in RNase free water and the concentration was measured with spectrophotometer (NanoDrop 2000, Thermal scientific, Wilmington, USA). Subsequently, first strand cDNA was reverse transcribed from RNA using iScript™ cDNA Synthesis Kit (Bio-Rad Laboratories, Inc., USA). Afterwards cDNA was further amplified and the expression of specific genes was quantified using IQ SYBR Green Supermix PCR kit (BioRad, Hemel Hempstead, United Kingdom) in a real-time PCR (BioRad, CFX96™ real-time system) (n=4). Osteogenic-related gene markers were evaluated, including runt-related transcription factor 2 (Runx2), collagen type 1 (Col1α1), alkaline phosphatase (Alp), and bone sialoprotein (Bsp) (Table 1). The specificity of the primers was confirmed separately before the real-time PCR reaction. The expression levels were compared to the housekeeping gene β-Actin and calculated via the 2-∆∆Ct method [7]. Table 1. Overview of the primer sequences Forward (5’ → 3’) Reverse (5’ → 3’) Runx-2 GAGCACAAACATGGCTGAGA TGGAGATGTTGCTCTGTTCG Col1α1 GAGCGATTACTACTGGATTGACCC CAAGGAATGGCAGGCGAGAT Alp GGGACTGGTACTCGGATAACGA CTGATATGCGATGTCCTTGCA Bsp TCCTCCTCTGAAACGGTTTCC GGAACTATCGCCGTCTCCATT β-Actin TTCAACACCCCAGCCATGT TGTGGTACGACCAGAGGCATAC 7 2.4.2 DNA content and ALP activity DNA content (n=3) of MSCS was quantified by QuantiFluorTM DNA system (Promega Corporation, Madison, USA) according to the manufacturer’s instructions. Culture medium and upper inserts was removed, and cells were washed twice with PBS. One ml of milliQ water was added to each well, after which 2 freeze-thaw cycles were performed. A DNA standard curve was made from the Picogreen assay kit (Quant-iT PicoGreen dsDNA, Invitrogen). 100 μl sample or standard were added to 100 μl freshly made working solution in the wells. The results were read using a fluorescence microplate reader (Bio-Tek Instruments, Abcoude, The Netherlands) with an excitation wavelength at 485nm and an emission wavelength at 530nm. 137 Chapter 7 The ALP activity (n=3) was measured as a marker for osteogenic differentiation using the same cell extracts as for the DNA assay. 80 μl of sample or standard and 20μl buffer solution (5 mM MgCl2, 0.5 M 2-amino-2-methyl-1-propanol) was pipetted into a 96-well plate, and 100μl of substrate solution (5 mM paranitrophenylphosphate) was added per well. Subsequently, the plate was incubated for 1 h at 37 °C, after which the reaction was stopped by adding 100 μl of 0.3M NaOH. Serial dilutions of 4-nitrophenol (0–250 ng/ml) were used for making the standard curve. The plate was read in an ELISA reader (Bio-Tek Instruments) at 405 nm. ALP activity results were normalized by the amount of DNA. 2.5 PCR array of chemokine signaling related genes The expression profile of chemokine signaling related genes of MSCs was analyzed by pathway based quantitative PCR arrays (RT2Profiler PCR arrays, Rat Chemokines & Receptors, no. PAMM-022, SABiosciences) using the same RNA extration as described in 2.4.1. RT2 real-time SYBR Green PCR master mix (SABiosciences) was applied according to the manufacturer’s instructions. The tested genes for the PCR array were listed in Supplementary Table 1. The protocol used for PCR array was 95°C for 10 min, followed by 40 cycles of 95 °C for 15 sec and 60°C for 1 min. The data were analyzed using the SABiosciences PCR Array Data Analysis Software. Figure 2. Cell migration. Compared to bare-TCP scaffolds (c), chondro-TCP constructs (b) significantly promoted MSC migration, which resulted in a 4-fold change (p<0.05) of cell number across the transwell membrane (a). (staining: toluidine blue; error bars represent standard deviation of the mean, n=3) 3. Results 3.1 MSCs migration In vitro rat MSCs recruitment was tested using a transwell system. Toluidine blue staining showed that chondrogenically differentiated HA/TCP-cell constructs were 138 Roles of Cartilage Templates in Mediating MSCs Behavior able to induce MSCs migration across the transwell membrane (Fig. 2). Compared to the bare-TCP scaffolds, chondro-TCP significantly enhanced rat MSCs migration (p<0.05) by 4-fold. 3.2 Expression of osteogenic genes The expression of osteogenic gene markers of MSCs was detected after 4 days of co-culture. As shown in Fig. 3, for all detected genes, including Runx2, Col1a1, Alp and Bsp, there was no significant difference between MSCs co-cultured with chondro-TCP and MSCs with bare-TCP constructs. Figure 3. Osteogenic gene expressions of MSCs after 4 days of co-culture. 7 Figure 4. ALP activity (a) and DNA content (b) of MSCs after 4 days of co-culture. 139 Chapter 7 3.3 DNA content and ALP activity After 4 days of co-culture, ALP activity could be detected from MSCs which were co-cultured with chondo-TCP or bare-TCP constructs, and the amounts were comparable. Similar DNA content, which is related to the cell number, was also shown for these two groups (Fig. 4). 3.4 Analysis of chemokine signaling pathway PCR array was used to identify the genes that might be involved in the behavior change of MSCs in response to the cartilage template. The result revealed that 4-day co-culture of MSCs with chondro-TCP resulted in the expression level alteration of a number of genes for MSCs (Fig. 5 and supplementary Table. 1), among which several chemokines, i.e. Ccl3, Ccl7, Ccl19 and Nfkb1, and receptors, i.e. Cxcr2 and Tlr2, were considerably up-regulated (gene fold change > 4) (Table. 2). Figure 5. PCR array of chemokine signaling pathway. Each square represents one tested gene and its color indicates the gene fold change (up- or down-regulation) of MSCs co-cultured with chondro-TCP compared to MSCs co-cultured with bare-TCP. 4. Discussion Although endochondral ossification approach showed significant superiority in generating vascularized bone for tissue engineering applications, costly and lengthy process for creating cartilage template hampers its translation to clinic. A more comprehensive understanding of the crosstalk between implanted cartilage 140 Roles of Cartilage Templates in Mediating MSCs Behavior template and host cells is essentially beneficial to possibly simplify this procedure or develop substitutes for effective bone regeneration. Therefore, the aim of the current study was to analyze the effect of cartilage template on bone marrow derived MSCs and also the possible signaling pathways involved in these biological events. Table 2. Overview of up- and down-regulation of the genes in PCR array (Numbers indicate the gene fold change compared to the control group). Up-regulation (gene fold change > 2) Down-regulation (gene fold change > 2) Ccl19 (14.5), Tlr2 (10.1), Cxcr2 (8.5), Ccl3 (8.1), Ccl7 Bmp10 (5.0), Il13 (2.11) (7.1), Nfkb1 (4.8), Ccbp2 (3.5), Cmklr1 (2.8), Ccl6 (2.8), Tnf (2.7), Csf2 (2.5), Bdnf (2.3), Cxcl13 (2.3), C5 (2.2), Gdf5 (2.1) Our results showed that cartilage templates (chondro-TCP constructs) were able to induce MSCs migration when tested using a transwell culture system, which suggested their possible role for MSC recruitment in vivo. It is known that interactions of chemokines and their receptors mediate the migration of MSCs. Among various chemokines, stromal cell-derived factor-1 (SDF-1), also known as CXCL12, has been reported to play a pivotal role in MSCs homing and modulating [8, 9] . SDF-1 activates cell recruitment mainly through its receptor CXCR4 [10]. The interaction between SDF-1 and CXCR4 also stimulates the release of matrix metalloproteinases (MMPs) [11, 12], which are closely involved in the remodeling of cartilage matrix [13]. However, in our study, the presence of chondro-TCP constructs did not activate higher CXCR4 expression of the tested MSCs compared to the bareTCP control. Instead, a significant up-regulation was found for CXCR2 expression, the receptor of interleukin 8 (IL8), indicating that the observed MSCs migration might be associated with IL8 signaling. This result is supported by some previous investigations, which also demonstrated the effect of IL8 on MSCs homing [8, 14]. Overall, PCR array results demonstrated the up-regulation of a number of chemokine signaling related genes, indicating MSCs were actively stimulated by the cartilage templates. Amongst the significantly up-regulated genes, CCL7 (also known as monocyte-specific chemokine 3; MCP-3), CCL 19 (also known as macrophage inflammatory protein-3β; MIP-3β), CCL3 (also known as macrophage inflammatory protein-1α; MIP-1α), and NF-κB (nuclear factor kappa B) were of particular interest. The involvement of CCL 3, CCL7 and CCL19 in migration of MSCs and hematopoietic stem cells (HSCs) have been reported previously [15, 16]. Up-regulation of these three genes suggested that, besides being recruited by 141 7 Chapter 7 the cartilage templates, MSCs were also stimulated to express more inducers to recruit additional cells participating in the successive actions. In addition to cell recruitment, the up-regulated genes (CCL3, CCL19 and NF-κB) are also closely related to the regulation of osteoclast functions [17-19]. In particular, the activation of NF-κB pathway, which is commonly mediated by RANKL (receptor activator of the NF- κB ligand), might significantly promote osteoclastogenesis [19]. As a result, the remodeling of cartilage matrix could be directly enabled. These data suggested that, short-term release of soluble mediators from cartilage templates were able to effectively trigger a cascade of genes activations of MSCs, which would potentially contribute to the bone forming process by recruiting additional cells and regulating osteoclast activities. Regarding the osteogenic differentiation of MSCs, the presence of chondroTCP constructs did not show any significant effects in a short term. This was evidenced by the PCR results that comparable osteogenic gene expressions and ALP activities were detected for both chondro-TCP and bare-TCP group after 4 days of co-culture. Despite all that, the cartilage templates might influence the osteogenic differentiation of MSCs at a later time point. As demonstrated in some previous studies, chondrocytes were able to significantly induce the osteogenic differentiation of MSCs by continuous co-culture of 21 days [20, 21]. Another possibility is, the released soluble factors alone might not be sufficient to support osteo-induction of MSCs. A stronger osteogenic induction for MSCs would possibly necessitate soluble factors in conjunction with a direct cartilage-MSCs contact [22]. Some technical remarks should be made regarding the study set-up. Although coculture systems have been commonly explored in vitro to investigate the interactions between different cell populations in a manner of maximally simulating the in vivo complexity, it is still very difficult to replicate the real situation. One of the main concerns is the choice of co-culture medium. Varied culture conditions, i.e. fullor half-supplemented culture medium (chondrogenic or osteogenic), have been applied previously in co-cultures studying the interaction between chondrocytes and MSCs [21, 23]. However, this would inevitably bring too many variables, and also have the risk in eliciting artificially high response of MSCs from the supplements. Therefore, in our study, the co-culture was performed in a plain medium. In this way, the low-serum culture condition for cartilage template could be maintained, and the effects of cartilage templates on MSCs could be evaluated without the possible disturbances from the osteogenic supplements (i.e. ascorbic acid, Naβ-glycerolphosphate or dexamethasone). Nevertheless, it has also to be noted that such low nutrition culture condition, in some extent, might also influence the survival and function of both MSCs and cartilage templates. Therefore, the 142 Roles of Cartilage Templates in Mediating MSCs Behavior exploration of a more suitable co-culture condition (or system) is required for the future studies. Besides the soluble mediators, follow-up studies are warranted to investigate the influence of cartilage templates on MSCs in form of direct contact. Another limitation of our study is, the co-culture was only performed in a short period due to the consideration of cell survival in the plain culture medium. Therefore, long-term effect of cartilage constructs on MSCs need to be further investigated. Moreover, as the in vivo endochondral bone forming process involves several other cell types in addition to MSCs and chondrocytes, e.g. a variety of hematopoietic cells, a more comprehensive understanding about the interaction between cartilage template and other cell types is also necessary. When the crosstalk between implanted cartilage and host cells is thoroughly unveiled, endochondral bone engineering can be further simplified and more effectively applied, i.e. by using scaffolds loaded with biological cues that provoke host cell performance without implantation of exogenous cells. The data of this study represent the first step in this process. 5. Conclusion By using an indirect co-culture system, our results demonstrated that chondrogenically differentiated MSC-HA/TCP constructs (chondro-TCP) were able to induce MSCs migration by possibly activating the IL8/CXCR2 signaling pathway. Additionally, they could activate a considerable elevation of signaling factors responsible for MSCs recruitment and osteoclast stimulation, including CCL3, CCL7, CCL19 and NF-κB. The osteogenic differentiation of MSCs was not significantly influenced after short exposure to chondro-TCP templates. This study represents the first step determining the possible signaling pathways through which the cartilage templates exert their functions in generating endochondral bone tissue. Continued research in this direction would open the attractive possibility for the development of alternative “cartilage templates” to target endochondral bone regeneration. References 1. 2. Kronenberg, H.M., Developmental regulation of the growth plate. Nature, 2003. 423(6937): p. 332-6. Gawlitta, D., Farrell, E., Malda, J., Creemers, L.B., Alblas, J., and Dhert, W.J., Modulating endochondral ossification of multipotent stromal cells for bone regeneration. Tissue Eng Part B Rev. 16(4): p. 385-95. 143 7 Chapter 7 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 144 Scotti, C., Tonnarelli, B., Papadimitropoulos, A., Scherberich, A., Schaeren, S., Schauerte, A., Lopez-Rios, J., Zeller, R., Barbero, A., and Martin, I., Recapitulation of endochondral bone formation using human adult mesenchymal stem cells as a paradigm for developmental engineering. Proc Natl Acad Sci U S A, 2010. 107(16): p. 7251-6. Farrell, E., van der Jagt, O.P., Koevoet, W., Kops, N., van Manen, C.J., Hellingman, C.A., Jahr, H., O’Brien, F.J., Verhaar, J.A., Weinans, H., and van Osch, G.J., Chondrogenic priming of human bone marrow stromal cells: a better route to bone repair? Tissue Eng Part C Methods, 2009. 15(2): p. 285-95. Yang, W., Yang, F., Wang, Y., Both, S.K., and Jansen, J.A., In vivo bone generation via the endochondral pathway on three-dimensional electrospun fibers. Acta Biomater, 2013. 9(1): p. 4505-12. Tortelli, F., Tasso, R., Loiacono, F., and Cancedda, R., The development of tissue-engineered bone of different origin through endochondral and intramembranous ossification following the implantation of mesenchymal stem cells and osteoblasts in a murine model. Biomaterials, 2010. 31(2): p. 242-9. Livak, K.J. and Schmittgen, T.D., Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods, 2001. 25(4): p. 402-8. Chen, F.M., Wu, L.A., Zhang, M., Zhang, R., and Sun, H.H., Homing of endogenous stem/ progenitor cells for in situ tissue regeneration: Promises, strategies, and translational perspectives. Biomaterials, 2011. 32(12): p. 3189-209. Honczarenko, M., Le, Y., Swierkowski, M., Ghiran, I., Glodek, A.M., and Silberstein, L.E., Human bone marrow stromal cells express a distinct set of biologically functional chemokine receptors. Stem Cells, 2006. 24(4): p. 1030-41. Peled, A., Petit, I., Kollet, O., Magid, M., Ponomaryov, T., Byk, T., Nagler, A., Ben-Hur, H., Many, A., Shultz, L., Lider, O., Alon, R., Zipori, D., and Lapidot, T., Dependence of human stem cell engraftment and repopulation of NOD/SCID mice on CXCR4. Science, 1999. 283(5403): p. 845-8. Chiu, Y.C., Yang, R.S., Hsieh, K.H., Fong, Y.C., Way, T.D., Lee, T.S., Wu, H.C., Fu, W.M., and Tang, C.H., Stromal cell-derived factor-1 induces matrix metalloprotease-13 expression in human chondrocytes. Mol Pharmacol, 2007. 72(3): p. 695-703. Kanbe, K., Takagishi, K., and Chen, Q., Stimulation of matrix metalloprotease 3 release from human chondrocytes by the interaction of stromal cell-derived factor 1 and CXC chemokine receptor 4. Arthritis Rheum, 2002. 46(1): p. 130-7. Sternlicht, M.D. and Werb, Z., How matrix metalloproteinases regulate cell behavior. Annu Rev Cell Dev Biol, 2001. 17: p. 463-516. Anton, K., Banerjee, D., and Glod, J., Macrophage-associated mesenchymal stem cells assume an activated, migratory, pro-inflammatory phenotype with increased IL-6 and CXCL10 secretion. PLoS One, 2012. 7(4): p. e35036. Schenk, S., Mal, N., Finan, A., Zhang, M., Kiedrowski, M., Popovic, Z., McCarthy, P.M., and Penn, M.S., Monocyte chemotactic protein-3 is a myocardial mesenchymal stem cell homing factor. Stem Cells, 2007. 25(1): p. 245-51. Sordi, V., Malosio, M.L., Marchesi, F., Mercalli, A., Melzi, R., Giordano, T., Belmonte, N., Ferrari, G., Leone, B.E., Bertuzzi, F., Zerbini, G., Allavena, P., Bonifacio, E., and Piemonti, L., Bone marrow mesenchymal stem cells express a restricted set of functionally active chemokine receptors capable of promoting migration to pancreatic islets. Blood, 2005. 106(2): p. 419-27. Oba, Y., Lee, J.W., Ehrlich, L.A., Chung, H.Y., Jelinek, D.F., Callander, N.S., Horuk, R., Choi, S.J., and Roodman, G.D., MIP-1alpha utilizes both CCR1 and CCR5 to induce osteoclast formation and increase adhesion of myeloma cells to marrow stromal cells. Exp Hematol, 2005. 33(3): p. 2728. Toh, K., Kukita, T., Wu, Z., Kukita, A., Sandra, F., Tang, Q.Y., Nomiyama, H., and Iijima, T., Possible involvement of MIP-1alpha in the recruitment of osteoclast progenitors to the distal tibia in rats with adjuvant-induced arthritis. Lab Invest, 2004. 84(9): p. 1092-102. Nakamura, H., Aoki, K., Masuda, W., Alles, N., Nagano, K., Fukushima, H., Osawa, K., Yasuda, H., Nakamura, I., Mikuni-Takagaki, Y., Ohya, K., Maki, K., and Jimi, E., Disruption of NF-kappaB1 prevents bone loss caused by mechanical unloading. J Bone Miner Res, 2013. 28(6): p. 1457-67. Roles of Cartilage Templates in Mediating MSCs Behavior 20. Gerstenfeld, L.C., Cruceta, J., Shea, C.M., Sampath, K., Barnes, G.L., and Einhorn, T.A., Chondrocytes provide morphogenic signals that selectively induce osteogenic differentiation of mesenchymal stem cells. J Bone Miner Res, 2002. 17(2): p. 221-30. 21. Thompson, A.D., Betz, M.W., Yoon, D.M., and Fisher, J.P., Osteogenic differentiation of bone marrow stromal cells induced by coculture with chondrocytes encapsulated in threedimensional matrices. Tissue Eng Part A, 2009. 15(5): p. 1181-90. 22. Ball, S.G., Shuttleworth, A.C., and Kielty, C.M., Direct cell contact influences bone marrow mesenchymal stem cell fate. Int J Biochem Cell Biol, 2004. 36(4): p. 714-27. 23. Acharya, C., Adesida, A., Zajac, P., Mumme, M., Riesle, J., Martin, I., and Barbero, A., Enhanced chondrocyte proliferation and mesenchymal stromal cells chondrogenesis in coculture pellets mediate improved cartilage formation. J Cell Physiol, 2012. 227(1): p. 88-97. 7 145 Chapter 7 Supplementary Table 1: Gene table of PCR array (rat chemokines & receptors) Position Gene Description symbol Gene Fold up- or reference down- NM_031349 NM_012513 NM_053303 NM_001031824 NM_021670 NM_013107 NM_016994 XM_342421 NM_078621 NM_019205 NM_001105822 NM_057151 NM_001108661 NM_031530 NM_019233 NM_013025 NM_053858 NM_031116 NM_001004202 NM_001007612 NM_001012357 NM_020542 NM_001106872 NM_021866 NM_053958 NM_133532 NM_053960 NM_001013145 NM_199489 XM_236704 NM_172329 NM_001108191 NM_001013142 NM_001106034 NM_022218 NM_053352 NM_001106164 NM_023981 XM_340799 NM_134455 NM_133534 NM_030845 NM_139089 NM_182952 NM_053647 NM_001007729 NM_022214 NM_153721 2 2.29 1.85 -5.01 2 -1.62 1.69 2.19 3.52 1.2 2 2 14.48 2 1.38 8.13 1.12 1.7 2.8 7.14 2 -1.15 -1.06 2 1.01 2 1.1 2 2 2 2 2 -1.18 1.99 2.75 1.37 1.28 1.8 2.5 -1.27 1 2 1.13 1.28 -1.42 1.19 2 1.41 regulation A01 A02 A03 A04 A05 A06 A07 A08 A09 A10 A11 A12 B01 B02 B03 B04 B05 B06 B07 B08 B09 B10 B11 B12 C01 C02 C03 C04 C05 C06 C07 C08 C09 C10 C11 C12 D01 D02 D03 D04 D05 D06 D07 D08 D09 D10 D11 D12 146 Aplnr Bdnf Cxcr5 Bmp10 Bmp15 Bmp6 C3 C5 Ccbp2 Ccl11 Ccl12 Ccl17 Ccl19 Ccl2 Ccl20 Ccl3 Ccl4 Ccl5 Ccl6 Ccl7 Ccl9 Ccr1 Ccr1l1 Ccr2 Ccr3 Ccr4 Ccr5 Ccr6 Ccr7 Ccr8 Ccr9 Ccrl2 Cmtm2a Cmtm5 Cmklr1 Cxcr7 Cmtm3 Csf1 Csf2 Cx3cl1 Cx3cr1 Cxcl1 Cxcl10 Cxcl11 Cxcl2 Pf4 Cxcl5 Ppbp Apelin receptor Brain-derived neurotrophic factor Chemokine (C-X-C motif) receptor 5 Bone morphogenetic protein 10 Bone morphogenetic protein 15 Bone morphogenetic protein 6 Complement component 3 Complement component 5 Chemokine binding protein 2 Chemokine (C-C motif) ligand 11 Chemokine (C-C motif) ligand 12 Chemokine (C-C motif) ligand 17 Chemokine (C-C motif) ligand 19 Chemokine (C-C motif) ligand 2 Chemokine (C-C motif) ligand 20 Chemokine (C-C motif) ligand 3 Chemokine (C-C motif) ligand 4 Chemokine (C-C motif) ligand 5 Chemokine (C-C motif) ligand 6 Chemokine (C-C motif) ligand 7 Chemokine (C-C motif) ligand 9 Chemokine (C-C motif) receptor 1 Chemokine (C-C motif) receptor 1-like 1 Chemokine (C-C motif) receptor 2 Chemokine (C-C motif) receptor 3 Chemokine (C-C motif) receptor 4 Chemokine (C-C motif) receptor 5 Chemokine (C-C motif) receptor 6 Chemokine (C-C motif) receptor 7 Chemokine (C-C motif) receptor 8 Chemokine (C-C motif) receptor 9 Chemokine (C-C motif) receptor-like 2 CKLF-like MARVEL transmembrane domain containing 2A CKLF-like MARVEL transmembrane domain containing 5 Chemokine-like receptor 1 Chemokine (C-X-C motif) receptor 7 CKLF-like MARVEL transmembrane domain containing 3 Colony stimulating factor 1 (macrophage) Colony stimulating factor 2 (granulocyte-macrophage) Chemokine (C-X3-C motif) ligand 1 Chemokine (C-X3-C motif) receptor 1 Chemokine (C-X-C motif) ligand 1 Chemokine (C-X-C motif) ligand 10 Chemokine (C-X-C motif) ligand 11 Chemokine (C-X-C motif) ligand 2 Platelet factor 4 Chemokine (C-X-C motif) ligand 5 Pro-platelet basic protein (chemokine (C-X-C motif) ligand 7) Roles of Cartilage Templates in Mediating MSCs Behavior E01 E02 E03 E04 E05 E06 E07 E08 E09 E10 E11 E12 F01 F02 F03 F04 F05 F06 F07 F08 F09 F10 F11 F12 G01 Cxcl9 Cxcr3 Cxcr4 Tymp Epo Gdf5 Ccr10 Gpr77 Hif1a Il13 Il16 Il18 Il1a Il4 Il8ra Cxcr2 Inha Inhbb Lif Cxcl13 Ltb4r2 Mmp2 Mmp7 Myd88 Nfkb1 Chemokine (C-X-C motif) ligand 9 Chemokine (C-X-C motif) receptor 3 Chemokine (C-X-C motif) receptor 4 Thymidine phosphorylase Erythropoietin Growth differentiation factor 5 Chemokine (C-C motif) receptor 10 G protein-coupled receptor 77 Hypoxia-inducible factor 1, alpha subunit Interleukin 13 Interleukin 16 Interleukin 18 Interleukin 1 alpha Interleukin 4 Interleukin 8 receptor, alpha Chemokine (C-X-C motif) receptor 2 Inhibin alpha Inhibin beta-B Leukemia inhibitory factor Chemokine (C-X-C motif) ligand 13 Leukotriene B4 receptor 2 Matrix metallopeptidase 2 Matrix metallopeptidase 7 Myeloid differentiation primary response gene 88 Nuclear factor of kappa light polypeptide gene enhancer in NM_145672 NM_053415 NM_022205 NM_001012122 NM_017001 XM_001066344 NM_001108836 NM_001003710 NM_024359 NM_053828 NM_001105749 NM_019165 NM_017019 NM_201270 NM_019310 NM_017183 NM_012590 NM_080771 NM_022196 NM_001017496 NM_053640 NM_031054 NM_012864 NM_198130 XM_342346 2 2 2 2 -1.05 2.14 1.71 1.21 1.42 -2.11 1.45 1.25 2 1.13 -1.14 8.49 2 -1.11 2 2.25 1.85 2 1.22 1.77 4.79 G02 G03 G04 G05 G06 G07 G08 G09 G10 G11 G12 Rgs3 Sdf2 Slit2 Tapbp Tcp10b Tlr2 Tlr4 Tnf Tnfrsf1a Trem1 Xcl1 B-cells Regulator of G-protein signaling 3 Stromal cell derived factor 2 Slit homolog 2 (Drosophila) TAP binding protein T-complex protein 10b Toll-like receptor 2 Toll-like receptor 4 Tumor necrosis factor (TNF superfamily, member 2) Tumor necrosis factor receptor superfamily, member 1a Triggering receptor expressed on myeloid cells 1 Chemokine (C motif) ligand 1 NM_019340 NM_001105803 NM_022632 NM_033098 NM_001107461 NM_198769 NM_019178 NM_012675 NM_013091 NM_001106885 NM_134361 1.53 2 1.42 -1.06 1.4 10.09 1.12 2.66 -1.78 2.46 2 7 147 Chapter 8 Summary, closing remarks and future perspectives Summary, Closing Remarks and Future Perspectives 1. Summary and address to the aims Although great challenges remain, bone tissue engineering presents a promising strategy to replace the current autografts or allografts for the treatment of bone defects. The standard procedure of bone tissue engineering involves loading osteoprogenitor cells onto proper designed biomaterials to induce cell differentiation in a specific pathway, followed by transplantation of this cellscaffold complex into the defect site to regenerate bone tissue. Such a construct aims not only to replace the injured tissue, but also to induce and direct the healing process, and completely restore the functions of the native tissue. So far, the most common approach to generate bone has largely focused on intramembranous ossification (i.e. bone formation through direct osteoblastic differentiation), however without obtaining convincing results in humans. Until recently, a new approach, resembling endochondral ossification, has drawn attention. With this approach, bone is formed via the remodeling of a cartilage intermediate. Compared to osteogenically differentiated cells, chondrocytes are well adapted to limited nutrition supply and hypoxic conditions. In addition, they can produce considerable amounts of angiogenic factors during differentiation. These features give the endochondral approach a great advantage in maintaining viability of the implanted cells and generating vascularized bone tissue. Yet the available knowledge and investigations on this endochondral approach is still in its infancy. This thesis aimed to investigate and optimize the endochondral ossification approach to achieve high-quality bone regeneration. Specifically, its applicability with different scaffold materials, optimal in vitro priming time, as well as the mechanism of the transformation from cartilage to bone, were explored in a series of studies. Chapter 1 gives a general introduction on bone tissue engineering and its essential elements, as well as the aims of this thesis. Each following chapter presents a separated study. This summary addresses the research questions as described in the first chapter in successive order. 1. What is the current state of the art of combining stem cells and biodegradable materials for bone regeneration via the endochondral pathway? Thanks to the improving knowledge of stem cell sources, together with profound advances in materials sciences, the “engineering” of man-made bone grafts appears to be increasingly feasible to target the regeneration of large bone defects. Chapter 2 first elaborates on bone tissue and osteogenesis mechanisms, in order to reach a fundamental understanding of the processes involved in tissue engineering approaches for bone regeneration. Next, the preliminary investigations in utilizing biomaterials and cell culture strategies to achieve endochondral bone formation were highlighted. Different cells sources, e.g. chondrocytes, MSCs and ESCs have 151 8 Chapter 8 been reported to posses the ability of undergoing the endochondral differentiation pathway with or without specific stimulations. A variety of biomaterials, with the structure and composition closely mimicking the nature cartilage or bone tissue, are potential candidates to support the generation of osteochondral tissue, among which natural or synthetic polymers and ceramics are the top choices. The superiority of generating vascularized bone has been revealed by using the endochondral strategy, albeit standardization and optimization of this approach are essentially needed to achieve an efficient transformation from cartilage to bone tissue. Moreover, a number of issues still need to be addressed before a clinical trial can be started, including the appropriate time point for implantation, function and fate of implanted cells. 2. Is the porous aliphatic polyurethane/hydroxyapatite composite a suitable scaffold for bone tissue engineering applications? Biomaterial scaffolds meant to function as supporting structures to osteogenic cells play a pivotal role in bone tissue engineering. In Chapter 3, we synthesized an aliphatic polyurethane (PU) scaffold via a foaming method using non-toxic components. Through this procedure a uniform interconnected porous structure was created. Furthermore, hydroxyapatite (HA) particles were introduced into this process to increase the bioactivity of the PU matrix. To evaluate the biological performances of these PU based scaffolds, their influence on in vitro cellular behavior and in vivo bone forming capacity was investigated in this study. A simulated body fluid (SBF) test demonstrated that the incorporation of 40wt% HA particles significantly promoted the biomineralization ability of the PU scaffolds. Enhanced in vitro proliferation and osteogenic differentiation of the seeded MSCs were also observed on the PU/HA composite. After being implanted subcutaneously in a nude mice model for 8 weeks, a considerable amount of vascularized bone tissue with initial marrow stroma development was generated in both PU and PU/ HA40 scaffold. In conclusion, the PU/HA composite is a potential scaffold for bone regeneration applications. 3. Is it possible to fabricate a 3D scaffold by electrospinning technique and utilize it as cell carrier to enable effective in vivo endochondral bone generation? To further explore candidate scaffolds for endochondral bone tissue engineering applications, in Chapter 4, a novel wet-electrospinning system, using ethanol as collecting medium, was exploited to fabricate a cotton-like poly(lactic-coglycolic acid) (PLGA)/poly(ε-caprolactone) (PCL) scaffold. Such PLGA/PCL scaffold consisted of an accumulation of fibers in a very loose and uncompressed manner. Rat MSCs were seeded on these scaffolds and chondrogenically differentiated in vitro for 4 weeks followed by subcutaneous implantation in vivo for 8 weeks. Cell 152 Summary, Closing Remarks and Future Perspectives pellets were used as a control. The glycosaminoglycan (GAG) assay and Safranin O staining showed that the cells infiltrated throughout the scaffolds and deposited an abundant cartilage matrix after in vitro chondrogenic priming. Histological analysis of the in vivo samples revealed extensive new bone formation through the remodeling of the cartilage template. In conclusion, using the wet-electrospinning method, we are able to create a 3D scaffold in which bone tissue can be formed via the endochondral pathway. This system can be easily processed for various assays and histological analysis. Consequently, it is more efficient than the traditional cell pellets as a tool to study endochondral bone formation for tissue engineering purposes. 4. Can in vivo endochondral bone generation be achieved in various scaffold materials? Chapter 5 investigated the behavior of bone marrow derived mesenchymal stem cells (MSCs) in regard to in vitro cartilage formation and in vivo bone regeneration when combined with different three-dimensional (3D) scaffold materials, i.e. hydroxyapatite/tricalcium phosphate (HA/TCP) composite block, polyurethane (PU) foam, poly(lactic-co-glycolic acid)/poly(ε-caprolactone) electrospun fibers (PLGA/PCL) and collagen I gel. To this end, rat MSCs were seeded on these scaffolds followed by the same experimental procedures as described in Chapter 4. After in vitro chondrogenic priming, comparable cell amounts and cartilage formation was observed in all types of scaffolds. Nonetheless, the quality and maturity of in vivo ectopic bone formation appeared to be scaffold/material-dependent. Eight weeks of implantation was not sufficient to ossify the entire PLGA/PCL constructs, albeit that a comprehensive remodeling of the cartilage had occurred. For HA/ TCP, PU and collagen I scaffolds, more mature bone formation with rich vascularity and marrow stroma development could be observed. These data suggest that chondrogenic priming of MSCs in the presence of different scaffold materials allows the establishment of reliable templates for generating functional endochondral bone tissue without using osteoinductive growth factors in vivo. The morphology and maturity of bone formation can be dictated by the scaffold properties. 5. What is the effect of in vitro chondrogenic priming time of MSCs on in vivo endochondral bone formation? To enable an effective translation of this endochondral approach towards the clinic, it is imperative to possibly shorten the in vitro procedure for the creation of a cartilage template and meanwhile ensure sufficient bone formation. Therefore, in Chapter 6, we aimed to investigate the influence of in vitro chondrogenic priming time of MSCs on the in vivo endochondral bone formation, both qualitatively and quantitatively. To this end, rat bone marrow derived mesenchymal stem cells 153 8 Chapter 8 were seeded on two scaffolds with distinguished features, i.e. a fibrous polymerbased PLGA/PCL scaffold or a porous ceramic-based HA/TCP composite. Next the constructs were chondrogenically differentiated for 2, 3 and 4 weeks in vitro followed by subcutaneous implantation for up to 8 weeks. The results showed that, a longer chondrogenic priming time resulted in a significantly increased amount and homogeneous deposition of the cartilage matrix on both PLGA/ PCL and HA/TCP scaffolds in vitro. In vivo, all implanted constructs gave rise to endochondral bone formation, whereas the bone volume was not affected by a longer priming time. An unpolarized woven bone like structure, with significant amounts of cartilage remaining, was generated and maintained in fibrous PLGA/ PCL scaffolds, while porous HA/TCP scaffolds supported progressive lamellar-like bone formation with mature bone marrow development. 6. What is the role of cartilage templates in mediating MSC behavior in the process of endochondral bone formation? Costly and lengthy process for creating cartilage templates hampers the translation of the endochondral approach to clinic. A more comprehensive understanding of the crosstalk between implanted cartilage template and host cells is essentially beneficial to possibly simplify this procedure or develop substitutes for effective bone regeneration. Therefore, in Chapter 7, the role of cartilage template in mediating MSC behavior in the process of endochondral bone formation was explored, as well as the possible signaling pathways involved in these biological events. By using an indirect coculture system, our results demonstrated that chondrogenically differentiated MSC-HA/TCP constructs (chondro-TCP) were able to induce MSCs migration by possibly activating the IL8/CXCR2 signaling pathway. Additionally, they could trigger a considerable elevation of signaling factors responsible for MSCs recruitment and osteoclast stimulation, including CCL3, CCL7, CCL19 and NF-κB. The osteogenic differentiation of MSCs was not significantly influenced after short exposure to chondro-TCP templates. 2. Closing remarks and future perspectives In the context of skeletal tissue development and repair, endochondral ossification has inspired a shift of the bone tissue engineering concept from the conventional intramembranous to the endochondral pathway. Although great challenges still remain before this strategy can be eventually progressed into clinical application, this thesis contributes to a further understanding of endochondral bone regeneration. Based on the obtained results, several key conclusions can be drawn: 154 Summary, Closing Remarks and Future Perspectives (i) Compared to the use of cell pellets, chondrogenic priming of bone marrow derived MSCs in the presence of scaffolds allows a more reliable and efficient establishment of templates to generate functional endochondral bone tissue in vivo, which can be achieved by using a variety of biomaterials. Among the tested scaffold materials in this thesis, collagen I hydrogel and porous calcium phosphate scaffold are the top choices to target a vascularized lamellar-like bone formation with mature marrow development. (ii) By applying a chondrogenically differentiated MSC-scaffold construct as a cartilage template, the essential in vitro priming time can be shortened to 2 weeks in order to generate a substantial amount of vascularized endochondral bone in vivo. This is closely contributed by the significant formation of cartilage tissue and expression of VEGF early to 2 weeks time point. Moreover, our data also implied that a certain extent of in vitro chondrogenic priming of the MSCs, possibly with even shorter time than 2 weeks, might be sufficient to ensure the progression of endochondral differentiation of the seeded cells and enable successful bone formation. If proved by the further investigations, this would lead to an advance for applying the endochondral approach in the clinical practice, as a shortened in vitro culture procedure would significantly reduce the treatment cost and regulatory burden. (iii) The properties of scaffold materials play an imperative role in dictating the progression and structure of the endochondral bone formation. Among which, biomaterial chemistry deeply affects the cell fate at early stages, but the internal architecture of the scaffolds also plays a profound role, especially in a long term, since it influences bone structure and its deposition kinetics. For both polymer (e.g. PU) and ceramic (e.g. HA/TCP) based scaffolds, a confined porous structure could concentrate collagen fibers and serve as a polarizing template within the pores to trigger the formation of lamellar like bone. On the contrary, fibrous scaffolds (e.g. PLGA/PCL) possessed a high level of geometrical complexity, thus the new collagen fibers were distributed in an isotropic manner, resulting mostly in woven bone. (iv) After implantation, a cartilage template is able to exert a timely influence on the recruitment of the host MSCs and the promotion of osteoclastogenesis by activating a number of chemokine signaling related factors. Further efforts to determine the signaling pathways involved in endochondral bone forming process will possibly allow the development of alternative “cartilage templates” to target bone regeneration in a simple and effective manner. 155 8 Chapter 8 For the ultimate success of translating the endochondral bone tissue engineering approach to clinical practice, a number of limitations and necessary investigations need to be addressed. First of all, it is important to keep in mind that all discoveries on endochondral bone tissue engineering so far emerged from cell culture and animal experiments, which cannot be directly translated to human situation. In our studies, rat bone marrow derived MSCs were used as the main cell source. However, for the final clinical translation, there is no doubt that the efficacy of utilizing human cells to generate endochondral bone has to be elucidated. Another issue that needs to be addressed is the selection of the animal model. So far we have only tested the in vivo bone formation in an ectopic site of immunocompromised animals. By significantly reducing the number of variables involved in bone formation, this model allows us to obtain an easier understanding of the tendency of cartilage template undergoing endochondral bone formation. Still, it has to be noted that the conclusions need to be further evidenced in a more clinically relevant situation, i.e. in an orthotopic location of immunocompetent animals, where the mechanical load, biochemical/inflammatory factors and endogenous progenitor/ stem cells are closely involved. Although our studies have confirmed that the design of the scaffold can drive the endochondral bone forming pattern, the specific parameters that dominate this process and its mechanism could not be thoroughly concluded due to the differences of the selected scaffolds regarding their chemical compositions and physical structures. More complete understanding could be obtained if the scaffold properties are more precisely controlled, e.g. all structural parameters of the scaffolds are kept constant and meanwhile only the composition is varied, or vice versa. In addition, larger-sized scaffolds have also to be tested in future studies, where the cell invasion and vascularization will be more challenging. When the interaction between the cells and scaffold materials is unraveled, we are able to target endochondral bone regeneration in an efficient manner and on a larger scale with the help of selecting and tailoring appropriate scaffolds. Furthermore, because high cell density is crucial for the formation of cartilage templates, the clinical application of the endochondral approach by using autologous cells can be challenging due to the limits of such cell source or complexity of the in vitro cell expansion procedure, especially when a large bone defect is involved. At this point, the underlying mechanism about how the implanted cells induce a specific response of the host cells and the successive tissue regeneration is still largely unknown. Further investigations are warranted to elucidate the cellular and molecular events involved in this process before a reliable endochondral 156 Summary, Closing Remarks and Future Perspectives therapy can be developed. More importantly, only when the crosstalk between the implanted cells and the host cells is thoroughly elucidated, endochondral bone tissue engineering can be further simplified and more effectively applied. In summary, bone regeneration via the endochondral pathway requires a deliberation of many factors, including a variety of cells, matrixes, and signals, in an orderly temporal and spatial sequence. Although extensive efforts still have to be devoted, endochondral bone tissue engineering holds great promise for the final clinical application. The ultimate goal to implement this approach is not to solely replicate the natural ossification process, but to recapitulate and extract the essence of this process to target bone regeneration in a simple, cost-effective and controllable way, e.g. by using scaffolds which releases various biological cues at right timing that provoke host cell performance to target endochondral bone formation without implantation of exogenous cells. For this to happen, a thorough understanding of the underlying cellular and molecular mechanism of the endochondral approach, including cell-cell and cell-matrix/material interactions, is required, which needs collaboration among many disciplines, including cell and molecular biology, material science, pharmacology, chemistry and medicine. 8 157 Chapter 9 Samenvatting, slotopmerkingen en toekomstperspectieven Samenvatting, Slotopmerkingen en Toekomstperspectieven 1. Samenvatting en evaluatie van de doelstellingen Hoewel er nog veel uitdagingen overwonnen moeten worden, is botweefselengineering een veelbelovende aanpak om de huidige behandelmethoden via autologe of allogene bottransplantatie te vervangen. De standaardprocedure die gebruikt wordt bij botweefselengineering is het laden van osteogene voorlopercellen op biomaterialen die de differentiatie van cellen in een specifieke richting sturen, gevolgd door de implantatie van deze constructen in het lichaam. Een dergelijke aanpak beoogt niet alleen het beschadigde weefsel te vervangen, maar induceert en dirigeert ook het genezingsproces om tot een volledige genezing van het aangedane weefsel te komen. De huidige methoden om bot te genereren zijn grotendeels gericht op intramembraneuze ossificatie (d.w.z. botvorming via osteoblast differentiatie), vooralsnog is de werking van deze methode echter nog niet op overtuigende wijze resultaat getoond bij mensen. Recentelijk heeft een nieuwe aanpak, gebaseerd op endochondrale ossificatie, de aandacht getrokken. Deze aanpak richt zich op het vormen van bot door middel van kraakbeen. Vergeleken met osteogeen gedifferentieerde cellen zijn chondrocyten goed aangepast aan hypoxische condities met daarbij een beperkt aanbod van nutriënten. Bovendien produceren chondrocyten tijdens de differentiatie aanzienlijke hoeveelheden angiogene factoren. Door deze eigenschappen heeft de endochondrale aanpak een groot voordeel bij het levensvatbaar houden van de geïmplanteerde cellen en het genereren van gevasculariseerd botweefsel. Het onderzoek en de beschikbare kennis over de endochondrale methode staan echter nog in de kinderschoenen. Het onderzoek beschreven in dit proefschrift heeft als doel om botvorming via de endochondrale methode te optimaliseren en zodoende een hoge kwaliteit botherstel te bereiken. Meer specifiek werd daarbij in verschillende studies de toepasbaarheid van verschillende scaffold materialen, optimale in vitro kweektijd en het mechanisme van de omzetting van kraakbeen naar bot onderzocht. Hoofdstuk 1 geeft een algemene inleiding over botweefselengineering, daarnaast bevat dit hoofdstuk de doelstellingen van dit proefschrift. Elk volgend hoofdstuk beschrijft een deelonderzoek. In deze samenvatting worden de vraagstellingen zoals beschreven in het eerste hoofdstuk in opeenvolgende volgorde behandeld. 1. Wat is de huidige stand van de techniek van het combineren van stamcellen en biologisch afbreekbare materialen voor botherstel via de endochondrale route? Dankzij verbeterde kennis over de bronnen van stamcellen, samen met een brede vooruitgang in de ontwikkeling van materialen, lijkt het erop dat “engineering” in toenemende mate toepasbaar is voor het regenereren van grote botdefecten. 161 9 Chapter 8 Hoofdstuk 2 gaat eerst dieper in op de eigenschappen en mechanismen van botweefsel en osteogenese, om zo een fundamenteel begrip van de betrokken processen te verkrijgen. Vervolgens worden de in de literatuur beschreven onderzoeken over de toepassing van biomaterialen en celcultuur strategieën om endochondrale botvorming te verkrijgen toegelicht. Van verschillende type cellen, waaronder chondrocyten, MSCs en ESCs, is het bekend dat ze het vermogen bezitten om endochondrale differentiatie te ondergaan, met of zonder specifieke stimulatie. Verschillende biomaterialen, met een structuur en samenstelling die nauwkeurig de aard kraakbeen of botweefsel nabootsen, zijn potentiële kandidaten voor de vorming van osteochondrale weefsels. Natuurlijke of synthetische polymeren en keramische materialen zijn daarbij de belangrijkste kandidaten. De mogelijkheid tot het genereren van gevasculariseerd bot is aangetoond via de endochondrale methode, hoewel er standaardisatie en optimalisatie nodig zal zijn om voor een efficiënte omzetting van kraakbeen naar botweefsel te zorgen. Voordat een klinische test uitgevoerd kan worden moeten bovendien nog een aantal zaken uitgezocht worden, waaronder het juiste tijdstip voor de implantatie, de functie en het lot van de geïmplanteerde cellen. 2. Is het poreuze alifatische polyurethaan/hydroxyapatiet composiet een geschikte scaffold voor botweefselengineering toepassingen? Biomaterialen met als doel te functioneren als ondersteunende structuur voor osteogene cellen spelen een centrale rol in botweefselengineering. In hoofdstuk 3 beschrijven we hoe een alifatische polyurethaan (PU) schuim gesynthetiseerd werd met behulp van niet-giftige bestanddelen. Via deze procedure werd een uniforme poreuze structuur gecreëerd, waarbij de holtes in de structuur onderling verbonden waren. Bovendien werden hydroxyapatiet (HA) deeltjes aan dit proces geïntroduceerd om de bioactiviteit van de PU matrix te verhogen. Om de biologische prestaties van deze PU scaffolds te evalueren, werd hun invloed op het cellulaire gedrag in vitro en de bot vormende capaciteit in vivo onderzocht. Een test met “simulated body fluid” (SBF) toonde aan dat het inbedden van 40 gew% HA deeltjes het vermogen tot biomineralisatie significant verhoogde. Verbeterde cel proliferatie en osteogene differentiatie van de gekweekte MSCs werden waargenomen op de PU/HA composiet. Na subcutane implantatie in een naakte muis model voor 8 weken, werd een aanzienlijke hoeveelheid gevasculariseerd botweefsel en beenmerg ontwikkeling gegenereerd bij zowel de PU als de PU/ HA40 scaffolds. De conclusie is dat het PU/HA composiet materiaal veel potentiee heeft voor botregeneratieve toepassingen. 162 Samenvatting, Slotopmerkingen en Toekomstperspectieven 3. Is het mogelijk om een 3D-scaffold te fabriceren via elektrospinning en deze te gebruiken als een drager materiaal voor de cellen om in vivo endochondrale bot generatie effectief mogelijk te maken? Om verder te onderzoeken welke scaffolds geschikt zijn voor endochondrale botweefselengineering toepassingen, is in hoofdstuk 4 beschreven hoe een nieuwe natte-elektrospin techniek werd ingezet. Hierbij werd gebruik gemaakt van ethanol als medium om de vezels in te verzamelen. Via deze methode werd een katoenachtige poly (melkzuur-co-glycolzuur) (PLGA)/poly (ε-caprolacton) (PCL) scaffold gefabriceerd, bestaande uit een opeenstapeling van vezels op een zeer losse ongecomprimeerde wijze. Op deze scaffolds werden vervolgens ratten beenmergcellen gezaaid en deze werden chondrogeen gedifferentieerd in vitro gedurende 4 weken, gevolgd door subcutane implantatie in vivo voor 8 weken. Daarbij werden cel pellets gebruikt als controle. De analyse van glycosaminoglycaan (GAG) en safranine O kleuring toonde aan dat na in vitro chondrogene priming de cellen de gehele matrix hadden geïnfiltreerd en daarbij een overvloedige kraakbeenmatrix hadden gedeponeerd. Histologische analyse van de in vivo constructen demonstreerde uitgebreide nieuwe botvorming door hermodellering van het kraakbeen. Samenvattend, gebruik makend van de natte electrospinning methode, kunnen wij een 3D scaffold creëren die helpt om via de endochondrale route bot te vormen. Dit systeem leent zich uitstekend voor analyse met verschillende assays en histologische analyse. Daarom is dit systeem efficiënter voor het bestuderen van weefselregeneratie via de endochondrale methode dan het gebruik van de traditionele cel paletten. 4. Kan via de endochondrale route bot worden gevormd in verschillende materialen? In hoofdstuk 5 wordt het gedrag van mesenchymale stamcellen (MSCs) uit het beenmerg onderzocht met betrekking tot de in vitro vorming van kraagbeen en in vivo botregeneratie in combinatie met verschillende driedimensionale (3D) materialen, waaronder: hydroxyapatiet/tricalciumfosfaat (HA/TCP), polyurethaan (PU) schuim, poly (melkzuur-co-glycolzuur)/poly (ε-caprolacton) vezels (PLGA/ PCL, bereid via electrospinning) en collageen-I gel. Hiertoe werden ratten MSCs gezaaid op deze scaffolds gevolgd door dezelfde experimentele procedure als beschreven in hoofdstuk 4. Vergelijkbare hoeveelheden cellen en kraakbeen werden waargenomen na in vitro chondrogene priming voor alle soorten scaffolds. Toch leek de kwaliteit en de maturatie van het in vivo ectopische gevormde bot materiaal afhankelijk. Een implantatietijd van acht weken was niet voldoende om de volledige PLGA/PCL constructen te ossificeren, maar het kraakbeen was wel duidelijk geremodelleerd. Voor de HA/TCP, PU en collageen I scaffolds, werd een 163 9 Chapter 8 meer volwassen botvorming met rijke vascularisatie en beenmerg ontwikkeling waargenomen. Deze gegevens suggereren dat chondrogene priming van MSCs in de aanwezigheid van verschillende scaffold materialen het genereren van functioneel botweefsel via de endochondrale route zonder osteoinductie groeifactoren in vivo mogelijk maakt. De morfologie en de maturatie van botvorming worden mede gedicteerd door de eigenschappen van het scaffold materiaal. 5. Wat is het effect van in vitro chondrogene priming van de MSCs op in vivo endochondrale botvorming? Om een effectieve behandeling van deze endochondrale aanpak klinisch mogelijk te maken, is het noodzakelijk om de in vitro procedure voor het fabriceren van een kraakbeen substraat zo kort mogelijk te houden en ondertussen te zorgen voor voldoende botvorming. In hoofdstuk 6 is daarom onderzocht wat de invloed is van de in vitro chondrogene primings tijd van MSCs op de in vivo endochondrale botvorming, zowel kwalitatief als kwantitatief. Hiervoor werden van ratten MSCs gezaaid op twee scaffolds met verschillende kenmerken; een vezelachtige PLGA/ PCL polymeer of een poreuze keramische, HA/TCP composiet. Vervolgens werden de constructen voor 2, 3 of 4 weken chondrogeen gedifferentieerd in vitro, gevolgd door subcutane implantatie tot 8 weken. De resultaten toonden aan dat een langere chondrogene primings tijd resulteerde in een significant verhoogde hoeveelheid afzetting van kraakbeenmatrix voor zowel de PLGA/PCL en HA/TCP materialen in vitro. In vivo, gaven alle geïmplanteerde substraten uiteindelijk aanleiding tot endochondrale botvorming, terwijl het bot volume niet werd beïnvloed door een langere priming tijd. Een ongepolariseerde gewoven bot structuur, waarbij nog aanzienlijke hoeveelheden kraakbeen zichtbaar waren, werd gegenereerd bij de vezelachtige PLGA/PCL scaffolds, terwijl de poreuze HA/TCP scaffolds lamellairachtige botvorming met een rijpere beenmerg ontwikkeling ondersteunden. 6. Op welke wijze beïnvloeden kraakbeen substraten het gedrag van MSCs tijdens het endochondrale proces? Een kostbaar en langdurig proces voor het maken van kraakbeen substraten belemmert de klinische haalbaarheid van de endochondrale methode. Een completer begrip van de communicatie tussen het geïmplanteerde kraakbeen substraat en de cellen van de ontvanger is noodzakelijk om deze methode verder te verbeteren. Daarom is in hoofdstuk 7 de invloed van het kraakbeen substraat op het gedrag van MSCs tijdens endochondrale botvorming onderzocht en werd daarnaast onderzocht welke biologische signalen mogelijk betrokken zijn bij dit proces. Door gebruikt te maken van een indirect cocultuur systeem, toonden onze resultaten aan dat chondrogeen gedifferentieerde MSC-HA/TCP substraten (chondro-TCP) migratie van MSCs mogelijk veroorzaken door het activeren van het 164 Samenvatting, Slotopmerkingen en Toekomstperspectieven IL8/CXCR2 signaal transductie route. Daarnaast werd een aanzienlijke verhoging van de signalerings factoren, zoals CCL3, CCL7, CCL19 en NF-kB, die verantwoordelijk zijn voor het rekruteren van MSCs en het stimuleren van osteoclasten gevonden. De osteogene differentiatie van MSCs werd niet significant beïnvloed na korte blootstelling aan de chondro-TCP templates. 2. Slotwoord en toekomstperspectieven In het kader van de ontwikkeling van het skelet en weefselherstel, inspireert de endochondrale ossificatie een verschuiving in de botweefselengineering van de conventionele intramembraneuze naar de endochondrale route. Hoewel er nog steeds grote uitdagingen voordat deze strategie uiteindelijk kan worden toegepast in de kliniek, draagt dit proefschrift bij aan een beter begrip van het endochondraal botherstel. Op basis van de verkregen resultaten kan een aantal belangrijke conclusies worden getrokken: (i) In vergelijking met het gebruik van celpellets, biedt chondrogene priming van MSCs uit het beenmerg in de aanwezigheid van materialen veel voordelen waaronder het creëren van een betrouwbaar en efficiëntesubstraat dat via de endochondraal route botweefsel kan genereren. Dit kan worden bereikt worden met een verscheidenheid aan biomaterialen. Onder de in dit proefschrift geteste materialen geven collageen-I hydrogelen en poreus calciumfosfaat de beste resultaten om een gevasculariseerd lamellair-achtige botvorming met volwassen beenmerg ontwikkeling te verkrijgen. (ii) Door gebruik te maken van een construct met chondrogeen gedifferentieerde MSCs als een kraakbeen substraat, kan de essentiële in vitro priming tijd om een aanzienlijke hoeveelheid gevasculariseerde bot te genereren verkort worden tot 2 weken. Dit was duidelijk zichtbaar door de aanzienlijke vorming van kraakbeenweefsel en de expressie van VEGF binnen 2 weken. Bovendien, suggereren de gegevens ook dat eventueel een nog kortere primings tijd dan 2 weken, kan volstaan voor voldoende chondrogene differentiatie van de gezaaide cellen en uiteindelijk dus botvorming. Als dit inderdaad bewezen wordt door de verdere onderzoeken, zou dit een groot voordeel zijn voor de toepassing van de endochondrale aanpak in de klinische praktijk, omdat de verkorte in vitro cultuur tijd de kosten en administratieve lasten van de behandeling aanzienlijk verminderen. (iii) De eigenschappen van scaffold materialen spelen een belangrijke rol in het dicteren van de progressie en de structuur van de endochondrale botvorming. De chemische eigenschappen van het biomateriaal spelen vooral een rol op de 165 9 Chapter 8 korte termijn, terwijl de interne structuur ervan meer invloed heeft op de lange termijn aangezien het de botstructuur en de depositie kinetiek beïnvloedt. Voor zowel polymeren (bijv. PU) als keramiek (bijv. HA/TCP) gebaseerde scaffolds, kan een structuur met kleine poriën collageenvezels concentreren en dienen als een polariserend sjabloon in de poriën en zo de vorming van lamellair bot induceren. Vezelige scaffolds (bijv. PLGA/PCL) daarentegen bezitten een hoge geometrische complexiteit, waardoor de nieuwe collageen vezels zich op een isotrope manier verspreiden, wat meestal leidt tot gewoven bot. (iv) Na implantatie is een kraakbeen substraat in staat om invloed uit te oefenen op de MSCs van de ontvanger en zo osteoclast vorming te induceren door het activeren van een aantal aan chemokines gerelateerde factoren. Verdere studies om meer begrip te krijgen voor de signaal routes betrokken bij endochondrale botvorming, kunnen mogelijk de ontwikkeling van alternatieve “kraakbeen substraten” naar botregeneratie vereenvoudigen en efficiënter maken. Om de endochondrale aanpak voor botweefselengineering succesvol toe te passen in de klinische praktijk, moeten nog een aantal hordes genomen worden. Allereerst is het van belang om in het achterhoofd te houden dat alle resultaten m.b.t. de endochondrale methode zijn voortgekomen uit celkweek en dierproeven. Deze resultaten zijn dus niet direct te vertalen naar menselijke situaties. In onze studies bijvoorbeeld, werden MSCs gebruikt verkregen uit het beenmerg van ratten. Er is geen twijfel over dat voor de uiteindelijk klinische toepassing de effectiviteit van het gebruik van menselijke cellen opnieuw zal moeten worden opgehelderd. Een ander probleem dat moet worden aangepakt is de selectie van het diermodel. Tot dusver hebben we alleen de in vivo botvorming op een ectopische plaats van immunogecompromitteerde dieren getest. Dit model werkt door het aantal variabelen betrokken bij botvorming aanzienlijk te verminderen, waardoor er gemakkelijker er een simpelere vertaling is van de overgang van kraakbeen substraat naar bot . Daarom moeten de uitkomsten getoetst moeten worden in een klinisch relevante situatie, dwz een orthotopische locatie in immunocompetente dieren, waarbij de mechanische belasting, biochemische/ontstekingsfactoren en stamcellen nauw betrokken zijn. Hoewel onze studies hebben bevestigd dat de structuur van de scaffolds het patroon van botvorming kunnen beïnvloeden, is het moeilijk een uitspraak te doen over het precieze mechanisme daarvan. De chemische en fysische eigenschappen van de gebruikte scaffolds verschilden daarvoor teveel. Om dit beter te kunnen onderzoeken is het nodig om een serie materialen te ontwikkelen waarbij telkens slechts een parameter (fysisch, chemisch, enz) gevarieerd wordt. Daarnaast is het belangrijk om grotere scaffolds te testen, omdat de invasie van cellen en 166 Samenvatting, Slotopmerkingen en Toekomstperspectieven vascularisatie een grotere uitdaging zal zijn. Wanneer de interactie tussen de cellen en scaffold materialen is ontrafeld, is het mogelijk om op efficiëntere en grotere schaal endochondrale botregeneratie toe te passen mede door het gebruikt van afgestemde materialen. Een hoge celdichtheid is cruciaal is voor de vorming van de kraakbeen substraten, dit kan de klinische toepassing van de endochondrale benadering met autologe cellen uitdagend maken. De beschikbaarheid van deze cellen kan dan een beperkende factor worden, zeker wanneer het om een groot botdefect gaat. Op dit moment is het onderliggende mechanisme van het door de geïmplanteerde cellen induceerde reactie op de gastheercellen en de opeenvolgende weefselregeneratie nog grotendeels onbekend. Verder onderzoek is nodig om de cellulaire en moleculaire gebeurtenissen die betrokken zijn bij dit proces op te helderen, voordat een betrouwbare endochondrale therapie kan worden ontwikkeld. Belangrijker nog, is het begrijpen van de communicatie tussen de geïmplanteerde cellen en de gastheercellen, alleen wanneer dit tot in detail begrepen is kan endochondrale botweefselengineering verder worden vereenvoudigd en efficiënter worden toegepast. Samengevat, botherstel via de endochondrale route vereist begrip van vele factoren, waaronder een groot aantal verschillende cellen, matrices en signalen. Hoewel er nog grote inspanningen verricht moeten worden, houdt endochondrale botweefselengineering een grote belofte in voor klinische toepassing. Het uiteindelijke doel van deze aanpak is niet alleen het kopiëren van het natuurlijke ossificatieproces, maar ook het recapituleren van de essentie van dit proces op een simpele en kosten efficiënte manier. Dit kan bereikt worden door scaffold materialen die op het juiste moment de juiste biologische signalen afgeeft om endochondrale botvorming te verkrijgen zonder implantatie van cellen. Om dit te kunnen realiseren is een grondig begrip nodig van de onderliggende cellulaire en moleculaire mechanismen van de endochondrale route, met inbegrip van cel-cel en cel-matrix/materiaal interacties. Dit vereist de samenwerking van vele disciplines, met inbegrip van cel- en moleculaire biologie, materiaalkunde, farmacologie, chemie en geneeskunde. 9 167 Acknowledgements List of publications Curriculum Vitae Acknowledgements Life is like a box of chocolates. You never know what you are going to get. Before 2010, I could not imagine such a long journey in a country so far away from my family and friends. But now, to me, Netherlands has become not only the witness for my growth during the PhD, but also a sweet home where I discovered and harvested so much beauty and love. While writing the last chapter of my thesis, I cannot help bringing back a lot of nice and meaningful memories I gathered in the past four years. The help, guidance, trust, support and love from the people I have stayed with, certainly made this journey so unique, colorful, and exciting. Although it is not possible to mention everyone by name, I would like to specifically acknowledge the following people. First and foremost, I would like to express my sincere gratitude to my supervisor, Prof. dr. John Jansen. Dear John, thank you for giving me the opportunity to experience such a wonderful journey in your department. You opened the door of tissue engineering realm for me and I have learned so much from you in the past years, not only from your broad knowledge and solid expertise in biomaterial field, but also from your way of critical thinking and highly efficient working attitude. Besides the research, the experience of working with you in the clinic has also being very enjoyable and enlightening for me. Your enthusiasm for work, skillful and elegant clinical operations, as well as the great considerations and respect to the patients has set a superb model for me. All of these will absolutely encourage me and constitute a valuable asset in my future career. I am also deeply grateful to my co-supervisor, Dr. Fang Yang. Dear Fang, I am so lucky to have you around in the past few years. To me, you are not only a supervisor who gave solid supports and trust for me and my experiments, but also a friend, a sister, who are always willing to help me, and share the joyful or difficult moments in my life. Your expertise in biomaterial and critical attitude has helped me make a lot of progress in research and writing. Your warm-hearted character and optimistic attitude towards work and life always encouraged and inspired me. I would never adapt to the Dutch life so quickly and pleasantly without your guidance on how to understand and appreciate the difference between western and eastern culture. In the past 4 years, I could not count how many joyful moments we have shared together: Chinese dinners, table games, BBQ, Karaoke, sports, travelling,... I sincerely thank you, and also Ronald, for the warm company during my PhD. This friendship will never fade in the future. I would like to further express my gratitude to my other co-supervisor, Dr. Sanne Both. Dear Sanne, you lead me to the field of cell culture from the very beginning, and it turned out that I enjoyed a lot working with those tiny and energetic creatures 170 Acknowledgements during my PhD. Your solid knowledge in cell culture and broad experiences in many biological techniques provided helpful guidance for my experiments. I appreciate your great patience in teaching; your vivid explanations always made me easily get the points. You showed a good example for me, not only being precise and collaborative on work, but also being confident, open-minded and adventurous in life. It has been very pleasant to be your colleague and friend. Thank you for all of your supports, encouragements and trust during my PhD; and also thank you for taking me to all those wonderful trips: paragliding, Christmas night in Gouda, and so on. I would also like to extend my appreciation to Prof. Yining Wang, Prof. Gerjo van Osch and Prof. Pieter Buma. Dear Prof. Wang, I was lucky to be your student during my master study, and thank you for the continuous encouragements in the last few years. The knowledge and working attitude I learned from you have firmly supported me pursuing the goal of PhD. Dear Prof. van Osch, I would never finish my studies so smoothly without the help from you and also the technical support from your lab in Rotterdam. Thank you for all those valuable and efficient advices for my studies and articles. Dear Prof. Buma, it was my pleasure to have you as my mentor. Thank you for the inspiring discussions about my studies. My appreciation also goes to my colleagues Dr. Joop Wolke, Dr. Frank Walboomers, Dr. Jeroen van den Beucken, Dr. Sander Leeuwenburgh, and former colleague Dr. Marco Lopez-Heredia. Your sharp insight and solid expertise in biomaterials field strongly supported my research. Thank you for your advice in all kinds of discussions and group meetings. Dear Joop, I admire your critical attitude to research and the warm heart whenever I or anyone else needs your help or advice. Dear Frank, it has been always a pleasure to talk with you. I was inspired not only by your humorous stories, but also your positive attitude towards life and great passion for the family. Dear Jeroen, I was always impressed by your high working efficiency. Thank you for all the laughers you brought to us and also for the nice company during the trip in Greece. Dear Sander, I learned from you to be generous and modest as a person, precise and efficient in work, passionate and adventurous for life. Thank you for every encouragement after my presentations. Dear Marco, I would never forget the pleasant time working with you during my first trip in Nijmegen as an exchange student. You were such a patient teacher leading a novice like me approaching science, and also a good friend sharing all kinds of stories in life. Your earnestness and diligence towards work has certainly set an excellent example for me and encouraged me to start my PhD journey. My work would not be completed so smoothly without the help from all “superhero” technicians. I would like to express my gratitude to Natasja, Vincent, 171 Monique, Martijn and Rene for your patience and being always there helping us. Special thank goes to Natasja, not only for your great support in my work, but also the pleasant trips we went together. I will always remember the beautiful stainings we worked out together, and the joyful swimming lessons. I would also like to thank Wendy Koevoet and Nicole Kops from Erasmus MC, for the technical supports and advices for my experiments. My acknowledgement also goes to all the people from CDL who helped me arrange and perform animal studies. In particular, I would like to thank Debby, Jeroen and Daphne for your friendly assistance. Many people in dental school have given me all kinds of help during my PhD. Dear Kim, Henriëtte, Monique and Vera, I appreciate your professional and patient support in the past years. Dear Rachel, thank you for the friendly guide in the clinic. I also want to acknowledge Dr. Ruud Kasten, Dr. Aukje Bouwman, and Dr. Ewald Bronkhorst. Dear Ruud, I would never have the chance to start my journey in Nijmegen without your invitation in 2009. Thank you for all your warm encouragements in the past years. Dear Aukje and Ewald, thank you for your valuable advices and help for my experiments. My stay in Netherlands would have never been so pleasant and enjoyable without my wonderful colleagues in department of Biomaterials. Thank you all for creating such a warm and friendly environment. Thank you all for sharing with me the knowledge in experiments and every story in life. Because of you all, working was never just working; I enjoyed every cheerful and joyful day with my friends here. My first heartfelt appreciation goes to Xiangzhen and Na. Dear Xiangzhen, I can still vividly recall the very first day we flew thousands kilometers away from home and landed together in this new place. In the past 4 years, we took care of each other as colleagues, best friends and sisters. You showed me great character of persistence, honesty, courage and strong mind. I cannot count how many times I turned to you whenever I met happy, depressed, confusing or just bored moments in my life. Thank you for always being there listening and supporting me. Although we may not see each other often after the PhD, I will always cherish the fantastic trips we have made, the beautiful handcraft shops we have stopped, and the childish dreams we have talked. Dear Na, from bachelor to PhD, our friendship grows tighter and tighter. You are the witness for every step I made forward, and also such a loyal friend that always supports me with your best efforts. Thank you and also Zengping for hosting so many nice activities and delicious dinners. Thank you for every encouragement and wish you made for my work and life. Your homelike company made me never feel lonely even far away from my own family. 172 Acknowledgements My PhD journey could never be so cheerful without my lovely (ex)officemates from all different countries: Eva, Alexey, Kemal, Daniel, Kambiz, Mani, Rosa, Rania and Floor. Dear Eva, it has been such a pleasant journey for me to have you around from the beginning of my PhD. Your passion and sincerity to life, friends and family, always inspired and moved me. I will never forget our fantastic adventures in US, and so many other moments we have shared together; no matter happy or sad, we grow stronger and more confident together towards our life and future. Dear Alexey, you are such an easy-going and helpful person. I was lucky to have you around. Your optimism and generous help always calmed me down when I met difficulties. Your humors and stories made me never feel bored staying in the office. I will always remember the countless laughers we have had, the bets we have made, and also the cookies you have stolen from me. :) Thanks also go to Thordis for volleyball, BBQ and many other nice activities together. Dear Kemal and Daniel, from 2nd to 1st floor, we were always united in the same office. I like the office time that we were sharing those serious or non-serious knowledge; and I also keep a lot of happy memories from our hanging-out time after work. Thank you, Kemal, for all kinds of warm invitations and for letting me “beat” you in pingpang. Dear Kambiz and Mani, you created such a relaxing and “niiicee” atmosphere for me during the last period of my PhD. Your humors always released me from the stress of writing. Thank you for the friendly advices whenever I turned to you. Dear Rosa, Rania and Floor, thank you for all the warm encouragements and nice chats. I always miss our old cozy office and all the joyful moments filled there. I never feel alone also because I have a big Chinese family in the department supporting me at different periods of my PhD. Dear Huanan, thank you for your company playing all kinds of sports with me, searching dinners with me in the canteen after work, and having those relaxing chats. You set a good example for Chinese guys in our department: diligent in work, optimistic in life, passionate for family, but maybe not your cooking skills. :) Dear Jinling, you were my greatest helper from the beginning of my PhD and I learned a lot from your optimism and persistence. I can still vividly recall the days we explored the experiments together and encouraged each other, which made me feel more confident and supported. I also cherish all the good time we shared in the office, in the sports center and in many other corners of our life. Dear Wei and Jie, you were always so helpful whenever I discussed the experiment or the daily life with you. I also miss those girls’ talks and many joyful activities we spent together. Dear Jiankang, Xinjie, Yang and Yue, staying with you guys makes me always feel younger and more energetic. I will miss a lot the delicious meals you have cooked. 173 It would not be a complete memory for the days in Biomaterials, if I missed the names of many friendly colleagues that have been or still are in this big family: Dear Manuela and Ruggero, you started my exploration of this city; it was much fun to hang out with you. Thank you for all those warm encouragement and advice for my work and life. Dear Paula and Astghik, it has been always enjoyable to talk with you, and I learned a lot from your cooking lessons. Thanks also to dear colleagues Ljupcho, Matilde, Edwin, Hamdan, Cristina, Jan Willem, Bart, Daniël, Claire, Antonio, Pedro, Simone, Reza, Ana, Michele, Take, Paulo, Jing, Weihua, Hongbin, and “neighbor” colleagues Aysel and Niels. Thank you for the company along the journey. Besides work, I also shared a lot of cheerful moments with my housemates. Dear Ineke, I learned a lot from you to discover the simple beauty from life. Thank you for taking care of me when I was sick, and the nice talks during our dinners. Dear Lie and Yi, it was such a warm feeling when we built a home together. I will always remember the days we spend together in Vossenlaan. The company and support from friends outside the department also made my PhD journey more colorful. Thank you, Weiwei, Xuan, Shuo, Junru, Gu Yan, Sun Yan, Yaping, Shanshan, Xuehui, Xinhui, Jiabo, Siji, Haiteng, Shuiquan, Junxiao, Jinge, Daixiao, Bronte, Tina, Andrew, Steven, Johnny, Esther, Richard, Bobby, Steffie, and many others. Although the time we spent together might not be so long, but the nice memory will always stay. I could also not forget the warm heart and support from my dear friends in China: Sun Wei, Xiling, Shiqian, Zhejun, Huiting, Yuanyuan, and Manli. Although we are separated from such a long distance and heading for different life directions, our friendship will never change. I am grateful to the Nijhuis members. Dear Marianne, Henk, Carolien, Jelle and Marius, thank you for supporting me and taking me as your family. This always warms me. To my beloved parents and grandparents, I miss all of you! Your unconditional and endless love, trust and support is the most precious fortune I could have in my life. 亲爱的老爸老妈,感谢你们三十年来不辞辛苦的养育我,教育我。你们无 私的爱和信任为我撑起了一片温暖的天空,让我勇敢的飞的更高更远去寻 找自己的梦想;你们无微的呵护和牵念是一棵繁茂的大树,时时刻刻守护 着我为我遮风避雨。女儿一直漂泊在外,很愧疚未能陪在你们身边,唯有 在心里默默的感激你们对我的宽容和体谅。也许永远也比不上你们为我付 出的三十年,但是女儿希望能用后面的三十年,五十年,无论多长,报答 你们,也成为你们的树,让你们依靠。还有我亲爱的外爹外奶,忘不了你 们为我忙碌的身影,还有你们对我一次次的叮咛和挂念,祝愿你们一直健 康平安。 174 Acknowledgements I owe my last words to you, Arnold, my dearest colleague, friend, and love. To me, the biggest achievement of the PhD is not about the completion of this thesis; having met you and building a life together with you was the most meaningful and beautiful gift I got. Thank you for catching me into your world, for spoiling me with your powerful love, for supporting me unlimitedly in my work and life. I enjoyed every single second I spent with you along our journey, and I know this journey will never end. 175 List of publications Related to this thesis: 1. Yang, W.*, Bongio, M.*, Yang, F., Both, S.K., Leeuwenburgh, S.C.G., van den Beucken, J.J.J.P., and Jansen, J.A., Combining Osteochondral Stem Cells and Biodegradable Hydrogels for Bone Regeneration. (Stem Cells and Bone Tissue, 2013. CRC Press: p. 326-348.) 2. Yang, W., Yang, F., Wang, Y., Both, S.K., and Jansen, J.A., In vivo bone generation via the endochondral pathway on three-dimensional electrospun fibers. (Acta Biomaterialia, 2013. 9(1): p. 4505-12.) 3. Yang, W., Both, S.K., van Osch, G.J.V.M., Wang, Y., Jansen, J.A., and Yang, F., Performance of different three-dimensional scaffolds for in vivo endochondral bone generation. (European Cells & Materials, 2014. 27: p. 350-364.) 4. Yang, W., Both, S.K., Zuo, Y., Tahmasebi Birgani, Z., Habibovic, P., Li, Y., Jansen, J.A., and Yang,F., Biological evaluation of porous aliphatic polyurethane/ hydroxyapatite composite scaffolds for bone tissue engineering. (Journal of Biomedical Materials Research Part A, submitted.) 5. Yang, W., Both, S.K., van Osch, G.J.V.M., Wang, Y., Jansen, J.A., and Yang, F., Effects of in vitro chondrogenic priming time of bone marrow derived mesenchymal stem cells on in vivo endochondral bone formation. (Acta Biomaterialia, submitted) 6. Yang, W., Both, S.K., Jansen, J.A., and Yang, F., Roles of cartilage templates in mediating mesenchymal stem cell behavior in endochondral bone tissue engineering (manuscript in preparation). Other publications: 1. Yang, W.*, Yan, X.Z.*, Yang, F., Kersten-Niessen, M., Jansen, J.A., and Both, S.K., Effects of Continuous Passaging on Mineralization of MC3T3-E1 Cells with Improved Osteogenic Culture Protocol. (Tissue Engineering Part C Methods, 2014. 20(3): p. 198-204.) 2. Sariibrahimoglu, K.*, Yang, W.*, Leeuwenburgh, S.C.G., Yang, F., Wolke, J.G., Zuo, Y., Li,Y., Jansen, J.A., Development of porous polyurethane/strontiumsubstituted hydroxyapatite composites for bone regeneration. (Journal of Biomedical Materials Research Part A, 2014. accepted) 3. Lopez-Heredia, M.A., Yang, W.*, Sariibrahimoglu, K.*, Bohner, M., Yamashita, D., Kunstar, A., van Apeldoorn, A.A., Bronkhorst, E.M., Felix Lanao, R.P., 176 List of publications Leeuwenburgh, S.C.G., Itatani, K., Yang, F., Salmon, P., Wolke, J.G., and Jansen, J.A., Influence of the pore generator on the evolution of the mechanical properties and the porosity and interconnectivity of a calcium phosphate cement. (Acta Biomaterialia, 2012. 8(1): p. 404-14.) 4. Both, S.K., Yang, W., Yang, F., and Jansen, J.A., The influence of 3D scaffolds in endochondral bone formation. (Journal of Tissue Engineering and Regenerative Medicine, 2012. 6: p. 33.) 5. Yeatts, A.B., Both, S.K., Yang, W., Alghamdi, H.S., Yang, F., Fisher, J.P., and Jansen, J.A., In vivo bone regeneration using tubular perfusion system bioreactor cultured nanofibrous scaffolds. (Tissue Engineering Part A, 2014. 20(1-2): p. 139-46.) 6. Cai, X., Yang, F., Yan, X.Z., Yang, W., Yu, N., Oortgiesen, D.A.W., Wang, Y., Jansen, J.A., and Walboomers, X.F., Influence of bone marrow-derived mesenchymal stem cells differentiation in periodontal regeneration: an in vivo study. (Journal of Clinical Periodontology, submitted) * These authors contributed equally. 177 Curriculum Vitae Wanxun Yang (杨婉洵) was born on June 15, 1985 in Bengbu (China). In September 2003, she was enrolled in Wuhan University, majoring in Dentistry with sevenyear schooling system. She received basic medicine education, resident training as a general dentist, and postgraduate training as a prosthodontist in the Department of Prosthodontics under the supervision of Prof. Yining Wang. In 2009, she performed a twomonth internship in the Department of Biomaterials at Radboud University as an exchange student. In June 2010, she was conferred the Master Degree of Dental Medicine with Outstanding Graduate Award of Wuhan University. The same year, she obtained the national medical license as a dental practitioner. In October 2010, she was supported by Chinese Scholarship Council (CSC) and started her PhD study in the Department of Biomaterials at Radboud University Medical Center, under the supervision of Prof. John A. Jansen. Her project mainly focused on combining biomaterials and pluripotent stem cells for bone regeneration through the endochondral ossification pathway, among which two studies were in collaboration with Department of Orthopedics, Erasmus Medical Center Rotterdam. The results of her PhD studies are described in this thesis and published as separated research articles in scientific journals. Part of the results were also presented in several international conferences, including International Bone-Tissue-Engineering Congress (Hannover, 2011), International Society of Differentiation Conference (Amsterdam, 2012), Tissue Engineering and Regenerative Medicine International Society (TERMIS) EU Congress (Istanbul, 2013), and the 5th International Conference on Tissue Engineering (Kos, 2014) where she also received the Aegean Conference Travel Award in recognition of her performance in the rapid-fire session. 178 “Success is not final, failure is not fatal: it is the courage to continue that counts.” - Winston Churchill 179
© Copyright 2025 Paperzz