9

Applications of 3D Bioprinting

Abstract

Three-dimensional bioprinting has been a powerful tool in patterning and precisely placing biologics including living cells, nucleic acids, drug particles, proteins, and growth factors to recapitulate tissue biology. Since the first time of cytoscribing cells demonstrated by Klebe in 1986, bioprinting has made a substantial leap forward, particularly in the last 10 years, and been widely used in fabrication of living tissues for various application areas. The technology has been recently commercialized by a number of emerging businesses, and bioprinters and bioprinted tissues have gained significant interest in medicine and pharmaceutics. This chapter presents the application areas of bioprinting technology including tissue engineering and regenerative medicine, transplantation and clinics, drug testing and high-throughput screening, and cancer research.

Keywords

Cancer research; Drug testing; High-throughput screening; Tissue engineering and regenerative medicine; Transplantation

There are no such things as applied sciences, only applications of science

Louis Pasteur

9.1. Introduction

Bioprinting is a growing field that makes a revolutionary impact on medical and pharmaceutical sciences, and it has gained significant attention worldwide. It is a computer-aided transfer process for simultaneous writing of living cells and biomaterials with a prescribed layer-by-layer stacking organization to fabricate bioengineered constructs (Ozbolat, 2015b). It offers great precision on spatial placement of cells, proteins, DNA, drug particles, grow factors, and biologically active particles to better guide tissue generation and formation. This powerful technology appears to be more promising for advancing tissue fabrication toward physiologically relevant tissue constructs, tissue models, tissues and organs, and organs-on-a-chip models for a broad spectrum of application areas (Ozbolat and Hospodiuk, 2016).
Bioprinting technology has a broad utility in various application areas such as tissue engineering and regenerative medicine (Jakab et al., 2010; Moroni et al., 2006), transplantation and clinics (Ozbolat, 2015a), drug screening and high-throughput assays (Snyder et al., 2011), and cancer research (Perkins, 2007) as depicted in Fig. 9.1. Bioprinting for tissue engineering and regenerative medicine fields has been around for more than a decade, and anatomically correct cell-laden constructs and scaffolds have been fabricated for various tissue types from connective and epithelial tissues to muscle and nervous tissues. With its great advantage in patterning and precisely positioning multiple cell types, bioprinting has circumvented one of the major shortcomings of traditional scaffold fabrication techniques and has enabled fabrication of nativelike tissues with heterocellular microenvironment. Although the vast majority of the efforts have been geared toward the fundamental science behind major bioprinting techniques such as extrusion-based bioprinting (EBB) (see Chapter 4), droplet-based bioprinting (DBB) (see Chapter 5), and laser-based bioprinting (LBB) (see Chapter 6), a substantial focus has recently been given to bioprinting for functional tissue fabrication (Ozbolat, 2015a). Particularly, considerable work has been dedicated to bioprinting for animal transplantation, where bioprinted tissues have been implanted to various associated sites in vivo. With the latest advances in in situ bioprinting, bioprinting technology has become a highly attractive approach to build body parts in operating rooms. As further progress taking place in biomaterials, cells, and transplantation technologies, bioprinting will translate from bench to bedside when approved for human use and has a myriad of advantageous in operating rooms in the near future. Before transitioning into clinical practice, bioprinting has already made a great leap in pharmaceutical use as it does not entail any regulatory approvals and there is currently an emerging bioprinting market for tissue fabrication for drug testing and high-throughput assays. With the inclusion of multiple cell types and facilitated complex heterocellular physiologically relevant environment, bioprinted tissue models (i.e., liver) have been used in drug screening. In addition, bioprinting has recently been used in cancer research to investigate cancer pathology, growth, and metastasis in a physiologically relevant microenvironment (Knowlton et al., 2016).
image
Figure 9.1 Application areas of bioprinting technology including tissue engineering and regenerative medicine, transplantation and clinics, drug screening and high-throughput assays, and cancer research (Image courtesy of Christopher Barnatt, www.explaningthefuture.com).
In this chapter, application areas of bioprinting technology are presented with an in-depth discussion on successfully bioprinted tissue types in tissue engineering and regenerative medicine, transplantation and clinics, drug screening and high-throughput assays, and cancer research. For each application, limitations of existing technologies are discussed, and future prospects are provided to the reader.

9.2. Tissue Engineering and Regenerative Medicine

Bioprinting of functional organs at clinically relevant dimensions still remains elusive as there are several challenges such as but not limited to the integration of vascular network from arteries and veins down to capillaries, incorporation of various cell types to recapitulate complex organ biology, and limited structural and mechanical integrity and long-term functionality (Ozbolat and Yu, 2013). Despite these difficulties, a wide variety of tissues have been successfully bioprinted such as thin or hollow tissues [i.e., blood vessel (Itoh et al., 2015)] or tissues that do not need vascularization, i.e., cartilage (Yu et al., 2016).
image
Figure 9.2 Bioprinted tissue constructs for tissue engineering and regenerative medicine. (A) Inkjet bioprinting of human mesenchymal stem cells (hMSCs) in hydrogels for bone tissue engineering, where bioprinting resulted in uniform distribution of hMSCs in oppose to accumulated hMSCs at the bottom of the scaffold due to gravity when hMSCs were manually pipetted (Reproduced/adapted with permission from Gao et al. (2014)). (B1) Bioprinting of cardiac tissue constructs with connected ventricles using a modified-HP printer. (B2) SEM image of the cross section of the scaffold showing loaded cells (Reproduced/adapted with permission from (Xu et al. (2009a,b)). (C1) Bioprinting of a 4 mm-PEG cartilage tissue construct with (C2–C5) Safranin-O staining with limited glycosaminoglycan deposition without transforming growth factor beta-1 and fibroblast growth factor-2 treatment even if high density (20 million cells per mL) was used (Reproduced/adapted with permission from Cui et al. (2012b)). (D1) A bioprinted heart valve with encapsulated human aortic valvular interstitial cell and (D2) a representative immunohistochemistry image showing αSMA (green) and vimentin (red) expression (Reproduced/adapted with permission from (Duan et al. (2014)). (E1) A bioprinted 40-layer alginate ring-shape construct, where (E2) human embryonic stem cells differentiated toward hepatocyte-like cells showing positive for albumin expression (Reproduced/adapted with permission from Faulkner-Jones et al. (2015)). (F1) Bioprinted four-layer lung tissue model with highly organized distribution of a 549 cells (green) and endothelial cells (labeled with VE-cadherin in pink), where F-actin and nuclei were labeled in red and white, respectively. (F2) Histological cross section stained with Masson–Goldner trichrome coloration showing highly uniform thickness of a tissue sample. Cytoplasm, collagen fibers, and cell nuclei were stained in red, green, and dark brown, respectively. (F3) Sagittal cross section demonstrates uniform epithelial layer on the top and endothelial cell layer at the bottom (Reproduced/adapted with permission from Horváth et al. (2015)). (G1) Scaffold-free bioprinting of nerve tissue using bone marrow stem cell (BMSC) pellet and coculture of BMSC and Schwan cell pellet, where cell pellet was bioprinted within printed agarose mold for aggregation of cells (G2) for fabrication of multiluminal nerve grafts. (G3–G4) Bielschowsky's staining of histological sections showing axons in black dots (Reproduced/adapted with permission from Owens et al. (2013)). (H) Green fluorescent protein–transduced islets plotted in alginate/gelatin scaffolds kept their morphology over a week culture compared to the encapsulated counterparts in bulk hydrogels (control) (Reproduced/adapted with permission from Marchioli et al. (2015)). (I1) A bioprinted skin substitute consisting of 20-layer of fibroblasts and 20-layer of keratinocytes on the top of Matriderm® was implanted on dorsa of a mice (I2) resulted in complete wound closure on Day 11 (Reproduced/adapted from open-access source Michael et al. (2013)). (J1) Bioprinted vascular tissue constructs using a coaxial nozzle apparatus, where (J2) loaded human umbilical vein smooth muscle cells (HUVSMCs) generated smooth muscle matrix under 6-week differentiation period, particularly in the luminal and outer surface of the tissue constructs (Reproduced/adapted from Zhang et al. (2015)).

9.2.1. Bone Tissue

Bone tissue engineering has been widely studied using bioprinting as bioprinting has the ability to fabricate anatomically correct patient-specific tissue constructs. In a recent study (Gao et al., 2014), a thermal inkjet bioprinter was used to fabricate poly(ethylene glycol) dimethacrylate (PEGDMA) scaffolds. Bone marrow–derived human mesenchymal stem cells (hMSCs) were coprinted with nanoparticles of bioactive glass and hydroxyapatite (HA) under simultaneous polymerization. Bioprinting in that study enabled uniform distribution of hMSCs compared to manually pipetted hMSCs, which accumulated at the bottom of the scaffold due to gravity (see Fig. 9.2A). The results revealed that the bioprinted constructs encapsulating hMSCs and HA demonstrated the highest cell viability, collagen production, and alkaline phosphate activity with increased compressive modulus after 21-day culture in vitro. Bioprinting HA particles were also performed for in situ bioprinting purposes, where an LBB system was used to deposit HA nanoparticles (n-HAs) into mouse calvarial defects in a framework study (Keriquel et al., 2010), as detailed in Section 9.3.
In another study, Fedorovich et al. (2008) bioprinted heterocellular tissue constructs made of Matrigel™ and alginate hydrogels. Endothelial progenitor and multipotent stromal cells were bioprinted in a spatially controlled manner, and the bioprinted constructs were subcutaneously implanted into immune-deficient mice. By incorporating osteoinductive biphasic calcium phosphate microparticles, multipotent stromal cells were differentiated into osteogenic lineage and facilitated bone formation in 6 weeks. In addition to osteoinductive materials, incorporation of growth factors is also crucial in stem cell differentiation in bone tissue engineering. Phillippi et al. (2008) demonstrated the effect of bone morphogenetic protein (BMP-2) on stem cell fate. Using inkjet bioprinting of patterned BMP-2 on fibrin-coated coverslips, primary muscle-derived stem cells were differentiated toward osteogenic lineage on the pattern even if they were treated with myogenic differentiation conditions.

9.2.2. Cardiac Tissue

Cardiac tissue engineering has been a growing interest as heart failure is a devastating disease (Hirt et al., 2014). While myocardium tissue has a limited regeneration capability as myocytes proliferation rapidly ceases after birth (Eulalio et al., 2012), tissue engineering of such a structurally and functionally complicated organ is essential. In the literature, limited attempt has been made in bioprinting of cardiac tissue models. Jakab et al. (2008) demonstrated EBB of tissue spheroids comprising human vascular endothelial cells (HUVECs) and cardiac cells isolated from myocardial tubes of chicken embryos. Tissue spheroids, which were adhesive and scaffold-free and which possessed rapid self-assembly capabilities, were bioprinted next-to-each other on collagen type-I biopaper in a single-layer grid pattern. Upon bioprinting, tissue spheroids were fused together in approximately 70 h and formed a single cardiac tissue patch that can synchronously beat. In addition to the scaffold-free approach undertaken in the abovementioned work, scaffold-based bioprinting has been investigated in a few studies. Xu et al. (2009a,b) bioprinted cardiac tissue constructs in a half-heart shape with connected ventricles using inkjet-based bioprinting as shown in Fig. 9.2B1–B2. In their study, primary feline adult and H1 cardiomyocytes were encapsulated in alginate/gelation composite hydrogels, and the cross-linker (calcium chloride solution) was selectively sprayed layer by layer. The resulting tissue construct with connected ventricles was electrically stimulated, and functional excitation–contraction coupling was successfully demonstrated. In addition to these studies, patterning of cells was also applied in cardiac tissue engineering. Gaebel et al. (2011) utilized a laser-induced forward transfer (LIFT) technique to pattern HUVECs and hMSCs in a geometrically defined pattern on polyester urethane urea (PEUU), and the fabricated samples were transplanted to the infarcted zone of rat hearts after LAD ligation. In 8 weeks posttransplantation, samples with LIFT-derived patterns facilitated increased vessel formation compared to randomly bioprinted cells as control groups, and the resulted myocardium patch provided significant functional improvement. Besides primary cells, human cardiac–derived cardiomyocyte progenitor cells (hCMPCs) were also bioprinted in mesh pattern made of alginate hydrogel (Gaetani et al., 2012). Bioprinted hCMPCs demonstrated phenotypic properties of cardiac lineage with enhanced expression of early cardiac transcription factors Nkx2.5, Gata-4, and Mef-2c.

9.2.3. Cartilage Tissue

Current tissue engineering techniques for cartilage regeneration cannot produce cartilage tissue that is indistinguishable from native tissue in terms of zonal properties and architectures (Makris et al., 2015). Thanks to its great potential for precise spatial and temporal deposition of cells and biomaterials with sophisticated patterns, bioprinting has recently gained increasing attention for engineering cartilage tissues that can closely mimic native tissues with zonally differentiated cells and extracellular matrix (ECM) composition (Kesti et al., 2015). Due to absence of blood vessels, cartilage tissue bioprinting has been extensively studied using various bioprinting modalities. LBB of stem cell–differentiated chondrocytes was attempted by Gruene et al. (2010) in which a computer-aided biofabrication technique was used with the assistance of LIFT. They successfully bioprinted porcine bone marrow–derived mesenchymal stem cells (MSCs) with high viability, where cells maintained their functionality and differentiation ability into osteogenic and chondrogenic lineage.
Inkjet-based bioprinting has also been used in cartilage tissue engineering as well as for cartilage defect repair. Cui et al. (2012a) modified an HP desktop printer into a bioprinter, where they were able to bioprint human chondrocytes loaded in PEGDMA hydrogel in a layer-by-layer manner. The bioprinted cartilage construct had mechanical properties and biochemical composition close to native cartilage. Also by implanting bioprinted cartilage constructs into articular cartilage defects, integration with the native tissue was observed with enhanced interface strength, which significantly improved the quality of the repaired cartilage tissue. In another study using the above experimental setup, the same group fabricated PEG scaffolds (see Fig. 9.2C1) and investigated the effect of combined transforming growth factor beta-1 (TGF-β1) and fibroblast growth factor-2 (FGF-2) on cell proliferation and differentiation capability, and demonstrated that samples treated with TGF-β1 and FGF-2 facilitated the highest glycosaminoglycan (GAG) content (Cui et al., 2012b) and samples without growth factor treatment did not secrete GAG even in 4-week culture (see Fig. 9.2C2–C5). Most recently, Xu et al. (2013) created a hybrid bioprinting method to fabricate mechanically improved cartilage tissue constructs by combining three-dimensional (3D) bioprinting and electrospinning techniques. In that study, electrospinning of polycaprolactone (PCL) fibers together with inkjet bioprinting of rabbit elastic chondrocytes in fibrin–collagen hydrogel was demonstrated. After printing, cell viability was well maintained and fabricated constructs formed cartilage tissues both in vitro and in vivo. Furthermore, the printed structures showed improved mechanical properties compared to printed hydrogels alone.
In addition to above hydrogels, sodium alginate has been widely used for cartilage tissue bioprinting. The author's group demonstrated hybrid bioprinting of chondrocytes loaded in bioprinted filaments in tandem with bioprinting of chondrocyte spheroids to increase the cell density of the tissue constructs (Ozbolat et al., 2014). A Multi-Arm BioPrinter was used to facilitate such a complex hybrid architecture. Using sodium alginate and silver nanoparticles, McAlpine's group (Mannoor et al., 2013) successfully printed bionic ear models, which were composed of chondrocyte-loaded alginate in ear shape and a conductive coil with the ability to translate sound waves into digital data. In addition to bioprinting cartilage tissue constructs, there have been studies to improve the bioink capabilities as well. In this regard, Bonassar's group presented a mixing chamber to increase the homogeneity of crosslinked sodium alginate loaded with chondrocytes (Cohen et al., 2010). The presented study revealed that as mixing of cell-loaded precrosslinked alginate increased, the fidelity and mechanical properties of the bioprinted constructs improved in addition to the enhancement in their cell viability. Recently, Markstedt et al. (2015) demonstrated mixing alginate with nanocellulose, which has outstanding shear thinning properties enabling fabrication of anatomically correct ear and meniscus constructs. Apart from hydrogels mainly used as bioink, Cho and his coworkers (Pati et al., 2014) recently used Taylor's organ decellularization approach (Ott et al., 2008) to remove cells from native cartilage and chopped it into smaller fragments leaving gel-like material behind. The material was then loaded with chondrocytes and bioprinted in tandem with PCL supporting frame to generate cartilage tissue constructs. The new bioink demonstrated highly suitable environment for growth and proliferation of loaded chondrocytes.
Despite the great progress in bioprinting for cartilage tissue regeneration, bioprinting of zonally stratified articular cartilage tissues with different structural, biomechanical, and biological properties is still a challenge and further progress is needed to achieve articular cartilage tissue constructs with zonal differentiation including more horizontal and thinner collagen fibers with high cell density in superficial zone, and relatively vertical and thicker collagen fibers with less cell density in deeper zones.

9.2.4. Heart Valve

In addition to cardiac tissue engineering, engineering heart valves is also important as heart valves do not possess regeneration capability and dysfunctional heart valves, if the damage or disease is detrimental, need to be replaced by mechanical or biological prosthetic counterparts (Hockaday et al., 2014). Such replacement valves, however, are limited by thrombogenicity and calcification (Jana and Lerman, 2015). Despite its critical role in cardiovascular system, only a very limited number of work has been demonstrated in bioprinting of heart valves. Butcher's group at Cornell University demonstrated the bioprinting of a heart valve for the first time (Hockaday et al., 2012) using a dual-head bioprinter modified from a Fab@Home printer (Malone and Lipson, 2007). In that work, a dual crosslinking mechanism, consisting of ionic and physical crosslinking, was used to print polyethylene glycol diacrylate (PEGDA) mixed with sodium alginate. After printing, porcine aortic valve interstitial cells were seeded and cultured for up to 21 days. In their study, anatomically accurate axisymmetric aortic valve geometries, composed of a root wall and trileaflets, were demonstrated. Although the first-time presented work does not fall under bioprinting as cells were not involved during the printing process, the group later demonstrated bioprinting of aortic valves using different hydrogels and cell phenotypes. In another study by that group (Duan et al., 2013), dual-nozzle bioprinting of composite alginate/gelatin hydrogel was performed to fabricate thin hydrogel discs. Aortic root sinus smooth muscle cells (SMCs) and aortic valve interstitial cells (VICs) were spatially bioprinted, and samples were incubated for a week. The results revealed a cell viability of 81.4 ± 3.4% and 83.2 ± 4.0% SMCs and VICs, respectively. Acellular aortic valve constructs, on the other hand, exhibited reduced modulus, ultimate strength, and peak strain. In a recent study (Duan et al., 2014), the same group presented bioprinting of composite hydrogels using methacrylated hyaluronic acid (Me-HA) and methacrylated gelatin (Me-Gel) loaded with human aortic valvular interstitial cells (HAVICs). Fabricated samples (see Fig. 9.2D1) represented the designed trileaflet valve shape accurately, and the cells bioprinted with increased Me-Gel concentration exhibited better spreading. HAVICs encapsulated within the composite hydrogel expressed alpha smooth muscle actin (α-SMC) and vimentin (see Fig. 9.2D2) and remodeled ECM with deposition of collagen and GAGs.

9.2.5. Liver Tissue

Although liver has excellent regenerative and recuperative properties, liver tissue engineering has been a growing interest while liver failure, in association with failure of multiple organs, is a significant cause of morbidity and mortality (Yoon No et al., 2015). Engineering liver tissues stands as a promising direction for future organ transplantation needs in addition to the great potential of bioprinted liver tissue models in drug testing and high-throughput screening as liver tissue is highly sensitive to drug toxicity. Faulkner-Jones et al. (2015) demonstrated the bioprinting of human-induced pluripotent stem cells (hiPSCs), where bioprinted hiPSCs were stimulated and differentiated into hepatocytes for liver microorgan engineering. The presented work systematically analyzed the effect of the bioprinting process and parameters on stem cell fate and the influence of pressure and nozzle length on the viability of hiPSC and human embryonic stem cells, concluded that the utilized inkjet bioprinting process was gentle enough to maintain viability and pluripotency of cells, and directed their differentiation into hepatic lineage. A dual-head valve-based inkjet bioprinter was used to deposit sodium alginate and calcium chloride to fabricate multilayer tissue constructs (see Fig. 9.2E1), and the results demonstrated that bioprinted stem cells differentiated into hepatic lineages successfully after a 17-day differentiation period and expressed hepatocyte markers including HNF4α, albumin, and ZO-1, where albumin secretion peaked on day 21 (see Fig. 9.2E2).
In addition to bioprinting liver tissue constructs, liver carcinoma HepG2 immortal cells were bioprinted in larger tissue models. Bertassoni et al. utilized a modified NovoGen MMX Bioprinter™ and bioprinted HepG2 cells, and fibroblast within gelatin-methacrylamide (GelMA) hydrogel strands along with agarose strands. Upon printing, agarose solidified immediately due to rapid drop in the temperature, and UV light was applied to photo-crosslink GelMA precursor under 6.9 mW/cm2 of UV light (360–480 nm) up to a minute (Bertassoni et al., 2014a). After complete gelation of GelMA, agarose was completely removed to create perfusable channels. The study revealed that cells preserved their viability up to 8 days. The similar work was then extended using other hydrogels such as star poly(ethylene glycol-co-lactide) acrylate (SPELA), PEGDMA, and PEGDA hydrogels at different concentrations (Bertassoni et al., 2014b).
This section presented liver tissue bioprinting attempts for tissue engineering only, and further discussion on bioprinting of liver tissue models for drug testing and high-throughput screening will be provided to the reader in Section 9.4.

9.2.6. Lung Tissue

Bioprinting for lung tissue engineering is highly new as there is only a recent work attempted to fabricate a lung tissue model using bioprinting. Horváth et al. (2015) demonstrated bioprinting of an in vitro air–blood barrier model using BioFactory® by regenHU Ltd. In this regard, they bioprinted a zonally stratified tissue construct layer by layer. First, a thin layer of Matrigel™ was bioprinted as a basement membrane followed by bioprinting of a single layer of EA.hy926 endothelial cells to facilitate attachment of cells on the Matrigel™ layer. Later, on day 2, a new layer of Matrigel™ was bioprinted on the top of the previously built construct followed by bioprinting of a single layer of A549 epithelial cells. Manually deposited layers were also constructed as control samples. On day 5, the samples were fixed for characterization, and cell viability of >95% and ≥86% were achieved for epithelial and endothelial cells, respectively. As can be seen in Fig. 9.2F1 and F3, epithelial and endothelial cells were uniformly distributed with epithelial cells on the top and endothelial cells at the bottom. Sagittal histological sections also confirmed formation of a highly thin, packed, and uniform tissue layer (see Fig. 9.2F2) compared to manually pipetted control samples. The barrier quality, such as tightness of the constructs, was investigated through measuring the translocation of blue Dextran molecules from the apical to the basolateral compartment of the samples after 3 days of culture, and results revealed that the tightness of the bioprinted samples was better than that of the manually pipetted samples.

9.2.7. Neural Tissue

Engineering tissues for nervous system offers tremendous promise to replace diseased, aged, or injured components of nervous system; however, there is limited work done in the context of bioprinting nerve grafts. Lee et al. (2010) studied the effect of vascular endothelial growth factor (VEGF) release on proliferation and migration of murine neural stem cells (C17.2). In their study, C17.2 cells were bioprinted on a collagen layer next to a fibrin disk loaded with VEGF. The study showed that cells migrated toward VEGF-releasing fibrin gel and proliferated successfully in contrast to the cells that could not proliferate in collagen matrix. Recently, Hsieh et al. (2015) demonstrated bioprinting of a thermo-responsive polyurethane (PU) hydrogel with tunable stiffness and gelation ability at 37°C without need for a crosslinker. They showed the effectiveness of the bioink by loading it with neural stem cells and injecting it into a zebrafish embryo neural injury model. The results revealed that the injected gel rescued the function of impaired nervous system in 6 days.
Owens et al. (2013) used a scaffold-free approach, where pellet of Schwann cells and bone marrow stem cells were extruded within a 3D-printed agarose mold (see Fig. 9.2G1 for the schematic of the process). Cells in agarose mold aggregated and formed nerve tissue graft with three lumina in each as shown in Fig. 9.2G2. The fabricated grafts were then implanted into mice, and their histology (see Fig. 9.2G3–G4) and functionality were evaluated 10 months after implantation and compared with the functionality of autologous grafts and hollow collagen grafts. Although the number of samples was not enough to draw a definitive conclusion about the performance of the 3D-bioprinted grafts with respect to commercially available collagen grafts, the presented case demonstrated a proof-of-concept for bioprinting nerve grafts.

9.2.8. Pancreas Tissue

As primary pancreatic β-cells do not easily survive in vitro and only a very few attempts were taken place in differentiation of β-cells from human stem cells, regeneration of pancreas tissues is primarily embodied to the extent that only β-cells from mouse lines or insulinoma cells have been used to fabricate pancreatic islets (Pagliuca and Melton, 2013; Pagliuca et al., 2015; Raikwar et al., 2015). A very few work has been done in the context of bioprinting for pancreatic tissues. Recently, Marchioli et al. (2015) encapsulated human and mouse islets as well as rat insulinoma INS1E β-cells within alginate or alginate/gelatin hydrogels and bioprinted them in dual-layer scaffolds. The scaffolds were later implanted in diabetic mice and explanted 7 days thereafter. Although the viability and morphology of islets were not impaired by encapsulation and bioprinting processes in both alginate and alginate/gelatin hydrogels (see Fig. 9.2H), bioprinted islets and INS1E β-cells lost their functionality in 7 day as they were not responsive to the change in glucose level. This can be attributed to the high level of presence of calcium (Ca2+) ions within crosslinked alginate because transmembrane calcium ion gradient, mediated by voltage-gated calcium channels, stimulated higher insulin secretion at low glucose level (Proks and Ashcroft, 1995). In addition, a recent work demonstrated the microfabrication of scaffold-free tissue strands (with strong expression of insulin) for EBB (Akkouch et al., 2015), where tissue strands were made of rat fibroblasts and mouse insulinoma TC-3 β-cells in the core and shell, respectively. The authors envisioned to use the demonstrated tissue strands for scale-up tissue bioprinting purposes.

9.2.9. Skin Tissue

Several tissue engineering approaches have been applied in skin tissue fabrication and tissue substitutes including autologous split-thickness skin graft (gold standard) (Coruh and Yontar, 2012), allografts (Leon-Villapalos et al., 2010), acellular dermal substitutes, and cellularized graft-like commercial products (Leon-Villapalos et al., 2010), i.e., Dermagraft and Apligraf (Organogenesis Inc.) (Metcalfe and Ferguson, 2007). Recently, bioprinting technology has been adopted for skin tissue fabrication as well. Lee et al. (2013) presented bioprinting of skin tissue using an 8-channel valve-based bioprinter, where a 13-layer tissue construct was bioprinted layer by layer using collagen hydrogel. Keratinocytes were bioprinted on the top of alternating layers of human foreskin fibroblasts and acellular collagen layers, and the resulting constructs demonstrated densely packed cells in epidermis layers in oppose to the dermis with low density of cells and less ECM deposition. In addition to DBB, LBB has also been used for biofabrication of skin tissue substitutes (Koch et al., 2012). Cells from human immortalized keratinocyte cell line and NIH 3T3 fibroblasts were bioprinted in collagen matrix in alternating layers on a sheet of Matriderm™. Histological results demonstrated high density of keratinocytes and fibroblasts with expression of laminin protein, which is a major component of basement membrane in skin. The same group extended their work (Michael et al., 2013) and demonstrated implantation of the tissue constructs on dorsa of mice as shown in Fig. 9.2I1. Results revealed that the bioprinted tissues were engrafted with the hosts in 11 days (see Fig. 9.2I2) with stratified epidermis with early sign of differentiation and formation of stratum corneum as well as some blood vessels. Control samples at the air–liquid interface in vitro culture, on the other hand, demonstrated proliferation of cells with limited differentiation. In a recent study (Yanez et al., 2014), Boland's group demonstrated the effect of bioprinting endothelial cells within skin substitutes on formation of macrovasculature during new tissue remodeling. In this regard, they encapsulated neonatal human dermal fibroblast and neonatal human epidermal keratinocytes (NHEKs) in collagen and laid down the dermis layer followed by patterning human dermal microvascular endothelial cells (HMVECs) on dermis layer of the construct by selectively bioprinting thrombin-laden HMVECS on manually deposited fibrinogen layer. The process was completed by covering the fibrin layer with collagen-laden NHEK cells. Then, the fabricated skin substitutes were implanted on dorsa of mice and compared with implanted commercially available skin substitutes (control). The results revealed that bioprinted HMVECs formed microvessels and the bioprinted constructs barely generated contraction compared to the control groups. In abovementioned studies, tissue constructs were bioprinted in vitro and implanted to a host; however, Skardal et al. (2012) demonstrated in situ bioprinting of stratified skin substitutes by alternating layer of fibrinogen/collagen and thrombin loaded with amniotic fluid–derived stem cells due to their lack of immunogenicity. With in situ bioprinting, skin substitutes were 3D bioprinted directly onto full thickness wounds on pigs and recapitulated the native skin more closely than control groups including the bioink loaded with MSCs and acellular hydrogels. Despite the efforts in skin tissue bioprinting, biofabrication of skin substitutes that virtually mimic native skin is still a challenge as integrating sweat glands is remained elusive.

9.2.10. Vascular Tissue

Bioprinting of scale-up tissues and organs vitally depends on vascularization as integration of vascular network will essentially provide oxygen and media supply to cells for their survival and function (Ozbolat, 2015b). Vascular tissue fabrication has been performed by various bioprinting modalities including EBB (Zhang et al., 2013a,b; Yu et al., 2013; Norotte et al., 2009), DBB (Xu et al., 2012; Christensen et al., 2015; Blaeser et al., 2013), and LBB (Xiong et al., 2015b). In EBB, a wide variety of extrusion techniques have been utilized. The author's group used coaxial-nozzle extrusion, where hydrogels including sodium alginate and chitosan were bioprinted directly in tubular form with encapsulated cells (Zhang et al., 2013a,b). During coaxial (core–shell) flow, ejected crosslinker (flowing through the core) contacted the precursor hydrogel solution (flowing through the shell) and facilitated rapid gelation and formation of tubular constructs (Fig. 9.2J1). Six-week-cultured HUVSMC-laden samples demonstrated deposition of smooth muscle matrix (Fig. 9.2J2). That approach enabled direct bioprinting of vascular constructs in a practical manner. In addition, there are other direct vascular tissue bioprinting approaches, such as bioprinting droplets of cell-laden hydrogels layer by layer using inkjet-based bioprinting performed by Nakamura et al. (Nishiyama et al., 2008), Huang's group (Xu et al., 2012; Christensen et al., 2015), and Blaeser et al. (2013). With the ability of bottom-up construction, inkjet-based bioprinting enabled branched tubes built in both horizontal and vertical directions. A similar approach was also performed using LBB demonstrated by Huang's group (Xiong et al., 2015a,b). In abovementioned studies, scaffold-based approaches were utilized; however, Forgacs and his coworkers followed a scaffold-free approach in bioprinting vascular tissues, where tissue spheroids were bioprinted one by one and self-assembled into larger tissue units (Norotte et al., 2009). As agarose is inert to cell adhesion, the printed agarose mold facilitated rapid fusion of tissue spheroids and maturation of the tissue.
In addition to direct bioprinting of tubular vascular tissues, indirect bioprinting of perfusable tissue constructs has been performed using various hydrogels including fibrin (Lee et al., 2014), collagen (Zhao et al., 2012), and GelMA (Bertassoni et al., 2014a). In indirect approach, a fugitive ink [that is dissolvable or reversibly crosslinkable such as agarose (Bertassoni et al., 2014b), sugar (Miller et al., 2012), Pluronic (Wu et al., 2011), and gelatin (Zhao et al., 2012)] was used to create open channels. Upon removing the fugitive ink, endothelial cells were perfused and glued to create endothelium within open channels (Miller et al., 2012). This approach enables bioprinting of highly complex vascular constructs that can be perfused over long time depending on the degradation profile of the matrix. Although both approaches can be utilized, the former one is more appropriate in generating vascular grafts for transplantation and the second one is more appropriate for fabrication of perfusable channels for in vitro tissue engineering applications (Ozbolat, 2015a). Bioprinted vascular tissues should be designed and fabricated in a way that they can be easily sutured to a blood vessel in a host, and possess certain properties such as enough mechanical strength to satisfy suture retention and burst pressure, sufficient intactness of endothelium to prevent thrombosis, and a high patency rate to support occlusion-free circulation (Quint et al., 2011).

9.2.11. Composite Tissues

In addition to single tissue types, efforts have been geared toward bioprinting composite tissues to recapitulate the complex biology, anatomy, and functionality of organ-level structures. Merceron et al. (2015) recently demonstrated fabrication of muscle–tendon units by bioprinting hybrid constructs using a multihead nozzle assembly. In this regard, PCL and PU was 3D printed to construct a frame to support cellular constructs, where half of the unit was printed using PCL and the other half was printed using PU (Fig. 9.3A1–A4). A composite hydrogel-based bioink, comprising 3 mg/ml hyaluronic acid, 35 mg/ml gelatin, and 25 mg/ml fibrinogen in calcium-free high-glucose Dulbecco's Modified Eagle Medium (DMEM), was used to 3D bioprint 3T3 fibroblasts and myoblasts into the PCL and PU frames to construct tendon and muscle units, respectively. The results revealed cell viability of >80% with differentiated cells at the end of a 7-day culture. The final muscle–tendon units were elastic on the PU-C1C12 muscle section with an elastic modulus of 0.39 ± 0.05 MPa and stiff on the PCL-3T3 tendon side with an elastic modulus of 46.47 ± 2.67 MPa.
In addition to muscle–tendon units, bioprinting of osteochondral models has been an interest in tissue engineering. Fedovorich et al. (2011) demonstrated bioprinting of MSCs and chondrocytes using alginate in mesh pattern with two different cell types bioprinted within two opposite ends of the scaffold. MSCs were coextruded with osteoinductive biphasic calcium phosphate particles, HA, and β-tricalcium phosphate. The bioprinted samples were cultured with a mixture of chondrogenic and osteogenic medium up to 21 days. The results revealed that bioprinted osteochondral tissue constructs demonstrated differentiated characteristics of cells into both osteogenic and chondrogenic lineages along with related ECM deposition in vitro and in vivo. A similar approach was performed by Park et al. (2014), where a systematic analysis was performed to understand the effect of native ECM components on the fate of osteoblasts and chondrocytes. In this regard, they bioprinted osteoblasts in collagen type-I and chondrocytes in hyaluronic acid and compare the performance of bioprinted osteoblasts and chondrocytes when they were bioprinted in hyaluronic acid and collagen type-I, respectively. Fourteen-day culture in vitro suggested that osteochondral tissue regeneration could be successfully attained when proper hydrogel type was selected. The same group demonstrated another osteochondral model by depositing alginate hydrogel loaded with human chondrocytes and human MG63 osteoblasts within a PCL frame (Shim et al., 2012). Osteoblasts and chondrocytes were supplemented with osteogenic and chondrogenic growth factors loaded in hydrogels for differentiation purposes. Another approach was conducted using acoustic-based bioprinting for the same purpose, where bioprinted nanodroplets of MSCs in TGF-β1 and BMP-2 patterns (see Fig. 9.3B1–B4) resulted in localized differentiation of MSCs toward osteogenic and chondrogenic lineages shown by the gene expression study (Gurkan et al., 2014).
image
Figure 9.3 Bioprinted composite tissue constructs. (A1) A muscle–tendon unit frame made of polyurethane (PU) (upper-half) and polycaprolactone (PCL) (lower-half) at a higher magnification view of (A2) printed PU filaments, (A2) filaments at the interface, and (A3) printed PCL filaments (Reproduced/adapted with permission from Merceron et al. (2015)). (B1) Bioprinted fibrocartilage samples (B2–B3) patterned by selectively depositing human mesenchymal stem cell (hMSC)–laden methacrylated gelatin droplets loaded with either transforming growth factor beta-1 or bone morphogenetic protein (BMP-2). (B4) Bioprinted precursor gel droplets were photo-crosslinked, and droplets were demonstrated using Rhodamine B (red) and Dextran-Alexa Fluor 488 (green) (Reproduced/adapted from open source Gurkan et al. (2014)).
In addition to osteochondral models, Yu and Ozbolat (2014) demonstrated hybrid bioprinting of macrovascularized stromal tissue, where scaffold-free tissue strands made of fibroblasts were assembled around a perfusable macrovasculature loaded with SMCs extruded through a coaxial nozzle unit (Yu et al., 2014). Tissue strands quickly fused and assembled around the macrovasculature in a week, which can be further scaled-up by extending the macrovascular network.

9.2.12. Other Tissue Types

In addition to the presented tissue types, there is a few other work at the early fundamental study level for bioprinting of retinal and brain tissues. Lorber et al. (2014) presented piezoelectric inkjet bioprinting of retinal ganglion cells (RGCs) and glia and investigated the effect of bioprinting parameters on the viability of cells and their growth-promoting properties. They concluded that inkjet bioprinting did not adversely affect the cell viability and RGC neurite outgrowth, rather RGCs demonstrated further neurite growth when bioprinted on a glial substrate. Recently, Lozano et al. (2015) presented manual deposition of primary cortical neuron-laden gellan gum-RGD for brainlike tissue fabrication. Three-layer constructed tissue models, with cortical neurons encapsulated in the top and bottom layers, demonstrated axon growth and penetration toward the cell-free middle layer in 5 days. Although no computer-control motion system was applied, the presented work unveiled the first-time demonstration of layer-by-layer fabrication for brain tissue engineering.

9.3. Transplantation and Clinics

Bioprinting of living tissue and organ constructs has been widely studied, and performance of these constructs was assessed via animal transplantation. Several bioprinted tissue types, including but not limited nerve (Owens et al., 2013), cardiac (Gaebel et al., 2011), blood vessel (Itoh et al., 2015), bone (Keriquel et al., 2010), and skin (Yanez et al., 2014), have been implanted into associated locations on animals to evaluate their functionality, neovascularization, and anastomosis and engraftment with the host (Ozbolat, 2015a). In addition, various tissue constructs have been bioprinted and implanted subcutaneously to assess in vivo differentiation of cells and functionality of implanted tissue constructs (Fedorovich et al., 2011). Despite these attempts, none of bioprinted tissues has been clinically used for humans as no approval has been granted from Food and Drug Administration (FDA) yet. There are no regulations laid down for bioprinters or bioprinted products; however, with the increasing global interest and emerging businesses in the growing bioprinting market, the success with the first technology going through FDA regulations will be exemplary for preceding technologies and products. For details of the regulatory concerns of bioprinting, the reader is referred to Chapter 10.
While bioprinting technology is still in its infancy in clinics, 3D-printed plastic, ceramic, or metallic implants for bone tissue replacement (Bose et al., 2013) have been successfully transplanted into humans. In addition to permanent implants, a recent work published in the New England Journal of Medicine (Zopf et al., 2013) demonstrated a unique case of transplantation of a 3D-printed bioresorbable airway splint into an infant. The institutional review board of the University of Michigan consulted with FDA and approved the use of 3D-printed device under the emergency-use exemption and the written consent of the patient's parents. No unforeseen problems have been observed with the splint, and full degradation of the device is expected to take around 3 years. This was an exemplary case for clinical use of 3D-printed scaffolds and hopefully will open up similar success with the bioprinted tissues and organs. Despite the accomplishments in bioprinting research, bioprinting for transplantation in a clinical setting for humans requires further advances and translational efforts (Ozbolat, 2015a). Organs and tissues that do not need significant vascularization (i.e., skin and cartilage) are expected to be translated into clinical use sooner. Tissues and organs that are metabolically highly active (i.e., heart, pancreas, and liver) are immensely challenging. No bioprinting technology so far facilitated fabrication of a vascular hierarchical network spanning arteries and veins down to capillaries. Since it is difficult to bioprint capillaries at the submicron scale using the current technology, an alternative could be to bioprint macrovasculature and then leave the nature to create capillaries by itself. The other approach in facilitating vascularization is in situ bioprinting of tissue and organ constructs directly into the defect sites in surgery settings rather than bioprinting tissue constructs and maturating and assessing them in vitro before transplantation. With in situ bioprinting, bioprinted tissue constructs can recruit endothelial cells from the host and facilitate neovascularization followed by anastomosis of newly formed vessels with the vascular network of the host. Therefore, in situ bioprinting has a great advantage over traditional two-step bioprinting approach and can be applied for regeneration of a wide array of tissues and organs (i.e., maxillo- and craniofacial reconstruction, plastic surgery, skin tissue, and flap tissue). In situ bioprinting has been recently used in skin regeneration for large wounds on pig models (Skardal et al., 2012) and calvarium defects in rodents (Ozbolat, 2015a; Keriquel et al., 2010). Fig. 9.4A demonstrates laser-assisted printing of n-HA particles into critical size defects on a mouse model, where one of the defects left empty for the control group. The histology and microtomography results for bone regeneration were not consistent (see Fig. 9.4B–C), which was due to immobilization of printed n-HA particles within the defects. Although the ink in the demonstrated work did not comprise any biologics, it unveiled the application of LBB systems in operating rooms, which will enable translation of bioprinting technologies from bench to bedside.
image
Figure 9.4 (A) Laser-assisted bioprinting setup for in situ bone printing of hydroxyapatite nanoparticles into critical-size calvarium defects in mice model. (B) Histology results of three-month in vivo culture revealing new tissue formation (marked with star) in oppose to an empty defect. (C1) X-ray microtomography images of bone tissue formation (C1) in 1 week, (C2) 1 month, and (C3) 3 months (Reproduced/adapted with permission from Michael et al. (2013)).
One of the controversies of the clinical translation potential of bioprinted tissues and organs originates from the bioprinting technology itself as bioprinting involves living cells in bioink and the use of patient-specific cells is fairly new in bioprinting. Stem cells, such as embryonic stem cells and induced-pluripotent stem cells (Yoshida and Yamanaka, 2010), have been potential unlimited sources of patient-specific cells for fabrication of tissues and organs. Patient-specific cells can be differentiated, then bioprinted or bioprinted, and then differentiated toward multiple lineages to fabricate tissues and organs that will have minimum immunogenicity risk as discussed in Chapter 8.

9.4. Drug Screening and High-Throughput Assays

Drug discovery entails a time-consuming and costly endeavor that requires substantial investment in financial and human resources. Despite continuing efforts to improve the productivity of drug development process, only 1 out of an estimated 10,000 new chemical entities and 1 out of 10 drug candidates entering the clinical trial reach the final approval stage and enter the market (Nam et al., 2015). Improving the ability to predict the efficacy and toxicity of drug candidates earlier in drug discovery process will speed up the translation of new drugs into clinics. Recent attempts in 3D in vitro assay systems is an ideal way to resolve this bottleneck since 3D-printed tissue models can closely mimic the native tissue and have the capability to be used in high-throughput assays as they can be bioprinted in microarrays. Bioprinted tissue and organ models have been increasingly considered for the potential of pharmaceutics use such as drug toxicology and high-throughput screening (Peng et al., 2016). Among the three different modalities discussed in previous chapters, DBB has been most common for pharmaceutical use due to its simplicity, versatility, and high-throughput capability (Gudapati et al., 2016). Table 9.1 shows the major strengths and limitations of each modality, within the application domain of pharmaceutics.
In the literature, liver and tumor tissues have been a primary focus in fabrication of tissue models for pharmaceutics. Sun's group investigated the fabrication of a bioprinted liver microorgan model for drug metabolism (Chang et al., 2008) (see Fig. 9.5A1). In that study, an automated syringe-based direct cell writing process was applied for extrusion of hepatocyte (HepG2) cell–encapsulated alginate strands as demonstrated in Fig. 9.5A2. Polydimethylsiloxane (PDMS) elastomer soft lithography was combined with a micromolding technique to fabricate 3D microfluidic chambers housing aforementioned constructs. The presented 3D liver microorgan with a sinusoidal flow pattern was an in vitro 3D microfluidic, microanalytical, microorgan (3DM) device for simulation of the physiological liver response to drug administrations and toxic chemical exposure. Effective drug metabolism in microliver chamber was demonstrated by metabolizing a nonfluorescent prodrug, 7-ethoxy-4-trifluoromethyl coumarin, to an effluent fluorescent metabolite 7-hydroxy-4-trifluoromethyl coumarin. Additionally, they schemed and created dual-microtissue microfluidic chips, which were connected to facilitate multicellular interaction and downstream effects of metabolism on the target tissue (Chang et al., 2010; Snyder et al., 2011). Epithelial cells and hepatocytes encapsulated in Matrigel™ were used to show the path of drug diffusing from blood stream to the tissue as epithelial cells are the cells lining along the lumen through which the drugs pass from blood stream to the target, where hepatocytes were used as the target cell type. The antiradiation drug amifostine was used as a prodrug, which can be converted to an active form by epithelial cells. The results showed that the percentage of radiation-damaged cells for the single tissue was more than twice the dual tissue. Based on the presented tissue model, researchers developed a computational macroscale model for such in vitro tissue models using a convection–diffusion–cell kinetics numerical framework, which is helpful for future research in 3D microorgan pharmacokinetics and toxicity (Tourlomousis and Chang, 2015).

Table 9.1

Bioprinting Modalities and Their Performance Comparison in Pharmaceutical Applications

BackgroundStrengthsLimitationsApplications in PharmaceuticsReferences
Extrusion-based Bioprinting (EBB)

• Introduced in early 2000s

• The most common and affordable bioprinting modality

• Driven by pneumatic or mechanical forces

• Print materials in the form of filaments

• Compatible with a wide range of bioink properties

• Compatibility with viscosities in a wide range (30 mPa/s to >6 × 107 mPa/s)

• Enables bioprinting of scaffold-free bioink such as tissue spheroids, which is not currently feasible using other modalities

• Facilitates vascularization using direct or indirect (with fugitive ink) bioprinting

• Suitable to extrude three-dimensional tissue constructs or organ-on-a-chip for drug testing and toxicity analysis

• Commercially available with moderate cost

• Substantial cell damage due to shear stress of highly viscous fluids, small nozzle diameter and high dispensing pressure

• Not practical for high-throughput bioprinting of tissue models

• Limited bioprinting resolution preventing direct fabrication of microcapillary network

• Limited control on cell–cell and cell–matrix interactions

• Liver-on-a-chip on a polydimethylsiloxane (PDMS) bioreactor for testing hepatic toxicity of acetaminophen

• Valve- and pneumatic-based extrusion of liver microorgan on a PDMS chamber for assaying drug metabolic properties

• Extrusion of breast cancer neotissues in a multiwell plate to test antitumor drugs

Chang et al. (2010); King et al. (2014); Bhise et al. (2016); Snyder et al. (2011)
Droplet-based Bioprinting (DBB)

• First introduced in early 2000s

• Inkjet printers are the most commonly used type of DBB

• Driven by thermal, piezoelectric, or acoustic forces

• Print materials in the form of liquid droplets

• Compatibility with small viscosities in the range of 3.5–12 mPa/s

• High speed (1–10,000 droplets/s), high resolution (1–300 picoliter in volume)

• Compatibility with many biological materials including living cells, DNA, RNA, biochemicals

• Suitable to drop cell populations on microarrays or organ-on-a-chip for HTS

• Affordable, versatile, and commercially available

• No uniformity in droplet size

• Inconstancy in encapsulating a single cell in each droplet on microarrays for high throughput screening (HTS)

• Nozzle clogging in high cell densities and fibrous bioink solutions

• Cross-contamination when bioprinting of multiple bioink solutions takes place simultaneously

• Thermal inkjet bioprinting Escherichia coli–laden alginate for high-throughput antibiotics screening

• Piezoelectric jetting of Sac6-EGFP yeast cells as microarrays for analysis of drug dose–response of latrunculin A

Rodríguez-Dévora et al. (2012); Saunders and Derby, (2014)
Table Continued

image

BackgroundStrengthsLimitationsApplications in PharmaceuticsReferences
Laser-based Bioprinting (LBB)

• First introduced in 1999

• Less popular than DBB or EBB

• Consists of a pulsed laser beam with a focusing system, a donor slide including two layers (energy-absorbing layer and biological material layer), and a collector substrate

• Stereolithography and its modifications also enable bioprinting of cells

• Driven by laser generated shock waves

• Compatibility with viscosities in range of 1–300 mPa/s

• Nozzle-free

• Generating negligibly cell damage

• Facilitates deposition of cells in the densities of 108 cells/ml with a resolution of one cell per droplet

• High-resolution feature of stereolithography and its modifications enables integration of vascular channels within tissue constructs

• Labor intensive and time-consuming preparation

• Difficulty of accurately targeting and depositing cells

• High cost and no commercial availability

• Not practical to bioprint heterocellular models

• LBB has not been applied to pharmaceutical use yet

Peng et al., (2016)

image

image
Figure 9.5 Bioprinted tissue models for drug testing and high-throughput screening: (A1) bioprinted liver tissue model loaded into a perfusion chamber (Reproduced/adapted with permission from Chang et al. (2010)), (A2) where HepG2-laden alginate filaments were patterned allowing media flow for drug testing (Reproduced/adapted with permission from Tourlomousis and Chang (2015)). (B1) A schematic showing inkjet-bioprinting of three-layer constructs, where agar and bacteria were bioprinted on a glass slide followed by printing of alginate and blend of antibiotics and CaCI2, (B2) light microscopy, and (B3) fluorescence imaging of the bioprinted samples (Reproduced/adapted with permission from Ref Rodríguez-Dévora et al. (2012)). (C1) A 3 × 3 mm bioprinted liver tissue model for drug testing with (C2) hematoxylin and eosin stain showing parenchymal (P) and nonparenchymal (N) regions (Image courtesy of Organovo Holdings, Inc.).
Emerging microengineering technologies enable versatile fabrication of 3D cell-based microarrays including soft lithography, surface patterning, microfluidic-based manipulation, and bioprinting (Feng et al., 2011). Among them, bioprinting technology has numerous advantages, including high precision control over size, microarchitecture and cellular composition, high-throughput capability, coculture and vascularization ability, and low risk of cross-contamination, where multiple tissue types need to be located separately with minimum cross-migration of cells (see Table 9.2 for the comparison) (Peng et al., 2016). A recent study developed a novel inkjet-based bioprinting method for assembling a high-throughput miniature drug-screening platform as presented in Fig. 9.5B1 (Rodríguez-Dévora et al., 2012). The authors applied a modified Hewlett Packard model 5360 compact disk printer with picoliter per droplet resolution and bioprinted Escherichia coli–laden alginate to array a chip on coverslips (see Fig. 9.5B2–B3). Droplets of three antibiotics were printed on the spots of cells in a layer-by-layer fashion. Results demonstrated similar cell viability, functionality, and antibacterial effects of antibiotics in both inkjet-bioprinted and micropipetted samples, which confirmed that inkjet-bioprinted high-throughput array is an effective method to minimize the typical drug screening test. Demirci's group presented cell-based biosensors (CBBs), where acoustic-based bioprinting, studied by the same group earlier for high-throughput screening purposes with various cells (including mouse embryonic stem cells, fibroblasts, AML-12 hepatocytes, human Raji cells, and HL-1 cardiomyocytes) (Demirci and Montesano, 2007), was used to create microarray of SMC droplets in collagen. The bioprinted microarray was then stimulated with different environmental conditions (e.g., temperature), and the effect of the applied stimuli as well as the bioprinting process parameters on cell viability was evaluated. With the bioprinted CBBs, the effects of analytes, such as pathogens, contaminants, toxins, and drug candidates, on living systems can be obtained (Xu et al., 2009a,b). Besides 3D bioprinting of high-throughput cell microarrays, controlled delivery of drug candidates into cell microarrays is also a key factor for successful drug screening. Various methods have been developed for controlled drug delivery onto cell microarrays, including drug patterning, drug stamping, aerosol sprays, and microfluidic drug loading (Feng et al., 2011).

Table 9.2

Comparison of Bioprinting With Other Three-Dimensional In Vitro Technologies

MethodsHanging Drop MethodMicrowell-Based MethodMicrofluidicsMagnetic Force–Based PatterningBioprinting
MechanismsCellular spheroids are formed by gravitational forceMicrowells are fabricated by nonadhesive materials to forming cellular spheroidsMicroflow mediates stacking cells in layers or forming cell spheroids using trappingMagnetically labeled cells are compacted in spheroids form under magnetic forcesCells are deposited in scaffold-based or scaffold-free manner
Size uniformity++++++++++++++
Microarchitectural controllability++++++++++++
Scalability+++++++++
Coculture ability++++++++++
High-throughput capability+++++++++++++
Low risk of cross-contamination+++++++++

image

+++, high;++, medium;+, low.

Recently, bioprinting companies have developed bioprinted tissue models for high-throughput drug screening (Vaidya, 2015; Nelson, 2015). In November 2014, Organovo began offering its 3D-printed “exVive3D” liver tissue models to screen drugs for liver toxicity. A scaffold-free tissue bioprinting approach was performed using NovoGen MMX Bioprinter™, where pellet of cocultured human hepatocytes, hepatic stellate, and endothelial cells were bioprinted into a temporary mold structure with building units in hexagonal shape (see Fig. 9.5C1) (Roskos et al., 2015). After bioprinting, cells further aggregated and the tissue construct matured toward a nativelike tissue model (see Fig. 9.5C2), which maintains normal function for at least 42 days. Formation of microcapillaries took place at certain locations. The bioprinted tissues were characterized by ATP production and secretion of liver-specific albumin protein. The results revealed that ATP production increased over time during 4-week culture period, and the enzyme-linked immunosorbent assay (ELISA) results showed that albumin secretion increased after fabrication and stabilized in 21 days. Organovo's system demonstrated very clear signals confirming the commercial drug's safety and the failed drug's toxicity. Other companies, including Aspect Biosystems from Vancouver, Canada, and Texas-based Nano3D Biosciences, are also developing technology for the similar purpose.

9.5. Cancer Research

Two-dimensional (2D) tumor models have been widely used in cancer research; however, they do not represent the physiologically relevant environment of cancer tissues as they lack cell–cell and cell–matrix interactions in 3D. Thus, bioprinting has offered great advantages to recapitulate cancer microenvironment to precisely locate various cell types and microcapillaries to study cancer pathogenesis and metastasis; however, bioprinting for cancer research is new, and only a few research work has been performed in this emerging application area (Knowlton et al., 2015).
Demirci's group was the first to demonstrate bioprinting of tumor tissue models for in vitro assays (Xu et al., 2011). In their study, human ovarian cancer (OVCAR-5) cells and MRC-5 fibroblasts were bioprinted using an inkjet-based bioprinting platform with dual ejectors. Multiple cell types were spontaneously bioprinted on Matrigel™ to form multicellular acini in a high-throughput and reproducible manner with a spatially mediated microenvironment with controlled cell density and cell–cell distance. The presented approach did not only demonstrate a tool for cancer research but also provided a great platform for high-throughput screening. Sun's group recently demonstrated bioprinting of HeLa cells to form cervical tumor models (Zhao et al., 2014). In this regard, HeLa cells were extruded and bioprinted in a gelatin/alginate/fibrinogen composite hydrogel in patterned form with >90% cell viability. In 5–8 days, HeLa cells migrated toward each other and formed cell aggregates within hydrogel filaments (see Fig. 9.6) in oppose to the control groups, where cells in 2D culture formed cell sheets with lower chemoresistance and lower level expression of metalloproteinase. Although these two studies presented biofabrication of tumor spheroids using bioprinting technology, biofabrication of a larger tissue model to study cancer cell migration and metastasis is also important. Huang et al. (2014) demonstrated a laser-based 3D projection printing system to bioprint HeLa cells and noncancerous 10 T1/2 fibroblasts in PEGDA along with microvascular network with channel widths of 25, 45, and 120 μm to reflect blood vessel diameters. The results revealed that bioprinted fibroblasts were not affected by the morphology of the channel width; however, HeLa cells migrated significantly when the channel diameter decreased. In addition to scaffold-based approaches, Organovo company demonstrated scaffold-free bioprinting of breast cancer model using NovoGen Bioprinting™ platform, where cancer cells were surrounded by a physiologically relevant stromal milieu comprising of MSC-differentiated adipose cells, mammary fibroblasts, and endothelial cells (see Fig. 9.6B1) (King et al., 2014). Histomorphological analysis revealed that bioprinted neotissues were viable for 2 weeks in vitro with a clear compartmentalization of adipose, stromal, and epithelial components with localized signs of microcapillary formation. The effects of chemotherapeutic drug tamoxifen were assessed for viability by ATP luciferase assay and concluded that isolated 2D cancer cells were more susceptible to tamoxifen-induced toxicity than cells incorporated into 3D tissue models when treated with same dose of the drug for the same period in vitro. Despite the recent attempts in engineering cancer microenvironment using bioprinting technology, further advancements are needed for fabrication of philologically relevant complete microenvironment, comprising tumor site, healthy site, and microvascular network in between, to study cancer metastasis research.
image
Figure 9.6 Bioprinted cervical tumor model. (A) Phase-contrast images showing bioprinted HeLa cells forming spheroids within gelatin/alginate/fibrinogen on day 5, where spheroids got larger with further aggregation and proliferation of cells on day 8. (B) Immunofluorescence images showing f-actin and DAPI (nuclei) of the forming aggregates on days 5 and 8 (Reproduced/adapted with permission from Zhao et al. (2014)).
Table 9.3 summarizes the application areas of bioprinting technology and lists the tissue types that have been successfully bioprinted along with the information of cell and bioink types, bioprinting modalities, and bioprinters used in fabrication of each tissue type.

9.6. Limitations

Despite the great progress in the field of bioprinting and its benefits in the presented application areas, there are several challenges to be circumvented for fabrication of functional tissues. Bioprinted tissues and bioprinting technologies entail different shortcomings for each distinct application as discussed in detail below.

9.6.1. Limitations in Bioprinting for Tissue Engineering and Regenerative Medicine

Bioprinting for bone tissue fabrication has been extensively studied as bone is one of the most widely engineered tissue types in regenerative medicine. The majority of bone tissue bioprinting attempts were mainly for the investigation of the science basis of stem cell differentiation toward osteogenic lineage within bioprinted tissue constructs (Kim et al., 2010). In vivo implantation has been performed for calvarium reconstruction or bone formation under subcutaneous tissue to assess in vivo performance and engraftment into the host (Bose et al., 2013); however, bioprinting of scale-up bone tissues for significant defects or substantial bone loss is still a challenge due to the lack of integration of blood vessels. In addition, most bioprinting approaches utilize soft materials such as hydrogels for bone tissue fabrication, which is not easy to implant into load-bearing sites of human body. Therefore, hydrogels should be reinforced with mechanically strong materials or integrated with supporting frames for large-size regeneration of bone tissue in vivo (Temple et al., 2014).

Table 9.3

Applications of Bioprinting Technologies

ApplicationTissue TypeCell Types BioprintedBioink or Substrate UsedBioprinting Modalities UsedBioprinters UsedRemarks
Tissue engineering and regenerative medicineBoneBone marrow–derived human mesenchymal stem cells (Gao et al., 2014); endothelial progenitor and multipotent stromal cells (Fedorovich et al., 2008); primary muscle-derived stem cells (Phillippi et al., 2008)PEGDMA (Gao et al., 2014); n-HA slurry (Keriquel et al., 2010); Matrigel™ and alginate (Fedorovich et al., 2008); bone morphogenetic protein (BMP-2) and fibrin (substrate) (Phillippi et al., 2008)Thermal inkjet (Gao et al., 2014); laser-induced droplet ejection (Keriquel et al., 2010); EBB (pneumatic) (Fedorovich et al., 2008); piezoelectric drop-on-demand (Phillippi et al., 2008)Hewlett–Packard deskjet (Gao et al., 2014); HT-BioLP workstation (Keriquel et al., 2010); Bioplotter (Fedorovich et al., 2008); MicroJet™ (Phillippi et al., 2008)Bioprinting bone tissue for critical-size defects is currently feasible, but bioprinting of scale-up vascularized bone tissues still remains elusive.
CardiacCardiac cells and HUVECs (Jakab et al., 2008); primary feline adult cardiomyocytes and HL1 cardiac muscle cells (Xu et al., 2009a,b); HUVEC and human mesenchymal stem cell (Gaebel et al., 2011); human cardiac–derived cardiomyocytes progenitor cells (Gaetani et al., 2012)Tissue spheroids and collagen type-I (biopaper) (Jakab et al., 2008); alginate (Xu et al., 2009a,b; Gaetani et al., 2012); PU (Gaebel et al., 2011)EBB (mechanical) (Jakab et al., 2008; Gaetani et al., 2012); thermal inkjet (Xu et al., 2009a,b); LIFT (Gaebel et al., 2011)nScrypt (Jakab et al., 2008); HP DeskJet 550 printers (Xu et al., 2009a,b); custom laser bioprinter (Gaebel et al., 2011); BioScaffolder (Gaetani et al., 2012)As cardiac cells do not have proliferation capability, scaffold-free bioprinting with high cell density is advantageous.
CartilageHuman chondrocytes (Cui et al., 2012a); rabbit elastic chondrocytes (Xu et al., 2013); bovine articular chondrocytes (Ozbolat et al., 2014); calve articular chondrocytes (Mannoor et al., 2013); human nasoseptal chondrocytes (Markstedt et al., 2015)PEGDMA (Cui et al., 2012a); fibrin–collagen type I (Xu et al., 2013); alginate25,26; alginate/nanocellulose (Markstedt et al., 2015)Thermal inkjet (Cui et al., 2012a); solenoid inkjet (Xu et al., 2013); EBB (pneumatic) (Ozbolat et al., 2014); EBB (microvalve) (Markstedt et al., 2015); EBB (mechanical) (Mannoor et al., 2013)HP Deskjet 500 printer (Cui et al., 2012a); XYZ plotter (Xu et al., 2013); MABP (Ozbolat et al., 2014); Fab @ home26; regenHU (Markstedt et al., 2015)Considerable work has been performed; however, zonally stratified articular cartilage is still a challenge and a great need in clinical use.
Table Continued

image

ApplicationTissue TypeCell Types BioprintedBioink or Substrate UsedBioprinting Modalities UsedBioprinters UsedRemarks
Heart valveAortic root sinus smooth muscle cells and aortic valve interstitial cells (Duan et al., 2013); aortic valvular interstitial cells (Duan et al., 2014)PEGDA and alginate (Hockaday et al., 2012); alginate and gelatin (Duan et al., 2013); methacrylated gelatin (Duan et al., 2014)EBB (mechanical) (Hockaday et al., 2012; Duan et al., 2013; Duan et al. 2014)Fab@home (Hockaday et al., 2012; Duan et al., 2013; Duan et al. 2014)Although anatomically accurate tissue models have been bioprinted, no performance evaluation has been done in vivo.
LiverHuman-induced pluripotent stem cells and human embryonic stem cells (Faulkner-Jones et al., 2015); HepG2 (Bertassoni et al., 2014a)Alginate (Faulkner-Jones et al., 2015); GelMA (Bertassoni et al., 2014a)Valve-based inkjet (Faulkner-Jones et al., 2015); EBB (mechanical) (Bertassoni et al., 2014a)Custom cell printer (Faulkner-Jones et al., 2015); NovoGen MMX Bioprinter™ (Bertassoni et al., 2014a)Limited progress has been made in bioprinting of liver tissues for regenerative medicine and patient-specific cells with long-term viability are still a concern.
LungEndothelial and epithelial cells (Horváth et al., 2015)Matrigel™ (substrate) (Horváth et al., 2015)Valve-based inkjet (Horváth et al., 2015)BioFactory® (Horváth et al., 2015)Although lung is hollow and reasonable easy to survive compared to some other organ types, human airway models for cytotoxicity testing seem to be in the near horizon.
NeuralMurine neural stem cells (Lee et al., 2010; Hsieh et al., 2015); Schwann cells and bone marrow stem cells (Owens et al., 2013)Collagen type I (substrate) (Lee et al., 2010); polyurethane (Hsieh et al., 2015); cell pellet and agarose (support) (Owens et al., 2013)Microvalve-based inkjet (Lee et al., 2010); EBB (mechanical) (Owens et al., 2013; Hsieh et al., 2015)Custom 4-head dispenser (Lee et al., 2010); NovoGen MMX Bioprinter™ (Owens et al., 2013)Nerve grafts are commercially available for short damages, but bioprinting has the capability to generate longer counterparts.
PancreasINS1E β-cells, mouse islets and human islets (Marchioli et al., 2015)Alginate and alginate/gelatin (Marchioli et al., 2015)EBB (pneumatic) (Marchioli et al., 2015)BioScaffolder (Marchioli et al., 2015)Beta cell source, its long-term functionality and viability, and availability of associated cells are still a challenge.
Table Continued

image

ApplicationTissue TypeCell Types BioprintedBioink or Substrate UsedBioprinting Modalities UsedBioprinters UsedRemarks
SkinHuman foreskin fibroblast and HaCaT keratinocytes (Lee et al., 2013); HaCaT keratinocyte cells and NIH 3T3 fibroblasts (Koch et al., 2012; Michael et al., 2013); human dermal microvascular endothelial cells (Yanez et al., 2014); amniotic fluid–derived stems (Skardal et al., 2012)Collagen type-I (Lee et al., 2013); collagen type-I on Matriderm TM (substrate) (Koch et al., 2012; Michael et al., 2013); thrombin (Yanez et al., 2014); collagen/fibrinogen and thrombin (Skardal et al., 2012)Microvalve-based inkjet (Lee et al., 2013; Skardal et al., 2012); LIFT (Koch et al., 2012; Michael et al., 2013); thermal inkjet (Yanez et al., 2014)Custom 8-head dispenser (Lee et al., 2013); LaBP (Koch et al., 2012; Michael et al., 2013); Modified HP Deskjet 340 (Yanez et al., 2014); skin printer (Skardal et al., 2012)Great progress has been made in skin bioprinting, but advancements are needed for further improvement in scarless tissue formation and integration of sweat glands.
VascularHUVSMCs (Zhang et al., 2015; Dolati et al., 2014; Norotte et al., 2009); chondrocytes (Zhang et al., 2013b; Yu et al., 2013); 3T3 mouse fibroblasts (Christensen et al., 2015); HUVEC (Lee et al., 2014; Zhao et al., 2012) and normal human lung fibroblast (Lee et al., 2014); human skin fibroblasts (Norotte et al., 2009)Alginate (Zhang et al., 2015; Christensen et al., 2015); alginate and chitosan (Zhang et al., 2013b; Yu et al., 2013), alginate with carbon nanotubes (Dolati et al., 2014); fibrin (Lee et al., 2014); collagen (Zhao et al., 2012); GelMA (Bertassoni et al., 2014a); tissue spheroids (Norotte et al., 2009)Coaxial nozzle extrusion (Zhang et al., 2013b; 2015; Yu et al., 2013; Dolati et al., 2014) piezo-inkjet (Christensen et al., 2015); valve-based inkjet (Lee et al., 2014; Zhao et al., 2012)Nordson (Zhang et al., 2013b; 2015; Yu et al., 2013; Dolati et al., 2014); Microfab (Christensen et al., 2015); custom multihead dispenser (Lee et al., 2014; Zhao et al., 2012); NovoGen MMX Bioprinter™(Bertassoni et al., 2014a; Norotte et al., 2009)Long-term in vivo efficacy of bioprinted blood vessel has not been tested yet. For organ fabrication, enabling technologies are needed to bioprint vascular network in multiscale.
Composite3T3 fibroblasts and myoblasts (Merceron et al., 2015); MSCs (Feng et al., 2011; Fedorovich et al., 2011) and chondrocytes (Fedorovich et al., 2011); osteoblast and chondrocytes (Park et al., 2014; Shim et al., 2012); HUVSMCs and fibroblasts (Yu et al., 2014)Hyaluronic acid/gelatin/fibrinogen and PU (Merceron et al., 2015); alginate (Fedorovich et al., 2011; Shim et al., 2012; Yu et al., 2014); collagen type I and hyaluronic acid (Park et al., 2014); tissue strands (Yu et al., 2014)EBB (pneumatic) (Merceron et al., 2015; Fedorovich et al., 2011; Park et al., 2014; Shim et al., 2012; Yu et al., 2014); acoustic-based droplet (Feng et al., 2011)Custom multinozzle head (Merceron et al., 2015; Fedorovich et al., 2011; Feng et al., 2011); MtoBS (Park et al., 2014; Shim et al., 2012); MABP (Yu et al., 2014)Bioprinting of composite tissues is highly vital, and a substantial progress is needed to generate organ-level constructs by integrating tissues such as bone, muscle, tendon, nerve, blood vessels, and skin together.
Table Continued

image

ApplicationTissue TypeCell Types BioprintedBioink or Substrate UsedBioprinting Modalities UsedBioprinters UsedRemarks
Pharmaceutics and drug testingLiverHepG2 (Chang et al., 2010; 2008); epithelial cells and hepatocytes (Snyder et al., 2011); human hepatocytes, hepatic satellite cells, and endothelial cells (Roskos et al., 2015)Alginate (Chang et al., 2010; 2008); Matrigel™ (Snyder et al., 2011); cell pellet (Roskos et al., 2015);EBB (valve) (Snyder et al., 2011),(Chang et al., 2010; 2008); EBB (mechanical) (Roskos et al., 2015)Multinozzle system (Snyder et al., 2011),(Chang et al., 2010; 2008); NovoGen MMX Bioprinter™ (Roskos et al., 2015)Bioprinted liver tissue models have a great potential in early drug discovery, but a standard model is yet to be developed.
Cell droplets for high-throughput arraysEscherichia coli (Rodríguez-Dévora et al., 2012); primary smooth muscle cells from rat bladder (Xu et al., 2009a); mouse embryonic stem cells, fibroblasts, AML-12 hepatocytes, human Raji cells, and HL-1 cardiomyocytes (Demirci and Montesano, 2007)Alginate and soy agar (substrate) (Rodríguez-Dévora et al., 2012); collagen (Xu et al., 2009a); sucrose and dextrose (Demirci and Montesano, 2007)Thermal inkjet (Rodríguez-Dévora et al., 2012); acoustic-based (Xu et al., 2009a; Demirci and Montesano, 2007)Modified-HP (Rodríguez-Dévora et al., 2012); A custom acoustic bioprinter (Xu et al., 2009a; Demirci and Montesano, 2007)Picoliter size of droplets can be generated with high accuracy in droplet size and location, which is highly efficient for high-throughput arrays for drug testing.
Transplantation and clinicsBone, cartilage and skinAmniotic fluid–derived stem cells and Bone marrow–derived MSCs (Skardal et al., 2012); bone marrow stem cells (Ozbolat, 2015a)Polycaprolactone (Zopf et al., 2013); nHA(Keriquel et al., 2010); collagen–fibrin (Skardal et al., 2012); alginate and Pluronic collagen (Ozbolat, 2015a)Laser-based printing (Zopf et al., 2013); laser-based bioprinting (Keriquel et al., 2010); piezo-inkjet (Skardal et al., 2012); extrusion (pneumatic) (Ozbolat, 2015a)EOS P 100 Formiga system (Zopf et al., 2013); HT-BioLP workstation (Keriquel et al., 2010); custom inkjet printer (Skardal et al., 2012); MABP (Ozbolat, 2015a)Only transplantation of a splint (using a nonbioprinting technique) into a human and in situ bioprinting on animal models have been achieved in operating rooms
Table Continued

image

ApplicationTissue TypeCell Types BioprintedBioink or Substrate UsedBioprinting Modalities UsedBioprinters UsedRemarks
Cancer researchOvarian cancerHuman ovarian cancer cells and MRC-5 fibroblasts (Xu et al., 2011)Matrigel™ (substrate) (Xu et al., 2011)Solenoid-valve ejection (Xu et al., 2011)A custom dual-head bioprinter (Xu et al., 2011)Only a technological platform has been demonstrated so far, but bioprinting of biomimetically developed ovarian cancer model is yet to be researched.
Cervical cancerHeLa (Zhao et al., 2014; Huang et al., 2014); 10 T1/2 fibroblasts (Huang et al., 2014)Gelatin/alginate/fibrinogen (Zhao et al., 2014); PEGDA (Huang et al., 2014)EBB (mechanical) (Zhao et al., 2014) and laser-based projection printing (Huang et al., 2014)Cell assembly system I (Zhao et al., 2014); DMD-PP (Huang et al., 2014)Only a very few attempts, at the basic research level, have been made for bioprinting of cervical cancer models.
Breast cancerMSC-differentiated adipose cells, mammary fibroblasts, and endothelial cells (King et al., 2014)Cell pellet (King et al., 2014)EBB (mechanical) (King et al., 2014)NovoGen MMX Bioprinter™ (King et al., 2014)Further substantial development is needed to use the bioprinted breast cancer model for cancer screening and drug testing.

image

EBB, Extrusion-based bioprinting; GelMA, gelatin-methacrylamide; n-HA, hydroxyapatite nanoparticle; HUVEC, human vascular endothelial cell; LIFT, laser-induced forward transfer; MSC, mesenchymal stem cell; PEGDA, polyethylene glycol diacrylate; PEGDMA, poly(ethylene glycol) dimethacrylate; PU, polyurethane.

Bioprinting for cardiac tissue fabrication has been performed at cardiac patch level and successful regeneration of a cardiac patch in vivo has demonstrated so far, but bioprinting of the entire heart is still considered science fiction (Dababneh and Ozbolat, 2014). While bioprinting of hierarchical vascular network is the critical step toward fabrication of organs at the clinically relevant size, bioprinting of a whole-heart model vitally depends on the integration of vascular network as heart needs substantial vascularization due to its hectic metabolic activity. One of the major hurdles in cardiac tissue regeneration is that cardiomyocytes do not proliferate, and it is not trivial to obtain sufficient number of cardiomyocytes (Mollova et al., 2013); therefore, stem cell engineering for sufficient and practical derivation of stable cardiac cells will be the key factor in further advancements for cardiac tissue bioprinting.
Bioprinting for cartilage tissue fabrication has been well studied in the literature; however, bioprinting of zonally stratified articular cartilage has yet to be demonstrated. Articular cartilage is the lining on articulating surfaces of diarthroidal joints, and it functions as a shock absorber to distribute the load from weight and daily activities (Aydelottea et al., 1988). The superficial zone takes up to 20% of the total cartilage thickness and contains densely packed collagen fibers in parallel to the articulating surface, and cells in that zone secrete lubricants to minimize wear and tear to the joint. The deeper zones including middle zone, deep zone, and calcified zone are relatively less in cell density and have thicker collagen bundles, which are perpendicular to the articulating surface to resist compression force. This unique arrangement is formed due to the external loads over time, which is transmitted through the matrix of the tissue and converted into a biochemical signal, alerting cells to either produce more or catabolize existing ECM (Makris et al., 2014). Therefore, currently existing strategies are not sufficient in enabling controlled zonal differentiation for nativelike engineered articular cartilage tissues (Kock et al., 2012; Makris et al., 2014). Although not attempted yet, integration of bioprinted cartilage tissue with subchondral bone may be another challenge as it is a common problem in cartilage tissue repair (Makris et al., 2015).
Bioprinting for heart valve fabrication has been attempted by one research group so far, and different combinations of hydrogels, including alginate, PEGDA, gelatin, and Me-Gel, have been used for heart valve bioprinting (Jana and Lerman, 2015). Currently, there is no in vivo implantation demonstrated so far to test the efficacy and performance of the bioprinted cardiac valves. While only hydrogel-based materials have been used in the literature, long-term cultivation of bioprinted samples and their biological, mechanical, and functional characterization have not been reported so far. To transplant bioprinted heart valves, the constructs should possess sufficient mechanical properties to withstand the physiological blood flow pressure while providing an appropriate microenvironment to promote survival and growth of seeded cells.
Bioprinting for liver tissue fabrication has been primarily performed using human tumor–derived cell lines (i.e., HepG2) (Bertassoni, Cardoso, et al., 2014a; Tourlomousis and Chang, 2015). Compared to primary hepatocytes, HepG2s have reduced functional performance such as low level of ammonia removal, aminoacid metabolism, cytochrome P450s, and lack of a myriad of drug-metabolizing functions (Palakkan et al., 2013). Although no transplantation has been pursued with tumor-derived cell lines, these cells have potential to transmit the tumorigenic products or possible complications arising from their transmission (Knowles et al., 1980). Although primary hepatocytes are preferred, their limited availability and poor in vitro proliferation capability reduce their applicability in bioprinting research. In addition to cell source–related issues, bioprinting of scale-up liver tissue is still a challenge as there has not been any work reported yet. Transplantation of bioprinted liver tissue models has not been demonstrated yet, and long-term functionality and performance of such a tissue still remain elusive.
Bioprinting for lung tissue fabrication is still in its infancy, and there has been only a single work reported so far (Horváth et al., 2015). As lung is a highly complicated organ in its morphology comprising around half a billion alveoli to facilitate sufficient gas exchange. As the demonstrated multilayer pulmonary tissue model represents the air barrier mimicking the alveolus wall, bioprinting of an anatomically correct pulmonary tissue model with a highly dense capillary network is still a major cornerstone toward fabrication of a functional lung parenchyma (Nichols et al., 2009). Such a parenchymal tissue model should be mechanically stable; structurally strong and elastic enough to withstand air pressure; permeable enough to facilitate gas exchange across the tissue barrier; and vascularized enough to allow transfer of the gas from and to capillaries.
Bioprinting for nervous tissue fabrication is still in its infancy, and bioprinting for functional central nervous system components is yet far from the reality; however, recent work on peripheral nervous system by understanding neural stem growth through bioprinting (Lee et al., 2010) and fabricating heterocellular nerve grafts (Owens et al., 2013) for in vivo testing demonstrated the promise of bioprinting in this challenging field. A wide array of nerve grafts has been studied, including autologous, allogeneic, and xenogeneic neural tissues, naturally derived polymers (i.e., collagen, laminin, fibrin, fibronectin, hyaluronan), and polysaccharides (chitosan, alginate, and agarose); however, each of them has certain limitation such as but not limited to donor site morbidity, limited availability, mismatch in size, immunogenic issues, limited functionality, and degradation and associated by-products (Gu et al., 2014). Nerve conduits support generation of axon growth, but the regenerated tissue, beyond a certain length, does not imitate the native anatomy and functionality closely. Bioprinting thus has the ability to recapitulate the complex nerve anatomy, with conduits of nerve bundles along with blood vessels and other connective tissue, which can be further imitated with electrical or biochemical stimulation (Ghasemi-Mobarakeh et al., 2009).
Bioprinting for pancreatic tissue fabrication is relatively new, and a limited number of research have been done in this regard (Marchioli et al., 2015; Akkouch et al., 2015). Cells used in these studies were limited to mouse insulinoma cell, or islets of Langerhans were directly bioprinted as an alternative approach to manual islet encapsulation. As primary β-cells and islets of Langerhans are not stable in vitro, pancreatic regeneration work has been embodied to the extent where insulinoma cells or mouse cell lines used only, which is far from the human islet physiological conditions. In addition, pancreatic islets are comprised of multiple cells types (including α-cells, endothelial cells, δ-cells, and PP cells), which is highly crucial for islet biology and functionality (Migliorini et al., 2014), and current tissue engineering approaches do not enable successful derivation and aggregation of these cells as observed in islets of Langerhans.
Bioprinting for skin tissue fabrication has aimed at compartmentalization of dermal and epidermis layers by depositing fibroblasts and keratinocytes layer by layer in addition to a recent attempt in stimulating formation of microcapillaries by incorporating endothelial cells within layers (Yanez et al., 2014). Although bioprinting stratified zones has been demonstrated extensively, generating nativelike complete skin tissue with all components and appendages, such as hair follicles and sweat glands, and minimum wound contraction and scar formation, still remains elusive (Michael et al., 2013). Although recent regenerative medicine practices focus on understanding the regeneration of de nova hair follicles using human dermal papilla cells in 3D spheroid culture model (Higgins et al., 2013), incorporation of such a model has not been attempted in skin tissue bioprinting yet.
Bioprinting for vascular tissue fabrication has been well studied in the literature; several challenges are yet to be circumvented such as but not limited to the formation of physiologically relevant vascular grafts and their long-term evaluation in vivo (Ozbolat, 2015a). Although a recent effort demonstrated the functionality, suture retention and in vivo endothelialization of the vascular grafts on a murine model (Itoh et al., 2015), long-term functionality and durability, as one of the common problems in engineered vascular grafts (Kurobe et al., 2012), still remain elusive. For vascular network bioprinting in scale-up engineered tissues, bioprinting of multiscale hierarchical vascular networks from arteries and veins down to capillaries has not been demonstrated yet (Ozbolat, 2015b). Although integration of macroscale channels with the microcapillaries in fibrin has been demonstrated through gluing endothelial cells (Lee et al., 2014), generation of microcapillaries around free-standing macrovasculature still needs to be demonstrated toward bioprinting of scalable tissues and organs.
Bioprinting for composite tissue fabrication will be a highly influential step in regenerative medicine, but current attempts could be able to demonstrate very simple structures, not even close to their native counterparts. As significant amount of work has been done in the area of osteochondral tissue fabrication (Di Luca et al., 2015), bioprinting of such structures will be one of the first composite tissue models in the near horizon. Despite the great effort in inducing osteogenic and chondrogenic differentiation locally for knee replacement, biomimetic development of such anisotropy with gradual transition in mechanical, structural, and biological properties of osteochondral tissue models with well integration at the interface of cartilage and bone is still a challenge.

9.6.2. Limitations in Bioprinting for Transplantation and Clinics

Bioprinting for transplantation and clinics is a trending area of application, and a number of bioprinted tissue types have been tested in vivo; however, translation of bioprinting technology into clinics for human has yet to be succeeded as there are several impediments down the road. First of all, due to the lack of hierarchical vascular network, bioprinting of scale-up tissues is still a challenge (Ozbolat and Yu, 2013). Tissues at the clinically relevant sizes, except cartilage and skin, have not been reported yet (Ozbolat, 2015a). In the literature, bioprinted tissues have been implanted into vascularized sites on animal models such as subcutaneous tissue (Marchioli et al., 2015), but they need to be implanted into less immune-responsive sites and connected to blood stream as an intravascular device when bioprinted in larger sizes. Although a myriad of bioprinted tissue constructs have been demonstrated, one of the major issues is the cell source considering immunogenicity issues. Therefore, stem cell use in bioprinted tissues is a key factor in enabling translation of bioprinted tissues into clinical practices. There is no bioprinted product undergone FDA approval yet, and FDA has not taken any consideration on the classification of bioprinters and bioprinted tissues but as the technology evolves further through clinical trials, then the first product going through the FDA regulations will be exemplary for preceding ones.

9.6.3. Limitations in Bioprinting for Drug Screening and High-Throughput Assays

Bioprinting for drug screening and high-throughput assays is a highly powerful technology as 3D tissue models can be precisely formed in a high-throughput manner. Both extrusion- and inkjet-based bioprinting technologies can be used to create heterocellular tissue environment to predict drug toxicity (Feng et al., 2011). Currently employed techniques in liver fabrication employ HepG2 liver carcinoma cell line, which does not represent the human liver physiology closely (Palakkan et al., 2013). Although complex patterned structures have been successfully integrated within a perfusion system to assess drug toxicity in a control manner, the cell type used as well as the lack of other associated cell types reduces the reliability of the tissue model in predicting the performance of drugs (Chang et al., 2010). In this regard, Organovo's liver tissue model stands as a promising tool to represent a biomimetically developed liver tissue model for drug testing (Roskos et al., 2015). The employed scaffold-free technique comprising stem cell–derived hepatocytes as well as hepatic satellite cells and endothelial cells facilitates physiologically relevant liver tissue, and further progress is needed to create a standardized bioprinted liver tissue model for pharmaceutical industry. In addition to liver tissue models, cells can be bioprinted in a high-throughput manner to study their behavior as well as bacterial cells can be bioprinted along with drugs to understand the interactions between applied drugs and bioprinted different bacterial strains (Rodríguez-Dévora et al., 2012). Different dosages can be bioprinted in high-throughput manner to effectively analyze the influence of different antibiotics on bacterial cells; however, a translational effort is yet to be demonstrated for the pharmaceutical industry.

9.6.4. Limitations in Bioprinting for Cancer Research

Bioprinting for cancer research is relatively new, and there are only a few efforts in fabrication of tumor models (i.e., ovarian, cervical, and breast tissue models) (Knowlton et al., 2015). To study cancer pathogenesis, inkjet-based bioprinting stands as a promising tool in generating 3D tumor models with multiple cell types in controlled composition; however, there is no well-established model in abovementioned tumor types that can be used in pharmaceutics industry yet. Formation of larger tissues as an in vitro cancer tissue model for exploration of cancer metastasis has not been attempted using bioprinting yet; however, there are various microfluidics-based platforms with closely recapitulated cancer microenvironment for cancer metastasis research (Jeon et al., 2015; Zervantonakis et al., 2012). These platforms lack precise placement of multiple cell types in a localized manner. Bioprinting has the capability of patterning the cellular microenvironment, but further research is required in angiogenesis work in bioprinting to establish standard cancer tissue models for cancer metastasis research.

9.7. Future Directions

Currently, there are around 15 different tissue types attempted in bioprinting field; however, there exist a number of other tissues in human body that have not been investigated yet. In this regard, diverging the focus into other tissue and organ models, which are challenging but can revolutionize the medicine, is necessary in the near horizon. This also depends on the advancements in tissue engineering field as bioprinting research vitally depends on our understanding of unexplored tissue types. In addition, bioprinting of new types of organs, such as bionic organs or organs that do not exist in nature, will be one of the new directions in bioprinting.
Majority of bioprinting research has evolved around homocellular tissue construct using a single cell type; however, native tissues have a heterocellular nature with multiple cell type patterned in a highly complex anatomy. Although bioprinting-simplified tissue models are relatively acceptable in basic research, functional tissue fabrication for clinics or pharmaceutics necessitates inclusion of multiple cell types as some of the functionality of cells can be enabled or further boosted by cell–cell interactions. Bioprinting multiple cell types requires further understanding of the optimum culture conditions including the right medium and cocktail to support growth and behaviors of multiple cell types.
Current efforts in translation of bioprinting into transplantation are mainly limited to murine models. As murine models are highly small and their physiological conditions are not closely relevant to that of humans. For example, a small representative pancreatic model with bioprinted pancreatic islets can be implanted into a diabetic murine model and can regulate insulin secretion, but average volume of mouse model is approximately 100,000 smaller than that of human model (Bock et al., 2003; Ionescu-Tirgoviste et al., 2015). Therefore, bioprinting of scale-up tissues and organs is highly important in fabricating tissues in clinically relevant dimensions. Larger animals such as a pig model can be a transitional step toward trials on humans as these models represent human physiology closer than small animals (Meurens et al., 2012). To scale-up bioprinted organs, bioprinting of hierarchical vascular network is vital as detailed out in Chapter 8. To generate clinically relevant tissue and organ models, mechanical strength and elasticity as well as long-term structural stability is highly important. In addition, integration of nerve tissues and establishing innervation is a vital step toward functional tissue and organ fabrication as tissues such as cardiac, muscle, and skin need innervation, which is currently a challenge in bioprinting. Majority of the bioprinting research entails the use of hydrogel-based bioink due to their favorable environment for 3D growth of cells; however, hydrogels are highly weak in mechanical properties when they are used at a concentration compatible with cell proliferation. In addition, biodegradable materials such as thermoplastics can be used as a supporting frame for the scale-up tissues and organs. Current biomaterials using synthetic polymers are highly strong, but their degradation takes prolonged times and does not synchronize with the tissue regeneration process. Therefore, new biomaterial development is highly crucial for the advancement in bioprinting of tissues and organs for clinical use. In addition to biomaterials, selection of the right cell source is also a critical factor in translation of bioprinting technologies into clinics and pharmaceutics. For personalized medicine, as it is highly challenging to acquire different primary cell types. Stem cell stands as a promising source of cells, and further advances are needed to establish highly standard protocols for differentiation of stem cells into various stable and functional cell lineages.
In addition, bioprinting for cancer research needs further advancement to enable philologically relevant microenvironment for cancer pathology and metastasis. There exist 3D tumor models that need to be integrated with vascular network and the rest of the parenchymal tissue using bioprinting. Tumor tissue models should be precisely placed within bioprinted vascularized parenchymal tissues to study tumor growth and tumor cell intravasation and extravasation. Physiologically relevant cancer microenvironment has been well established in microfluidics-based organ-on-chip-models, but bioprinting can bring precise control on positioning tumor spheroids and cells as well as tight control on the formation of microvascular network.

9.8. Summary

This chapter presented a detailed discussion on the application areas of 3D bioprinting technology. The noble effort by Klebe in cytoscribing cells has opened up great avenues in bioprinting of living tissue and organ constructs, which has been adopted into various fields including basic research in tissue engineering, regenerative medicine and cancer pathogenesis, tissue printing for transplantation and clinics, and some recent efforts in beginning translational technologies in pharmaceutics for drug testing and high-throughput screening. With the current advancements in various bioprinting technologies as well as progress in cell and biomaterial research, fabrication of physiologically relevant scalable tissues is expected to be feasible in the near future.

References

Akkouch A, Yu Y, Ozbolat I.T. Microfabrication of scaffold-free tissue strands for three-dimensional tissue engineering. Biofabrication. 2015;7(3):31002.

Aydelottea M, Greenhilla R, Kuettnerab K. Differences between sub-populations of cultured bovine articular chondrocytes. II. Proteoglycan metabolism. Connective Tissue Research. 1988;18(3):223–234.

Bertassoni L.E, Cardoso J.C, et al. Direct-write bioprinting of cell-laden methacrylated gelatin hydrogels. Biofabrication. 2014;6(2):024105.

Bertassoni L.E, Cecconi M, et al. Hydrogel bioprinted microchannel networks for vascularization of tissue engineering constructs. Lab on a Chip. 2014;14(13):2202–2211.

Bhise N.S, et al. A liver-on-a-chip platform with bioprinted hepatic spheroids. Biofabrication. 2016;8(1) doi: 10.1088/1758-5090/8/1/014101.

Blaeser A, et al. Biofabrication under fluorocarbon: a novel freeform fabrication technique to generate high aspect ratio tissue-engineered constructs. BioResearch Open Access. 2013;2(5):374–384.

Bock T, Pakkenberg B, Buschard K. Increased islet volume but unchanged islet number in ob/ob mice. Diabetes. 2003;52(7):1716–1722.

Bose S, Vahabzadeh S, Bandyopadhyay A. Bone tissue engineering using 3D printing. Materials Today. 2013;16(12):496–504.

Chang R, et al. Biofabrication of a three-dimensional liver micro-organ as an in vitro drug metabolism model. Biofabrication. 2010;2(4):45004.

Chang R, Nam J, Sun W. Direct cell writing of 3D microorgan for in vitro pharmacokinetic model. Tissue Engineering Part C: Methods. 2008;14(2):157–166.

Christensen K, et al. Freeform inkjet printing of cellular structures with bifurcations. Biotechnology and Bioengineering. 2015;112(5):1047–1055.

Cohen D.L, et al. Increased mixing improves hydrogel homogeneity and quality of three-dimensional printed constructs. Tissue Engineering Part C: Methods. 2010;17(2):239–248.

Coruh A, Yontar Y. Application of split-thickness dermal grafts in deep partial- and full-thickness burns: a new source of auto-skin grafting. Journal of Burn Care and Research. 2012;33(3).

Cui X, Breitenkamp K, Finn M.G, et al. Direct human cartilage repair using three-dimensional bioprinting technology. Tissue Engineering Part A. 2012;18(11–12):1304–1312.

Cui X, Breitenkamp K, Lotz M, et al. Synergistic action of fibroblast growth factor-2 and transforming growth factor-beta1 enhances bioprinted human neocartilage formation. Biotechnology and Bioengineering. 2012;109(9):2357–2368. .

Dababneh A.B, Ozbolat I.T. Bioprinting technology: a current state-of-the-art review. Journal of Manufacturing Science and Engineering. 2014;136(6):061016.

Demirci U, Montesano G. Single cell epitaxy by acoustic picolitre droplets. Lab on a Chip. 2007;7(9):1139–1145.

Dolati F, et al. In vitro evaluation of carbon-nanotube-reinforced bioprintable vascular conduits. Nanotechnology. 2014;25(14):145101.

Duan B, et al. 3D Bioprinting of heterogeneous aortic valve conduits with alginate/gelatin hydrogels. Journal of Biomedical Materials Research Part A. 2013;101A(5):1255–1264.

Duan B, et al. Three-dimensional printed trileaflet valve conduits using biological hydrogels and human valve interstitial cells. Acta Biomaterialia. 2014;10(5):1836–1846.

Eulalio A, et al. Functional screening identifies miRNAs inducing cardiac regeneration. Nature. 2012;492(7429):376–381.

Faulkner-Jones A, et al. Bioprinting of human pluripotent stem cells and their directed differentiation into hepatocyte-like cells for the generation of mini-livers in 3D. Biofabrication. 2015;7(4):44102.

Fedorovich N.E, et al. Biofabrication of osteochondral tissue equivalents by printing topologically defined, cell-laden hydrogel scaffolds. Tissue Engineering Part C: Methods. 2011;18(1):33–44.

Fedorovich N.E, et al. Three-dimensional fiber deposition of cell-laden, viable, patterned constructs for bone tissue printing. Tissue Engineering Part A. 2008;14(1):127–133.

Feng X, et al. Microengineering methods for cell-based microarrays and high-throughput drug-screening applications. Biofabrication. 2011;3(3):34101.

Gaebel R, et al. Patterning human stem cells and endothelial cells with laser printing for cardiac regeneration. Biomaterials. 2011;32(35):9218–9230.

Gaetani R, et al. Cardiac tissue engineering using tissue printing technology and human cardiac progenitor cells. Biomaterials. 2012;33(6):1782–1790.

Gao G, et al. Bioactive nanoparticles stimulate bone tissue formation in bioprinted three-dimensional scaffold and human mesenchymal stem cells. Biotechnology Journal. 2014;9(10):1304–1311.

Ghasemi-Mobarakeh L, et al. Electrical stimulation of nerve cells using conductive nanofibrous scaffolds for nerve tissue engineering. Tissue Engineering Part A. 2009;15(11):3605–3619.

Gruene M, et al. Laser printing of stem cells for biofabrication of scaffold-free autologous grafts. Tissue Engineering Part C: Methods. 2010;17(1):79–87.

Gu X, Ding F, Williams D.F. Neural tissue engineering options for peripheral nerve regeneration. Biomaterials. 2014;35(24):6143–6156.

Gudapati H, Dey M, Ozbolat I. A comprehensive review on droplet-based bioprinting: past, present and future. Biomaterials. 2016;102:20–42.

Gurkan U.A, et al. Engineering anisotropic biomimetic fibrocartilage microenvironment by bioprinting mesenchymal stem cells in nanoliter gel droplets. Molecular Pharmaceutics. 2014;11(7):2151–2159.

Higgins C.A, et al. Microenvironmental reprogramming by three-dimensional culture enables dermal papilla cells to induce de novo human hair-follicle growth. Proceedings of the National Academy of Sciences. 2013;110(49):19679–19688.

Hirt M.N, Hansen A, Eschenhagen T. Cardiac tissue engineering: state of the art. Circulation Research. 2014;114(2):354–367.

Hockaday L.A, et al. 3D-Printed hydrogel technologies for tissue-engineered heart valves. 3D Printing and Additive Manufacturing. 2014;1(3):122–136.

Hockaday L.A, et al. Rapid 3D printing of anatomically accurate and mechanically heterogeneous aortic valve hydrogel scaffolds. Biofabrication. 2012;4(3):35005. .

Horváth L, et al. Engineering an in vitro air-blood barrier by 3D bioprinting. Scientific Reports. 2015;5:7974.

Hsieh F.-Y, Lin H.-H, Hsu S.-H. 3D bioprinting of neural stem cell-laden thermoresponsive biodegradable polyurethane hydrogel and potential in central nervous system repair. Biomaterials. 2015;71:48–57.

Huang T, et al. 3D printing of biomimetic microstructures for cancer cell migration. Biomedical Microdevices. 2014;16(1):127–132.

Ionescu-Tirgoviste C, et al. A 3D map of the islet routes throughout the healthy human pancreas. Scientific Reports. 2015;5:14634.

Itoh M, et al. Scaffold-free tubular tissues created by a Bio-3D printer undergo remodeling and endothelialization when implanted in rat aortae. PLoS One. 2015;10(9):e0136681.

Jakab K, et al. Tissue engineering by self-assembly and bio-printing of living cells. Biofabrication. 2010;2(2):22001.

Jakab K, et al. Tissue engineering by self-assembly of cells printed into topologically defined structures. Tissue Engineering Part A. 2008;14(3):413–421.

Jana S, Lerman A. Bioprinting a cardiac valve. Biotechnology Advances. 2015;33(8):1503–1521.

Jeon J.S, et al. Human 3D vascularized organotypic microfluidic assays to study breast cancer cell extravasation. Proceedings of the National Academy of Sciences. 2015;112(1):214–219.

Keriquel V, et al. In vivo bioprinting for computer- and robotic-assisted medical intervention: preliminary study in mice. Biofabrication. 2010;2(1):14101.

Kesti M, et al. A versatile bioink for three-dimensional printing of cellular scaffolds based on thermally and photo-triggered tandem gelation. Acta Biomaterialia. 2015;11:162–172.

Kim K, et al. Stereolithographic bone scaffold design parameters: osteogenic differentiation and signal expression. Tissue Engineering Part B: Reviews. 2010;16(5):523–539.

King S.M, Presnell S.C, Nguyen D.G. Abstract 2034: development of 3D bioprinted human breast cancer for in vitro drug screening. Cancer Research. 2014;74(Suppl. 19):2034.

Knowles B.B, Howe C.C, Aden D.P. Human hepatocellular carcinoma cell lines secrete the major plasma proteins and hepatitis B surface antigen. Science. 1980;209(4455):497–499.

Knowlton S, et al. Advancing cancer research using bioprinting for tumor-on-a-chip platforms. International Journal of Bioprinting. 2016;2(2):3–8.

Knowlton S, et al. Bioprinting for cancer research. Trends in Biotechnology. 2015;33(9):504–513.

Koch L, et al. Skin tissue generation by laser cell printing. Biotechnology and Bioengineering. 2012;109(7):1855–1863.

Kock L, van Donkelaar C.C, Ito K. Tissue engineering of functional articular cartilage: the current status. Cell and Tissue Research. 2012;347(3):613–627.

Kurobe H, et al. Concise review: tissue-engineered vascular grafts for cardiac surgery: past, present, and future. Stem Cells Translational Medicine. 2012;1(7):566–571.

Lee V, et al. Design and fabrication of human skin by three-dimensional bioprinting. Tissue Engineering Part C: Methods. 2013;20(6):473–484.

Lee V, et al. Generation of multi-scale vascular network system within 3D hydrogel using 3D bio-printing technology. Cellular and Molecular Bioengineering. 2014;7(3):460–472.

Lee Y.-B, et al. Bio-printing of collagen and VEGF-releasing fibrin gel scaffolds for neural stem cell culture. Experimental Neurology. 2010;223(2):645–652.

Leon-Villapalos J, Eldardiri M, Dziewulski P. The use of human deceased donor skin allograft in burn care. Cell and Tissue Banking. 2010;11(1):99–104. .

Lorber B, et al. Adult rat retinal ganglion cells and glia can be printed by piezoelectric inkjet printing. Biofabrication. 2014;6(1):15001.

Lozano R, et al. 3D printing of layered brain-like structures using peptide modified gellan gum substrates. Biomaterials. 2015;67:264–273.

Di Luca A, Van Blitterswijk C, Moroni L. The osteochondral interface as a gradient tissue: from development to the fabrication of gradient scaffolds for regenerative medicine. Birth Defects Research Part C: Embryo Today: Reviews. 2015;105(1):34–52.

Makris E.A, MacBarb R.F, Paschos N.K, Hu J.C, Athanasiou K.A. Combined use of chondroitinase-ABC, TGF-Beta1, and collagen crosslinking agent lysyl oxidase to engineer functional neotissues for fibrocartilage repair. Biomaterials. 2014;35(25):6787–6796.

Makris E.A, et al. Repair and tissue engineering techniques for articular cartilage. Nature Reviews Rheumatology. 2015;11(1):21–34.

Malone E, Lipson H. Fab@Home: the personal desktop fabricator kitnull. Rapid Prototyping Journal. 2007;13(4):245–255.

Mannoor M.S, et al. 3D printed bionic ears. Nano Letters. 2013;13(6):2634–2639.

Marchioli G, et al. Fabrication of three-dimensional bioplotted hydrogel scaffolds for islets of Langerhans transplantation. Biofabrication. 2015;7(2):25009.

Markstedt K, et al. 3D bioprinting human chondrocytes with nanocellulose–alginate bioink for cartilage tissue engineering applications. Biomacromolecules. 2015;16(5):1489–1496.

Merceron T.K, et al. A 3D bioprinted complex structure for engineering the muscle–tendon unit. Biofabrication. 2015;7(3):35003.

Metcalfe A.D, Ferguson M.W.J. Tissue engineering of replacement skin: the crossroads of biomaterials, wound healing, embryonic development, stem cells and regeneration. Journal of the Royal Society Interface. 2007;4(14):413–437.

Meurens F, et al. The pig: a model for human infectious diseases. Trends in Microbiology. 2012;20(1):50–57.

Michael S, et al. Tissue engineered skin substitutes created by laser-assisted bioprinting form skin-like structures in the dorsal skin fold chamber in mice. PLoS One. 2013;8(3):e57741.

Migliorini A, Bader E, Lickert H. Islet cell plasticity and regeneration. Molecular Metabolism. 2014;3(3):268–274.

Miller J.S, et al. Rapid casting of patterned vascular networks for perfusable engineered three-dimensional tissues. Nature Materials. 2012;11(9):768–774.

Mollova M, et al. Cardiomyocyte proliferation contributes to heart growth in young humans. Proceedings of the National Academy of Sciences. 2013;110(4):1446–1451.

Moroni L, De Wijn J.R, Van Blitterswijk C.A. 3D fiber-deposited scaffolds for tissue engineering: influence of pores geometry and architecture on dynamic mechanical properties. Biomaterials. 2006;27:974–985.

Nam K.-H, et al. Biomimetic 3D tissue models for advanced high-throughput drug screening. Journal of Laboratory Automation. 2015;20(3):201–215.

Nelson B. 3-dimensional bioprinting makes its mark: new tissue and organ printing methods are yielding critical new tools for the laboratory and clinic. Cancer Cytopathology. 2015;123(4):203–204.

Nichols J.E, Niles J.A, Cortiella J. Design and development of tissue engineered lung: progress and challenges. Organogenesis. 2009;5(2):57–61.

Nishiyama Y, et al. Development of a three-dimensional bioprinter: construction of cell supporting structures using hydrogel and state-of-the-art inkjet technology. Journal of Biomechanical Engineering. 2008;131(3):35001.

Norotte C, et al. Scaffold-free vascular tissue engineering using bioprinting. Biomaterials. 2009;30(30):5910–5917. .

Ott H.C, et al. Perfusion-decellularized matrix: using nature’s platform to engineer a bioartificial heart. Nature Medicine. 2008;14(2):213–221.

Owens C.M, et al. Biofabrication and testing of a fully cellular nerve graft. Biofabrication. 2013;5(4):45007.

Ozbolat I.T. Bioprinting scale-up tissue and organ constructs for transplantation. Trends in Biotechnology. 2015;33(7):395–400 Available at:. http://www.ncbi.nlm.nih.gov/pubmed/25978871.

Ozbolat I.T. Scaffold-based or scaffold-free bioprinting: competing or complementing approaches? Journal of Nanotechnology in Engineering and Medicine. 2015;6(2):24701.

Ozbolat I.T, Chen H, Yu Y. Development of “Multi-arm Bioprinter” for hybrid biofabrication of tissue engineering constructs. Robotics and Computer-Integrated Manufacturing. 2014;30(3):295–304.

Ozbolat I.T, Hospodiuk M. Current advances and future perspectives in extrusion-based bioprinting. Biomaterials. 2016;76:321–343.

Ozbolat I.T, Yu Y. Bioprinting toward organ fabrication: challenges and future trends. IEEE Transactions on Bio-Medical Engineering. 2013;60(3):691–699.

Pagliuca F.W, et al. Generation of functional human pancreatic β cells in vitro. Cell. 2015;159(2):428–439.

Pagliuca F.W, Melton D.A. How to make a functional β-cell. Development. 2013;140(12):2472–2483.

Palakkan A.A, et al. Liver tissue engineering and cell sources: issues and challenges. Liver International. 2013;33(5):666–676.

Park J.Y, et al. A comparative study on collagen type I and hyaluronic acid dependent cell behavior for osteochondral tissue bioprinting. Biofabrication. 2014;6(3):35004.

Pati F, et al. Printing three-dimensional tissue analogues with decellularized extracellular matrix bioink. Nature Communications. 2014;5:3935.

Peng W, Unutmaz D, Ozbolat I.T. Bioprinting towards physiologically relevant tissue models for pharmaceutics. Trends in Biotechnology. September 2016;34(9):722–732. doi: 10.1016/j.tibtech.2016.05.013.

Perkins J.D. Are we reporting the same thing?: comments. Liver Transplantation. 2007;13:465–466.

Phillippi J.A, et al. Microenvironments engineered by inkjet bioprinting spatially direct adult stem cells toward muscle- and bone-like subpopulations. Stem Cells. 2008;26(1):127–134.

Proks P, Ashcroft F.M. Effects of divalent cations on exocytosis and endocytosis from single mouse pancreatic beta-cells. The Journal of Physiology. 1995;487(2):465–477.

Quint C, et al. Decellularized tissue-engineered blood vessel as an arterial conduit. Proceedings of the National Academy of Sciences. 2011;108(22):9214–9219.

Raikwar S.P, et al. Human iPS cell-derived insulin producing cells form vascularized organoids under the kidney capsules of diabetic mice. PLoS ONE. 2015;10(1):e0116582.

Rodríguez-Dévora J.I, et al. High throughput miniature drug-screening platform using bioprinting technology. Biofabrication. 2012;4(3):35001.

Roskos K, et al. Essentials of 3D Biofabrication and Translation. Elsevier; 2015.

Saunders R, Derby B. Inkjet printing biomaterials for tissue engineering: bioprinting. International Materials Review. 2014;59(8):430–448.

Shim J.-H, et al. Bioprinting of a mechanically enhanced three-dimensional dual cell-laden construct for osteochondral tissue engineering using a multi-head tissue/organ building system. Journal of Micromechanics and Microengineering. 2012;22(8):85014.

Skardal A, et al. Bioprinted amniotic fluid-derived stem cells accelerate healing of large skin wounds. Stem Cells Translational Medicine. 2012;1(11):792–802. .

Snyder J.E, et al. Bioprinting cell-laden matrigel for radioprotection study of liver by pro-drug conversion in a dual-tissue microfluidic chip. Biofabrication. 2011;3:034112.

Temple J.P, et al. Engineering anatomically shaped vascularized bone grafts with hASCs and 3D-printed PCL scaffolds. Journal of Biomedical Materials Research Part A. 2014;102(12):4317–4325.

Tourlomousis F, Chang R.C. Numerical investigation of dynamic microorgan devices as drug screening platforms. Part I: Macroscale modeling approach & validation. Biotechnology and Bioengineering. 2015;113(3):612–622.

Vaidya M. Startups tout commercially 3D-printed tissue for drug screening. Nature Medicine. 2015;21(1):2.

Wu W, DeConinck A, Lewis J.A. Omnidirectional printing of 3D Microvascular networks. Advanced Materials. 2011;23(24):H178–H183.

Xiong R, Zhang Z, Chai W, et al. Freeform drop-on-demand laser printing of 3D alginate and cellular constructs. Biofabrication. 2015;7(4):45011.

Xiong R, Zhang Z, Huang Y. Identification of optimal printing conditions for laser printing of alginate tubular constructs. Journal of Manufacturing Processes. 2015;20:450–455.

Xu F, Moon S, Emre A, Lien C, Turali E, Demicri U. Cell bioprinting as a potential high-throughput method for fabricating cell-based biosensors (CBBs). IEEE Sensors. 2009:387–391.

Xu C, et al. Scaffold-free inkjet printing of three-dimensional zigzag cellular tubes. Biotechnology and Bioengineering. 2012;109(12):3152–3160.

Xu F, et al. A three-dimensional in vitro ovarian cancer coculture model using a high-throughput cell patterning platform. Biotechnology Journal. 2011;6(2):204–212.

Xu T, et al. Fabrication and characterization of bio-engineered cardiac pseudo tissues. Biofabrication. 2009;1(3):35001.

Xu T, et al. Hybrid printing of mechanically and biologically improved constructs for cartilage tissue engineering applications. Biofabrication. 2013;5(1):15001.

Yanez M, et al. Vivo assessment of printed microvasculature in a bilayer skin graft to treat full-thickness wounds. Tissue Engineering Part A. 2014;21(1–2):224–233.

Yu Y, Ozbolat I.T. Tissue strands as “bioink” for scale-up organ printing. In: Conference Proceedings: Annual International Conference of the IEEE Engineering in Medicine and Biology Society. IEEE Engineering in Medicine and Biology Society. Annual Conference. 2014:1428–1431.

Yoon No D, et al. 3D liver models on a microplatform: well-defined culture, engineering of liver tissue and liver-on-a-chip. Lab on a Chip. 2015;15(19):3822–3837.

Yoshida Y, Yamanaka S. Recent stem cell advances: induced pluripotent stem cells for disease modeling and stem cell–based regeneration. Circulation. 2010;122(1):80–87.

Yu Y, et al. Evaluation of cell viability and functionality in vessel-like bioprintable cell-laden tubular channels. Journal of Biomechanical Engineering. 2013;135(9):91011.

Yu Y, et al. Three-dimensional bioprinting using self-assembling scalable scaffold-free “tissue strands” as a new bioink. Scientific Reports. 2016;6:28714.

Yu Y, Zhang Y, Ozbolat I.T. A hybrid bioprinting approach for scale-up tissue fabrication. Journal of Manufacturing Science and Engineering. 2014;136(6):61013.

Zervantonakis I.K, et al. Three-dimensional microfluidic model for tumor cell intravasation and endothelial barrier function. Proceedings of the National Academy of Sciences. 2012;109(34):13515–13520.

Zhang Y, Yu Y, Chen H, et al. Characterization of printable cellular micro-fluidic channels for tissue engineering. Biofabrication. 2013;5(2):025004. .

Zhang Y, et al. In vitro study of directly bioprinted perfusable vasculature conduits. Biomaterials Science. 2015;3(1):134–143.

Zhang Y, Yu Y, Ozbolat I.T. Direct bioprinting of vessel-like tubular microfluidic channels. Journal of Nanotechnology in Engineering and Medicine. 2013;4:020902.

Zhao L, et al. The integration of 3-D cell printing and mesoscopic fluorescence molecular tomography of vascular constructs within thick hydrogel scaffolds. Biomaterials. 2012;33(21):5325–5332.

Zhao Y, et al. Three-dimensional printing of Hela cells for cervical tumor model in vitro. Biofabrication. 2014;6(3):35001.

Zopf D.A, et al. Bioresorbable airway splint created with a three-dimensional printer. New England Journal of Medicine. 2013;368(21):2043–2045.


 With minor contributions by Dr. Weijie Peng, The Pennsylvania State University.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.118.189.251