Next Article in Journal
Experimental Proof of a Transformation Product Trap Effect with a Membrane Photocatalytic Process for VOC Removal
Previous Article in Journal
Bio-Based Ceramic Membranes for Bacteria Removal from Water
Previous Article in Special Issue
The Use of Biocompatible Membranes in Oral Surgery: The Past, Present & Future Directions. A Narrative Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

3D Printed and Bioprinted Membranes and Scaffolds for the Periodontal Tissue Regeneration: A Narrative Review

by
Irina-Georgeta Sufaru
1,
Georgiana Macovei
2,*,
Simona Stoleriu
3,*,
Maria-Alexandra Martu
1,
Ionut Luchian
1,
Diana-Cristala Kappenberg-Nitescu
1 and
Sorina Mihaela Solomon
1
1
Department of Periodontology, Grigore T. Popa University of Medicine and Pharmacy, Universitatii Street 16, 700115 Iasi, Romania
2
Department of Oral and Dental Diagnostics, Grigore T. Popa University of Medicine and Pharmacy, Universitatii Street 16, 700115 Iasi, Romania
3
Department of Cariology and Restorative Dental Therapy, Grigore T. Popa University of Medicine and Pharmacy, Universitatii Street 16, 700115 Iasi, Romania
*
Authors to whom correspondence should be addressed.
Membranes 2022, 12(9), 902; https://doi.org/10.3390/membranes12090902
Submission received: 12 August 2022 / Revised: 13 September 2022 / Accepted: 15 September 2022 / Published: 19 September 2022
(This article belongs to the Special Issue Biocompatible Membranes for Bone Regeneration)

Abstract

:
Numerous technologies and materials were developed with the aim of repairing and reconstructing the tissue loss in patients with periodontitis. Periodontal guided bone regeneration (GBR) and guided tissue regeneration (GTR) involves the use of a membrane which prevents epithelial cell migration, and helps to maintain the space, creating a protected area in which tissue regeneration is favored. Over the time, manufacturing procedures of such barrier membranes followed important improvements. Three-dimensional (3D) printing technology has led to major innovations in periodontal regeneration methods, using technologies such as inkjet printing, light-assisted 3D printing or micro-extrusion. Besides the 3D printing of monophasic and multi-phasic scaffolds, bioprinting and tissue engineering have emerged as innovative technologies which can change the way we see GTR and GBR.

1. Introduction

Periodontitis is a periodontal tissues inflammatory disease, of multifactorial etiology [1,2,3,4], in which the host’s immune response to the aggression of periodontopathogenic bacteria plays a key role [5]. This pathology is characterized by the progressive loss of periodontal attachment, destruction of the alveolar bone, phenomena that, over time, can lead to tooth loss.
Periodontal treatment includes the elimination or modulation of the factors which led to the appearance and evolution of periodontitis, as well as the correction, within the limits of the case and the available technologies, of periodontal soft and hard tissues loss [6,7]. The reconstruction of these defects is essential for the restoration of masticatory, aesthetic, phonation functions and for improving the patients’ quality of life. Therefore, the purpose of complete periodontal therapy is to regenerate the entire periodontal system, which includes bone and cement neo-formation, as well as restoring the attachment of periodontal fibers [8].
The healing following non-surgical periodontal therapy can only provide periodontal repair, represented by the formation of a long epithelial attachment by migrating epithelial cells [9,10]. Therefore, a complete restoration of periodontal anatomy and functionality cannot be ensured [11]. These aspects can be covered by tissue regeneration, with the mobilization and involvement of cellular elements—fibroblasts, osteoblasts and cementoblasts, as well as the signals needed to direct regenerative processes [12].
Guided tissue and bone regeneration (GTR/GBR) are periodontal surgical interventions whose main purpose is to restore the periodontal architecture [13,14]. Primarily, these procedures involve the use of a barrier membrane that may or may not be associated with bone regeneration materials and whose primary function is to prevent epithelial cell migration into the bone defect [15,16]. Moreover, the barrier membrane also serves to maintain the space, creating a protected area in which tissue regeneration is favored. There are four principles considered necessary for successful regeneration, included in the acronym PASS: (P) primary closure of the wound to ensure optimal healing; (A) angiogenesis for adequate vascularization of newly formed tissues, with the supply of oxygen, nutrients, and pro-healing cells; (S) maintaining space for bone neo-formation, preventing inadequate epithelial cell proliferation; (S) wound stability to include blood clot formation [17].
In addition to strength and stability, a number of other characteristics of the barrier membrane can influence the regenerative success. These include pore size, permeability or its architecture [18]. The pore size of barrier membranes has long been a controversial issue. Membrane pore size can influence cell adhesion as well as progenitor cell migration [19]; these processes facilitate not only clot formation but also membrane stabilization, preventing micro-displacements that could disrupt tissue neoformation [20]. Studies have shown that excessively large pores make membranes less effective against soft tissue cells migration and proliferation [21]. To date, an optimal size of barrier membrane porosity has not been confirmed [22].
Over time, there has been a continuous search for ideal materials for tissue regeneration and numerous methods have been developed to stimulate the periodontal tissues repair [23]. The membranes can be classified, depending on their biodegradation capacity, into resorbable membranes (membranes made of synthetic or natural polymers) and non-resorbable membranes (metallic membranes; membranes made of synthetic polymers, polytetrafluoroethylene—PTFE) [24]. Furthermore, another classification includes first generation (non-resorbable), second generation (resorbable) and third generation (membranes as a product of tissue engineering) membranes [25].
GTR and GBR techniques have recently benefited from remarkable advances in the field of three-dimensional (3D) printing, tissue engineering and biofabrication [16,26,27], creating extensive conditions and possibilities for regenerative therapies.
Moreover, 3D printing technology has led to major innovations in periodontal regeneration methods, allowing the printing not only of biocompatible membranes and scaffolds, but also of living cells and supporting components in complex 3D functional tissues, defined as “bioprinting” [28]. The concept of periodontal tissue engineering has thus emerged, being defined as ”the use of physical, chemical, biological, and engineering processes to control and direct the aggregate behavior of cells” [29].
Bioprinting technology is a state-of-the-art tool for rendering biofunctional hierarchical architecture through 3D printing, with one or more types of living cells embedded. Bioprinting refers to the printing of components that form a specific tissue, including living cells embedded in matrix materials, to generate analogous tissue structures [29]. Bioprinting techniques bring a biological functionality to a conventional 3D printed scaffold, as it mimics a cell-to-cell and cell-to-matrix interaction in construction [30]. The material used in 3D bioprinting involves the usage of living biological cells, hydrogels, chemical factors and biomolecules, under the name of ”bioink”. A difference must be made regarding the types of printing material, between ”bioinks” and ”biomaterial inks”, as those two terms do not assimilate each other. While a bioink is cell-laden, acting as a cell carrier and deliverer during formulation and bioprinting processing, biomaterial inks are cell-free and can only be seeded with cells after printing [31].
The aim of this paper is to review the present data in the literature, related to 3D printing and bioprinting techniques and materials, as well as applications in the form of membranes and scaffolds in the regeneration of the periodontal apparatus.

2. 3D Printing and Bioprinting Techniques

3D printing is a revolutionary technology also known as additive manufacturing [29]. Figure 1 shows a series of milestones in the chronological evolution of 3D printing and bioprinting technologies. In 1984, Charles Hull invented stereolithography (SLA) for printing 3D objects from digital data [32]. In 1986 Carl Deckard and Joe Beaman invented selective laser sintering (SLS) [33].
Bioprinting was first demonstrated in 1988 by Klebe with a standard Hewlett-Packard (HP) inkjet printer to deposit cells by cytowriting technology [34]. Moreover, in 1988, the first SLA 3D printer was available [35]. In 1989, the fused deposition modeling (FDM) technique was developed by Scott Crump [36]. In 1999, the first organ used for transplantation (bladder) was printed [37].
In the year 2000, the first extrusion-based 3D printer (3D-Bioplotter) appeared, and two years after that, the first kidney was printed by micro-extrusion [38]. In 2003 Wilson and Boland developed the first inkjet bioprinter by modifying a standard HP inkjet printer [39]. In 2005, the RepRap (Replicating Rapid prototyper) project was founded at the University of Bath, Somerset, UK, with the aim of creating a low-cost 3D printer that would print most of its components [40]. The first SLS-based 3D printer was available in 2007 [41].
The first prosthetic limb was printed in 2008 [42] and the first blood vessels were 3D printed in 2009 [43]. The first 3D printed jaw was made in 2012 [44]. The year 2014 marks the first 3D printing of liver tissue [45] but also the appearance of the first desktop 3D printer [46]. Rasperini et al. developed and applied the first 3D-printed scaffold for periodontal repair in 2015 [26]. In 2018, the first commercial 3D-printed full human (skin) product became available [47] and in 2019, Noor et al. succeeded in generating a perfusable heart with reduced dimensions [48]. The 2020s pave the way for the personalized use of 3D printing in medicine [49].
3D printing technology has found numerous applications in dentistry and periodontics, in particular. 3D printing methods have been investigated in the treatment of gingival lesions (treatment of gingival recessions, gingivectomies or restoration of the smile design), as well as regeneration of periodontal tissues: alveolar bone, periodontal ligaments (PDL) and cement [50,51]. Most commonly used biomaterials in periodontal tissue regeneration 3D assisted are presented in Table 1, along with their main advantages and disadvantages.
The main 3D printing techniques in GTR and GBR include droplet-based printing, micro-extrusion and light-assisted printing [28].

2.1. Droplet-Based Printing

Droplet-based 3D printed products are the result of independent and discrete droplets as basic unit. These techniques can be divided inro inkjet 3D printing and electrohydrodynamic jetting. Moreover, inkjet printing can be performed either by continuous inkjet printing or drop-on-demand printing [43].
Inkjet printers used to be the most frequently approached method in both non-biological and biological applications [28]. The principle of operation involves the controlled passage of a low-viscosity solution flow, an absolutely necessary condition due to the excessive force required to release droplets when using solutions at higher viscosities [52]. The fluid subsequently breaks into very small-volume droplets (1–100 pL) after passing through the holes in the printhead; these droplets are then distributed under the influence of pressure pulses in predefined locations, to form three-dimensional structures after solidification [53].
The development of inkjet printers in biological printing is based on classic, two-dimensional inkjet printers [28]; in the modified version, the ink, the cartridge and the paper are replaced with biocompatible materials and the control is performed three-dimensionally, in the three axes (OX, OY and OZ) [54]. Subsequently, printheads with several hundred individual nozzles were developed [11].
Continuous inkjet printing is based on a natural phenomenon called Rayleigh-plateau instability, which exhibits the natural tendency for a stream of liquid to undergo a morphological transformation to a train of discrete drops (Figure 2). On the other hand, drop-on-demand printer produces a droplet when required; droplet deposition is performed by displacing the nozzle above the desired location before a droplet is ejected [43]. Modulation and conditions of drop-on-demand printing can also be done thermally, acoustically or electromagnetically (Figure 2).
A number of major disadvantages, however, make the use of inkjet printers limited. In the case of thermal inkjet printers there is an increased risk of thermal and mechanical stress; moreover, low directionality and uneven droplet size, as well as frequent nozzle clogging, have been reported [55]. The disadvantages encountered in the case of thermal inkjet printers have tried to be overcome by introducing acoustic inkjet printers or electromagnetic inkjet printers [56]. The concern in the case of electromagnetic inkjet printers is related to the used frequencies (15–25 kHz) which can induce damage to the cell membrane [57].
Moreover, a high pressure is required when using inkjet printers, an aspect which can affect the cell viability within the bioink. In order to counteract this particular disadvantage, electrohydrodynamic jetting avoids such pressures by using an electric field (Figure 2) [58]. Furthermore, electrohydrodynamic jetting is particularly suitable for printing bioink with high cell concentration and weight/volume ratio.
The metallic nozzle is filled with bioink and a spherical meniscus is formed at the tip of the nozzle due to surface tension. A high voltage is applied between nozzle and substrate and droplets are ejected when the electrostatic stresses overcome the surface tension under a sufficiently high voltage [59]. Different voltages will generate different types of inkjets (micro-dripping, intermittent jetting, breakdown, etc.), but the most frequently used are independent, discrete droplets [43]. Also, higher voltages will produce smaller droplets [60].

2.2. Light-Assisted 3D Printing

Light-assisted 3D printing uses laser-assisted, photocuring-based or stereolithography techniques in order to generate printed and bioprinted products.
Laser-assisted printing involves nozzle-free and non-contacting techniques, which include laser-induced forward transfer (LIFT), laser-induced forward transfer supported by an absorption film (AFA-LIFT), matrix-assisted laser evaporation direct writing (MAPLE-DW), biological processing (BioLP) or laser guidance direct writing (LGDW).
Generally, a laser-assisted printer is composed of: a pulsed laser source, usually a nanosecond laser with ultraviolet wavelength; optics necessary for the beam delivery; a target in the form of a glass/quartz ribbon coated with the bioink; and a receiving substrate coated with biopolymer or cell medium (Figure 3) [31]. The laser beam is directed at the absorbing layer of the ribbon, generating local evaporation and the formation of high-pressure bubbles, propelling the cell-containing material towards the receiving substrate [43]. Laser parameters need to be accurately set, in order to control the usual photothermal, photochemical and photomechanical laser effects.
Laser-assisted techniques allow bioprinting with biomaterials of high cell density and viscosity, at a high resolution. Being nozzle-free instruments, they also avoid clogging or high shear stress problems generated by other techniques [61]. Their main disadvantages include the high cost and the complex printing technique.
Stereolithography (SLA) was the first commercially available 3D printing technology. The basic principle of SLA is given by the photopolymerization of a highly crosslinked viscous polymer or prepolymer under exposure to a light energy source (laser or UV), to create layered structures [62].
The SLA printing process involves three main stages: exposure to the light radiation source, platform movement, and polymer filling (Figure 4) [63]. The polymerization layers are bonded from the bottom up, following the exposure of the resin to light radiation; as one layer is polymerized, the platform descends for a distance equal to the thickness of one layer and builds the next layer until the printing of the digitized 3D object is completed [63,64]. In order to be used in regenerative therapy, SLA involves the use of a slowly photopolymerized prepolymer formulation, loaded with cells [64].
Digital projection printing (DLP) is similar to SLA procedure, with the difference that, while in the case of SLA the light beam moves, in the case of DLP it is stationary, which makes DLP a process with a higher printing speed, but less accurate [63]. DLP uses liquid photosensitive resins which, under the action of light curing, will form the 3D construction, layer by layer (Figure 5). DLP printers use a system of micro-mirrors that conduct light to the projection lens [65].
Selective laser sintering (SLS) uses a high energy laser beam to induce the fusion of the raw material in powder form, layer by layer. SLS does not require additional material support during printing because the support is provided by the surrounding powder (Figure 6) [66]. This particular technique uses a high energy laser which selectively fuses the powdered material by scanning cross-sections generated from the digital file.

2.3. Extrusion 3D Printing

Of the various 3D bioprinting technologies, extrusion is the most commonly used in medical applications [57]. Extrusion 3D printing can be performed under thermal and non-thermal conditions.

2.3.1. Thermal Extrusion 3D Printing

Thermal extrusion 3D printing can be done by fused deposition modelling (FDM) or melt electrowriting (MEW) (Figure 7). Fused deposition modelling (FDM) is based on layer-by-layer printing of thermoplastic polymers. The polymer is heated over its melting point and, due to a solid-semisolid transition, is then extruded into filaments with diameters of 160–700 μm, through a nozzle, according to the predefined design [16]. FDM is characterized by high printing speed; in addition, by adding multiple nozzle heads, various materials can be printed simultaneously [67]. Due to the fact that FDM uses high temperatures (140–250 °C), it cannot be used in bioprinting; this technique can be approached, however, for the realization of multiphase scaffolds of periodontal regeneration [68,69,70].
The main disadvantages of FDM include the low resolution and the surface finish that may be poor, following the spread of the material before it has cooled; it should be noted, however, that these drawbacks can be overcome by decreasing the nozzle diameter [71].
Melt electrowriting combines FDM with the electrospinning technique to obtain extremely thin filaments (generally 2–30 μm), even nanometric [72], directed in porous constructions with high degree of complexity and precision. MEW involves applying an electric field to continuously pull a molten polymer toward a computer-controlled static or rotating collector plate [73]. The essential parameters in the realization of such constructions are represented by the delivery rate of the polymer, the needle diameter, the collector speed, the distance from the needle tip to the collector, as well as the applied electrical voltage [16,74]; by adjusting these variables, fibers can be obtained on a scale very close to native collagen fibers [75]. Thus, MEW is characterized by a remarkable resolution of the fibers, with the ability to print 1 cm thick structures, which makes this technique ideal for scaffolding or cell bioprinting. To date, most research has focused on prints in the same plane as the substrate/construction plate; it has been shown, however, that constructions in anatomical forms can be obtained from MEW using biomaterials such as hydrogel or bioceramics [76].

2.3.2. Non-Thermal Extrusion 3D Printing

Non-thermal extrusion techniques allow 3D printing by extrusion without melting the material. This printing system can use air pressure, a pressurized piston without a valve or a screw to extrude cell-free or cell-loaded biomaterials on a substrate, avoiding the distortion of temperature-sensitive biomolecules and/or cell death (Figure 8) [16].
Extremely important parameters in this technique are represented by the properties of the material, the quantity and quality of the used additives. Non-thermal extrusion has been used mainly for making cell-loaded scaffolds at physiological temperatures [77].
Although significant progress has been reported in extrusion-based 3D bioprinting, further improvements in bioink availability are needed; hydrogels with suitable printing properties have been developed, which can ensure a maintenance of the post-printing 3D shape without compromising the viability of the obtained product [78].

3. Applications of 3D Printing in Periodontology

3.1. 3D Printed Scaffolds in Periodontal Defects

Monophasic scaffolds are characterized by the presence of a single compartment, having the characteristic functions of a barrier membrane: maintaining and stabilizing the bone defect subject to regeneration; ensuring bone proliferation, without epithelial interference in bone defect; control of the healing process in time and space [79]. Moreover, these scaffolds can be loaded with growth factors or cells to promote bone neo-formation [11]. Obtaining alveolar bone tissue, facilitated by monophasic scaffolds, is not sufficient for restitutio ad integrum. Thus, biphasic scaffolds have been developed for the periodontal ligaments’ regeneration. Triphasic scaffolds have emerged in order to promote the regeneration of cementum, together with the alveolar bone and periodontal ligaments. A schematic structure of printed scaffolds, along with their main characteristics, are presented in Figure 9.
Carrel et al. [80] developed an extrusion scaffold, consisting of cylindrical filaments of tricalcium beta-phosphate (β-TCP) and hydroxyapatite (HA), arranged in orthogonal layers (OsteoFlux®). It was placed in sheep calvary defects and was compared with bovine bone (Bio-Oss®) and β-TCP particles. In the first 8 weeks the authors observed a significant increase in bone growth, but no difference was registered in the total of four months [80].
Another scaffold, consisting of 30% HA, 60% β-TCP and 10% tricalcium alpha-phosphate (α-TCP), also made by extrusion, in the form of a mesh, with a macroporosity of 60%, was implanted in sheep sinus [81]. The scaffold was well tolerated, generating a peripheral bone remodeling in the first 45 days after implantation; at 90 days, the formation of a peripheral lamellar bone was observed, but after 90 days the scaffold continued to resorb, leaving an incompletely filled bone defect, with areas of fibrous tissue [81].
Cellular loading of hydrogel scaffolds was also attempted, mainly by electrospinning techniques; these forms of scaffolds have, however, important limitations related to cell source and culture [11,82,83]. Improved single-phase scaffolds with growth factors have also been developed. Cho et al. [84] used extrusion to make a polycaprolactone (PCL) scaffold, loaded with poly (lactic-co-glycolic acid) (PLGA) microspheres with morphogenetic protein 2 and 7 (BMP-2, BMP-7) and connective tissue growth factor (CTGF); the scaffold was implanted in vitro on the root surface of human teeth with the removed cementum, in a cementogenic/osteogenic environment. After 6 weeks, all groups delivered with growth factor showed surface recovery of the dentin with a layer similar to the newly formed cement, compared to the control. BMP-2 and BMP-7 showed de novo formation of significantly thicker tissue layers than all other groups, while CTGF and BMP-7 resulted in significantly improved integration on the dentin surface [84].
Shim et al. [85] compared 3D printed polycaprolactone (PCL) and tricalcium polycaprolactone/β-phosphate (PCL/β-TCP) membranes with a conventional commercial collagen membrane in terms of their ability to facilitate GBR, investigating the mechanical properties in dry and humid environment. Fibroblasts and pre-osteoblasts were seeded in membranes and proliferation rates and patterns were analyzed with scanning electron microscopy. Subsequently, the membranes were placed in alveolar defects in beagle dogs. CT and histological analyzes at eight weeks postoperatively showed that 3D-printed PCL/β-TCP membranes were more efficient than 3D-printed PCL and substantially better than conventional collagen membranes in terms of biocompatibility and bone regeneration [85].
Dubey et al. [86] designed, by infusing a PCL mesh made by MEW, with a hydrogel loaded with amorphous magnesium phosphate (AMP), a membrane reinforced with highly adjustable bioactive fibers for GBR. This membrane proved that the presence of PCL networks manufactured by MEW can delay the degradation of the hydrogel; thus, soft tissue invasion is prevented. At the same time, a mechanical barrier is generated to allow slower-migrating progenitor cells to participate in bone regeneration [86].
Hsieh et al. [87] investigated the biological characteristics of human PDL cell spheroids formed on chitosan and polyvinyl alcohol; PDL cell spheroids were cultured in 3D-printed polylactic acid scaffolds by FDM to assess mineralization capacity. Cellular spheroids formed on the chitosan membrane demonstrated an increased alkaline phosphatase activity, as well as an increase in mineralized matrix deposits [87].
Bai et al. [88] have developed an individualized titanium mesh (Ti-Mesh) using computer-aided design and sintering additive manufacturing technology to evaluate the effect of different thicknesses and sizes of titanium mesh pores on its mechanical properties. The authors observed that, as the mesh diameter increased (3 mm to 5 mm), the mechanical properties of the mesh decreased. The 0.4 mm thick titanium mesh proved to be strong enough with little mucosal stimulation.
Porous networks of 3D printed titanium-niobium alloy (Ti-Nb) have also been developed in order to maintain the space, to prevent the fibroblasts’ growth and inhibit the bacterial colonization [89]. In this technique, Ti-Nb alloy meshes were prepared by selective laser melting (SLM) and used as substrates for new surface coatings. Porous coatings of chitosan (CS)/gelatin (G)/doxycycline (Dox) were formed on the meshes, using electrophoretic deposition (EPD) and lyophilization. This membrane has been shown to be effective in preventing the growth of fibroblasts, while allowing nutrient infiltration; moreover, its antibacterial effect was observed.
A scaffold with two compartments was designed: bone and ligament, by combining extrusion with casting; the wax forms were made by extrusion, and the materials were then cast into these shapes [90]. The bone compartment was seeded with Ad-CMV-BMP7 transduced periodontal ligament cells and the ligament compartment, composed of three superimposed cylinders, was seeded with PDL cells. The authors observed that parallel and oblique oriented fibers were generated, which grew and traversed the designed PCL and poly (glycolic acid) (PGA) structures, forming ligaments and bone structures similar to the native ones; the disadvantage, however, is the inability to control and predict cell directionality in vivo [90].
Vaquette et al. designed a scaffold with a bone compartment made by β-TCP FDM, seeded with osteoblasts, and a ligament compartment produced by electrospinning, in which PDL cell sheets were placed [69]. Following the in vitro test, it was observed that osteoblasts formed a mineralized matrix in the bone compartment after 21 days of culture and that harvesting the PDL cell sheet did not induce significant cell death. The biphasic scaffold seeded with cells was placed on a block of dentin and implanted for 8 weeks in a subcutaneous model of athymic rat. The authors observed bone neoformation, ligament and cement regeneration (but with perpendicular non-orientated non-functional periodontal fibers) [69]. This model was taken over and modified by Costa et al. [91]; the bone compartment, with an increased pore size, was covered with a layer of calcium phosphate (CaP) to increase osteoconductivity, seeded with osteoblasts and cultured in vitro for 6 weeks. Then PDL cell sheets were placed in the obtained product, subsequently implanted subcutaneously in athymic rats for 8 weeks. These modifications led to better bone neo-formation, better oblique orientation of periodontal fibers (but poorly controlled) and increased vascularization [91].
Wang et al. [92] developed a biphasic scaffold model made of collagen and strontium-doped calcium silicate, manufactured by extrusion, loaded with gingival fibroblast cells; it was placed in calvary defects in rabbits. The authors found that biphasic scaffolds improved osteogenesis [92].
Lee et al. developed a triphasic scaffold with a precise architecture, enriched with biochemical gradients [70]. This scaffold was made up of three distinct, layered compartments, corresponding to the morphology of the periodontal complex: cementum, periodontal ligament and alveolar bone. Each component had a specific architecture, with variable pore sizes (100, 600 and 300 μm). Moreover, PGA microspheres loaded with growth factors specific to the regeneration of each tissue type (amelogenin, CTGF and BMP-2) were added to each compartment. The scaffold was digitally designed and 3D printed, but the addition of microspheres was done manually, by pipetting. The authors observed a discontinuous cementogenesis, with a notable osteogenesis in the bone compartment. Interposed connective tissue between these two mineralized formations was found, with an alignment of the fibers and a ligament attachment on the newly formed cementoid tissue [70].
Rasperini et al. [26] designed a customized scaffold, based on the patient’s computed tomography; the scaffold was made of PCL powder containing HA, using selective laser sintering. Although this scaffold proved effective in treating large periodontal defects, at 12 months it became exposed, which compromised the therapeutic result.
Therefore, triphasic scaffolds, with a three-compartment design, have an important potential in complete periodontal regeneration, which involves the formation of cement on the root surface, the formation and proper insertion of periodontal fibers in hard tissues and sufficient rigidity in general [93], but their design and implementation require further investigation.
One concern in making and functionalizing multiphase scaffolds was poor interphase cohesion, which affects the mechanical stability of the scaffold. Continuous additive fabrication or simultaneous multiphase crosslinking seem viable options to overcome this dilemma [94,95].

3.2. Socket Preservation

Bone resorption after dental extraction represents a common challenge and various techniques have been developed and investigated, 3D printing included, in order to preserve an adequate height of the alveolar bone. Park et al. [96] evaluated the efficiency of a 3D printed PCL scaffold implanted in artificially created saddle-type bone defects in beagle dogs, along with β-TCP. The authors observed that the PCL scaffold maintained the physical space, without any inflammatory infiltrates around the scaffold.
Goh et al. [97] inserted a 3D bioresorbable PCL scaffold in fresh extraction sockets in a randomized controlled clinical trial (RCT). The micro-CT and histological analysis revealed that this treatment option resulted in a lower vertical ridge resorption at the 6-months evaluation.
3D printing was used also for HA granules manufacturing, which were placed in alveolar sockets after atraumatic extractions and covered with collagen membrane [98]. At the 8-weeks evaluation, the grafted site was completely filled with woven bone, with vascular neo-formation and without any inflammatory signs.

3.3. Other Applications

Positive outcomes have been observed in bone augmentation for sinus lift procedures when using 3D applications [99]. Different materials have been tested, such as monolithic monetite (dicalcium phosphate anhydrous) or biphasic calcium phosphate [100]. The main advantages of using 3D printed products for sinus lift are represented by the low risk of infection transfer, as well as by shorter surgery time and higher comfort for the patient [101]. A synthesis of the main 3D applications in available studies is presented in Table 2.
Other applications include 3D-surgical guide precise implant placement, with both static and dynamic implantation guiding [111,112,113,114,115,116,117,118]. 3D printing has been also extensively used in education of dental school students and residents and even for patient motivation.

4. Bioprinting

The 3D bioprinting process comprises six major stages:
  • Data acquisition, using X-ray scanning and reconstruction techniques, computed tomography (CT), magnetic resonance imaging (MRI), or directly using computer-aided design (CAD) software. These data will be processed with the help of specific software. The file is converted to a printer-readable file [119]. The data is then translated to allow estimation of the amount of material to be extruded, which depends on the desired height and width of the layer according to the shape of the bioink (droplet or filament) [31,120].
  • The choice of bioink, which is made according to the printing technique and the requirements of the printed structures. Thus, the bioink must meet favorable mechanical properties, as well as biocompatibility and printability requirements. The bioink can contain isolated cells, growth factors and bioprinting materials. It is prepared according to the physiological temperature, pH and requirements of the printed structures [31].
  • Setting the appropriate printing parameters, depending on the bioink and the desired structure of the printed product.
  • The actual bioprinting, under close observation to make adjustments when necessary. Printing resolution is specific to the printer and the type of bioink. In cases of high resolution, the time to fabricate the object can be longer [121].
  • Post-printing stage, which can include spinning and microscopical assessment of the printed object. The bioprinted object is kept in an incubator or bioreactor.
  • Placement of the bioprinted product (in vivo or in vitro conditions).
3D bioprinting involves precise, layer-by-layer positioning of biological materials and living cells [57]. Some applications of 3D bioprinting include stem cell research [122], anti-cancer therapy [123], drug testing [124] and tissue engineering [125]. In tissue engineering, these technologies can control pore size, shape, distribution and interconnectivity. Moreover, by associating bioprinting technologies with imaging tests, such as cone-beam computed tomography (CBCT), specialized constructions can be made, adapted to the case and site-specific [126].
Bioprinting is based on three fundamental concepts: biomimicry, self-assembly and building blocks with mini-tissue [127]. Biomimicry involves the creation of exact replicas of the cellular and extracellular parts of a tissue and organ [128]. So far, bioprinting has been adopted for the manufacturing of bioartificial tissues and organs, such as skin, bones, cartilage, kidneys, heart, lungs [53,103,129,130,131,132,133]. Self-assembly involves the design of a scaffold-free method that mimics the behavior of embryonic stem cells and mini-tissues can be defined as the smallest structural and functional component of a tissue [130].
Although this concept is still in its infancy, a number of in vitro and in vivo studies have shown interesting perspectives that go beyond the principles of barrier membranes and scaffolds. At the same time, there are numerous limitations to overcome, which include production technology and bioinks optimization [134]. Photo-crosslinkable hydrogels, gelatin-methacryloyl hydrogel and poly (ethylene glycol) dimethacrylate have been proposed as basic materials in bioprinting, and the optimization of printing capacity, mechanical stability and cytocompatibility has been tried by testing different extrusion parameters and crosslinking methods [134].
Wang et al. [135] proposed an active tissue engineering scaffold for the repair of bone defects. This material was constructed with a 3D-BG scaffold composite, MSC and nanoparticles loaded with the BMP-2 gene (BMP/CS). The scaffold was subsequently implanted in the alveolar bone defect of rhesus monkeys and its capacity for osteogenesis was assessed. The authors found that this scaffold model promoted bone healing, with enhanced osteogenic properties in vivo [135].
Lin et al. [136] developed a biomimetic microfiber system, able to withstand functional load, to help regenerate PDL. Straight, collagen-based waveform microfibers to guide the growth of PDL cells were prepared by extrusion and a laminar flow-based bioreactor was used to generate fluid shear stress. PDL cells were seeded on those microfibers [136]. The authors found that microfibers maintained the viability of PDL cells and showed an improved tendency to promote healing and regeneration under shear stress.
Vurat et al. [137] developed a multicellular micro-tissue model similar to the periodontal ligament-alveolar bone biointerface. The periodontal ligament layer was modeled using methacryloyl gelatin bioink (Gel-MA) and ligament fibroblasts. The alveolar bone layer was modeled using a composite bioink composed of Gel-MA and HA-magnetic iron oxide nanoparticles (Gel-MA/HAp-MNPs), and osteoblasts [137]. The bioprinted self-supporting microfabric was cultured under flow in a microfluidic platform for more than 10 days, without significant loss of shape fidelity. Confocal microscopy analysis indicated that the encapsulated cells were homogeneously distributed inside the matrix and maintained their viability for more than 7 days under microfluidic conditions. Preliminary interaction study with tetracycline indicated absorption of the drug by cells inside the 3D microfabric [137].
Li et al. [138] developed a system for vascularized bone regeneration; they combined ethosomes loaded with deferoxamine (DFO) (Eth) with Gel-MA/gellan gum methacrylate (GGMA) to fabricate 3D printed scaffolds by photo and ionic crosslinking. This system had good cytocompatibility, generated sustained release of DFO, which significantly promoted endothelial cell migration and tube formation, mineralized matrix deposition, and alkaline phosphatase expression [138]. The scaffold was later placed in a rat cranial defect; the signaling pathway of inducible hypoxia factor 1-α (HIF1-α) has been activated, indicating that the composite scaffold could promote angiogenesis and bone regeneration [138].
Although 3D bioprinting has been identified as a technology capable of changing the paradigms of regenerative and reconstructive periodontal therapy, it still finds itself at the beginning of the road. Of course, further studies to translate in vitro models to experimental animals and, subsequently, why not, to human subjects, are an absolute necessity for the establishment of manufacturing and implementation protocols.

5. Future Directions

Periodontal regeneration involves a high degree of intricacy due to the complex nature and architecture of the periodontal apparatus. The coordination of the components, gingiva, periodontal ligaments, alveolar bone and root cementum, is of crucial importance in the complete regeneration of periodontal structures. In addition, the periodontal system is susceptible to microbial flora, as well as to local and systemic risk factors for periodontal disease. Translating in vitro and in vivo models into clinical models on human subjects is the main challenge associated with the development of technologies. Moreover, the present studies pay little attention to antimicrobial and immunomodulatory strategies. The integration of antimicrobial drugs and/or biomaterials with immunomodulatory capacity is an extremely interesting future direction.
An interesting approach that is still in its early stages is the reconstruction of highly organized living tissues and tissue interfaces in complementary converged 3D printing/bio-printing technologies, integrated into a single biofabrication platform, called multi-technology manufacturing. This field can generate scaffold systems with high print fidelity, spatial control over cell distribution and improved biomechanical properties, without compromising cell viability and functionality [139].
Studies already present in the literature have investigated the association of hydrogel printing with MEW [140]. The authors combined extrusion-based bioengineering and MEW in a single biofabrication platform, which allowed the fabrication of living constructs with spatial distribution of mesenchymal stromal cells, with improved functionality and biomechanics, without compromising cell viability or chondrogenic differentiation [140]. Diloksumpan et al. [141] also used convergent technologies to achieve hard-soft tissue interfaces, using hydrogel and bioceramics.
Dos Santos et al. [142] combined co-axial electrospinning and 3D printing techniques to generate zein-based bilayers as a potential platform for dual delivery of periodontal tissue regeneration drugs. In vitro experiments showed that the two-layer constructs provided sustained release of distinct drugs for 8 days and showed biocompatibility against human oral keratinocytes (Nok-si) (cell viability > 80%), as well as antibacterial activity against bacterial strains. distinct, including those from the red complex, such as Porphyromonas gingivalis and Treponema denticola [142].
Liu et al. [143] associated directly deposition 3D printing of PCL/gel/nano-hydroxyapatite (n-HA) scaffolds with polycaprolactone/gelatin nanofiber membranes (PCL/Gel) by electrospinning. The authors noted that after 20 weeks, the PCL/Gel/n-HA scaffold-treated sites showed a higher degree of new bone formation than in the control group, indicating that this is a favorable combination of GBR materials [143].
The fourth dimension was introduced in the bioprinting field; thus, smart scaffolds have emerged, programmed to undergo shape or functional changes according to the desired stimulation in time [144]. 3D-printed networks are stimulated by the environmental factor to fold into tubes, mimicking vascular-like tissue constructs [31].
Other promising technologies include external magnetic field stimulation [145] or atmospheric plasma functionalization [146]. Also, new perspectives are brought by real-time monitoring of the printing process through combinations of artificial vision, inspection sensors and feedback control systems supervised by high-efficiency artificial intelligence algorithms and the biofabrication of truly functional tissue equivalents [16,139].

6. Conclusions

In response to the requirements of periodontal GTR and GBR, new feeding materials and 3D printing methods have been developed, with the aim of accelerating production and improving performance, with favorable in vitro and even in animal models results. These aspects propel membrane printing and multiphasic scaffolds as a new and promising approach in the reconstruction of periodontal tissues.
Of course, many limitations related to the bioavailability of bioinks, the mechanical properties of the printed structure and its dimensional accuracy are still to be overcome. Moreover, translating the obtained models to human subjects remains an important challenge. Improving digitally assisted techniques and biomaterials, together with the combination of CBCT investigations, can facilitate this translation, with the production of patient- and site-specific scaffolds.
Tissue engineering is still in its early-stage development, due to the lack of in vivo and clinical evaluation in periodontal defect models, the available data being limited.

Author Contributions

Conceptualization, I.-G.S. and S.M.S.; methodology, G.M., I.L. and D.-C.K.-N.; resources, S.S. and M.-A.M.; writing—original draft preparation, I.-G.S., G.M. and S.S.; writing—review and editing, M.-A.M., I.L. and S.M.S.; visualization, I.-G.S.; supervision, S.M.S.; project administration, I.-G.S. and S.M.S.; funding acquisition, G.M. and S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. AlJehani, Y.A. Risk Factors of Periodontal Disease: Review of the Literature. Int. J. Dent. 2014, 2014, 182513. [Google Scholar] [CrossRef] [PubMed]
  2. Gibertoni, F.; Sommer, M.E.L.; Esquisatto, M.A.M.; Amaral, M.E.C.D.; Oliveira, C.A.; Andrade, T.A.M.; Mendonça, F.A.S.; Santamaria, M., Jr.; Felonato, M. Evolution of periodontal disease: Immune response and RANK/RANKL/OPG system. Braz. Dent. J. 2017, 28, 679–687. [Google Scholar] [CrossRef] [PubMed]
  3. Van Dyke, T.E.; Bartold, P.M.; Reynolds, E.C. The nexus between periodontal inflammation and dysbiosis. Front. Immunol. 2020, 11, 511. [Google Scholar] [CrossRef] [PubMed]
  4. He, L.; Liu, L.; Li, T.; Zhuang, D.; Dai, J.; Wang, B.; Bi, L. Exploring the imbalance of periodontitis immune system from the cellular to molecular level. Front. Genet. 2021, 12, 653209. [Google Scholar] [CrossRef]
  5. Song, J.; Zhao, H.; Pan, C.; Li, C.; Liu, J.; Pan, Y. Risk factors of chronic periodontitis on healing response: A multilevel modelling analysis. BMC Med. Inform. Decis. Mak. 2017, 17, 135. [Google Scholar] [CrossRef]
  6. Grzesik, W.J.; Narayanan, A.S. Cementum and periodontal wound healing and regeneration. Crit. Rev. Oral. Biol. Med. 2002, 13, 474–484. [Google Scholar] [CrossRef]
  7. Zhao, D.; Wu, M.Z.; Yu, S.Y.; Pelekos, G.; Yiu, K.H.; Jin, L. Periodontitis links to concurrent systemic comorbidities among ‘self-perceived health’ individuals. J. Periodontal. Res. 2022, 57, 632–643, Epub ahead of print. [Google Scholar] [CrossRef]
  8. Cirelli, J.A.; Fiorini, T.; Moreira, C.H.C.; de Molon, R.S.; Dutra, T.P.; Sallum, E.A. Periodontal regeneration: Is it still a goal in clinical periodontology? Braz. Oral Res. 2021, 35, 0097. [Google Scholar] [CrossRef]
  9. Cheng, X.; Yang, F. More than just a barrier-challenges in the development of guided bone regeneration membranes. Matter 2019, 1, 550–644. [Google Scholar] [CrossRef]
  10. Cho, Y.D.; Kim, K.H.; Lee, Y.M.; Ku, Y.; Seol, Y.J. Periodontal Wound Healing and Tissue Regeneration: A Narrative Review. Pharmaceuticals 2021, 14, 456. [Google Scholar] [CrossRef]
  11. Raveau, S.; Jordana, F. Tissue engineering and three-dimensional printing in periodontal regeneration: A literature review. J. Clin. Med. 2020, 9, 4008. [Google Scholar] [CrossRef] [PubMed]
  12. Han, J.; Menicanin, D.; Gronthos, S.; Bartold, P.M. Stem cells, tissue engineering and periodontal regeneration. Aust Dent. J. 2014, 59, 117–130. [Google Scholar] [CrossRef] [PubMed]
  13. Sanz, M.; Dahlin, C.; Apatzidou, D.; Artzi, Z.; Bozic, D.; Calciolari, E.; De Bruyn, H.; Dommisch, H.; Donos, N.; Eickholz, P.; et al. Biomaterials and regenerative technologies used in bone regeneration in the craniomaxillofacial region: Consensus report of group 2 of the 15th European Workshop on Periodontology on Bone Regeneration. J. Clin. Periodontol. 2019, 46, 82–91. [Google Scholar] [CrossRef] [PubMed]
  14. Dimitriou, R.; Mataliotakis, G.I.; Calori, G.M.; Giannoudis, P.V. The role of barrier membranes for guided bone regeneration and restoration of large bone defects: Current experimental and clinical evidence. BMC Med. 2012, 10, 81. [Google Scholar] [CrossRef] [PubMed]
  15. Bottino, M.C.; Pankajakshan, D.; Nör, J.E. Advanced scaffolds for dental pulp and periodontal regeneration. Dent. Clin. 2017, 61, 689–711. [Google Scholar] [CrossRef] [PubMed]
  16. Aytac, Z.; Dubey, N.; Daghrery, A.; Ferreira, J.; de Souza Araujo, I.J.; Castilho, M.; Malda, J.; Bottino, M.C. Innovations in craniofacial bone and periodontal tissue engineering—From electrospinning to converged biofabrication. Int. Mat. Rev. 2021, 2021, 1946236. [Google Scholar] [CrossRef]
  17. Wang, H.L.; Boyapati, L. “PASS” principles for predictable bone regeneration. Implant Dent. 2006, 15, 8–17. [Google Scholar] [CrossRef]
  18. Tayebi, L.; Rasoulianboroujeni, M.; Moharamzadeh, K.; Almela, T.K.D.; Cui, Z.; Ye, H. 3D-printed membrane for guided tissue regeneration. Mat. Sci. Eng. C Mater. Biol. Appl. 2018, C84, 148–158. [Google Scholar] [CrossRef]
  19. Murphy, C.M.; Haugh, M.G.; O’Brien, F.J. The effect of mean pore size on cell attachment, proliferation and migration in collagen–glycosaminoglycan scaffolds for bone tissue engineering. Biomaterials 2010, 31, 461–466. [Google Scholar] [CrossRef]
  20. Takata, T.; Wang, H.L.; Miyauchi, M. Attachment, proliferation and differentiation of periodontal ligament cells on various guided tissue regeneration membranes. J. Periodontal. Res. 2001, 36, 322–327. [Google Scholar] [CrossRef]
  21. Rakhmatia, Y.D.; Ayukawa, Y.; Furuhashi, A.; Koyano, K. Current barrier membranes: Titanium mesh and other membranes for guided bone regeneration in dental applications. J. Prosthodont. Res. 2013, 57, 3–14. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, H.Y.; Jiang, H.B.; Ryu, J.H.; Kang, H.; Kim, K.M.; Kwon, J.S. Comparing properties of variable pore-sized 3D-printed PLA membrane with conventional PLA membrane for guided bone/tissue regeneration. Materials 2019, 12, 1718. [Google Scholar] [CrossRef] [PubMed]
  23. Solomon, S.-M.; Sufaru, I.-G.; Teslaru, S.; Ghiciuc, C.M.; Stafie, C.S. Finding the perfect membrane: Current knowledge on barrier membranes in regenerative procedures: A descriptive review. Appl. Sci. 2022, 12, 1042. [Google Scholar] [CrossRef]
  24. Sasaki, J.I.; Abe, G.L.; Aonan, L.; Thongthai, P.; Tsuboi, R.; Kohno, T.; Imazato, S. Barrier membranes for tissue regeneration in dentistry. Biomat. Investig. Dent. 2021, 8, 54–63. [Google Scholar] [CrossRef]
  25. Lee, H.S.; Byun, S.H.; Cho, S.W.; Yang, B.E. Past, present, and future of regeneration therapy in oral and periodontal tissue: A review. Appl. Sci. 2019, 9, 1046. [Google Scholar] [CrossRef]
  26. Rasperini, G.; Pilipchuk, S.; Flanagan, C.L.; Park, C.H.; Pagni, G.; Hollister, S.J.; Giannobile, W.V. 3D—Printed bioresorbable scaffold for periodontal repair. J. Dent. Res. 2015, 94, 153S–157S. [Google Scholar] [CrossRef] [PubMed]
  27. Shen, C.; Witek, L.; Flores, R.L.; Tovar, N.; Torroni, A.; Coelho, P.G.; Kasper, F.K.; Wong, M.; Young, S. Three-dimensional printing for craniofacial bone tissue engineering. Tissue Eng. Part A 2020, 26, 1303–1311. [Google Scholar] [CrossRef] [PubMed]
  28. Ma, Y.; Xie, L.; Yang, B.; Tian, W. Three-dimensional printing biotechnology for the regeneration of the tooth and tooth-supporting tissues. Biotech Bioeng 2019, 116, 452–468. [Google Scholar] [CrossRef]
  29. Groll, J.; Boland, T.; Blunk, T.; Burdick, J.A.; Cho, D.W.; Dalton, P.D.; Derby, B.; Forgacs, G.; Li, Q.; Mironov, V.A.; et al. Biofabrication: Reappraising the definition of an evolving field. Biofabrication 2016, 8, 013001. [Google Scholar] [CrossRef]
  30. Yamada, S.; Shanbhag, S.; Mustafa, K. Scaffolds in periodontal regenerative treatment. Dent. Clin. N. Am. 2022, 66, 111–130. [Google Scholar] [CrossRef]
  31. Fatimi, A.; Okoro, O.V.; Podstawczyk, D.; Siminska-Stanny, J.; Shavandi, A. Natural hydrogel-based bio-inks for 3D bioprinting in tissue engineering: A review. Gels 2022, 8, 179. [Google Scholar] [CrossRef] [PubMed]
  32. US Patent for Apparatus for Production of Three-Dimensional Objects by Stereolithography Patent. U.S. Patent 4,575,330, 11 March 1986. Justia Patents Search. Available online: Patents.justia.com (accessed on 5 September 2022).
  33. Deckard, C. Method and apparatus for producing parts by selective sintering. U.S. Patent 4,863,538, 17 October 1986. published 5 September 1989. [Google Scholar]
  34. Klebe, R.J. Cytoscribing: A method for micropositioning cells and the construction of two- and three-dimensional synthetic tissues. Exp. Cell Res. 1988, 179, 362–373. [Google Scholar] [CrossRef]
  35. “Our Story”. 3D Systems. 3D Systems, Inc. Available online: https://www.3dsystems.com/our-story (accessed on 5 September 2022).
  36. Apparatus and Method for Creating Three-Dimensional Objects” (A System and a Method for Building Three-Dimensional Objects in a Layer-by-Layer Manner via Fused Deposition Modeling). U.S. Patent 5,121,329, 9 June 1989.
  37. Odde, D.J.; Renn, M.J. Laser-guided direct writing for applications in biotechnology. Trends Biotechnol. 1999, 17, 385–389. [Google Scholar] [CrossRef]
  38. Landers, R.; Hubner, U.; Schmelzeisen, R.; Mulhaupta, R. Rapid prototyping of scaffolds derived from thermoreversible hydrogels and tailored for applications in tissue engineering. Biomaterials 2002, 23, 4437–4447. [Google Scholar] [CrossRef]
  39. Wilson, W.C., Jr.; Boland, T. Cell and organ printing 1: Protein and cell printers. Anat. Rec. Part A 2003, 272, 491–496. [Google Scholar] [CrossRef]
  40. Jones, R.; Haufe, P.; Sells, E.; Iravani, P.; Olliver, V.; Palmer, C.; Bowyer, A. Reprap—The replicating rapid prototyper. Robotica 2011, 29, 177–191. [Google Scholar] [CrossRef]
  41. Charoo, N.A.; Ali, S.F.B.; Mohamed, E.M.; Kuttolamadom, M.A.; Ozkan, T.; Khan, M.A.; Rahman, Z. Selective laser sintering 3D printing—An overview of the technology and pharmaceutical applications. Drug Dev. Ind. Pharm. 2020, 46, 869–877. [Google Scholar] [CrossRef]
  42. The History of 3D Printing: 3D Printing Technologies from the 80s to Today. Available online: https://www.sculpteo.com/en/3d-learning-hub/basics-of-3d-printing/the-history-of-3d-printing/ (accessed on 5 September 2022).
  43. Gu, Z.; Fu, J.; Lin, H.; He, Y. Development of 3D bioprinting: From printing methods to biomedical applications. Asian J. Pharm. Sci. 2020, 15, 529–557. [Google Scholar] [CrossRef]
  44. Nickels, L. Worlds’s first patient-specific jaw implant. Metal Powder Rep. 2012, 67, 12–14. [Google Scholar] [CrossRef]
  45. Duan, B. State-of-the-art review of 3D bioprinting for cardiovascular tissue engineering. Ann. Biomed. Eng. 2017, 45, 195–209. [Google Scholar] [CrossRef]
  46. The Complete History of 3D Printing: From 1980 to 2022. Available online: https://www.3dsourced.com/guides/history-of-3d-printing/ (accessed on 4 September 2022).
  47. Beheshtizadeh, N.; Lotfibakhshaiesh, N.; Pazhouhnia, Z.; Hoseinpour, M.; Nafari, M. A review of 3D bio-printing for bone and skin tissue engineering: A commercial approach. J. Mat. Sci. 2020, 55, 3729–3749. [Google Scholar] [CrossRef]
  48. Noor, N.; Shapira, A.; Edri, R.; Gal, I.; Wertheim, L.; Dvir, T. 3D printing of personalized thick and perfusable cardiac patches and hearts. Adv. Sci. 2019, 6, 1900344. [Google Scholar] [CrossRef] [PubMed]
  49. Vaz, V.M.; Kumar, L. 3D printing as a promising tool in personalized medicine. AAPS PharmSciTech 2021, 22, 49. [Google Scholar] [CrossRef] [PubMed]
  50. Li, J.; Chen, M.; Wei, X.; Hao, Y.; Wang, J. Evaluation of 3D-printed polycaprolactone scaffolds coated with freeze-dried platelet- rich plasma for bone regeneration. Materials 2017, 10, 831. [Google Scholar] [CrossRef] [PubMed]
  51. Reddy, M.S.; Shetty, S.R.; Shetty, R.M.; Vannala, V.; Sk, S.; Rajasekar, S. Focus on periodontal engineering by 3D printing technology—A systematic review. J. Oral. Res. 2020, 9, 522–531. [Google Scholar] [CrossRef]
  52. Kim, J.D.; Choi, J.S.; Kim, B.S.; Chan, C.Y.; Cho, Y.W. Piezoelectric inkjet printing of polymers: Stem cell patterning on polymer substrates. Polymer 2010, 51, 2147–2154. [Google Scholar] [CrossRef]
  53. Zhu, W.; Ma, X.Y.; Gou, M.L.; Mei, D.Q.; Zhang, K.; Chen, S.C. 3D printing of functional biomaterials for tissue engineering. Curr. Op. Biotechnol. 2016, 40, 103–112. [Google Scholar] [CrossRef]
  54. Xu, T.; Kincaid, H.; Atala, A.; Yoo, J.J. High-throughput production of single-cell microparticles using an inkjet printing technology. J. Manuf. Sci. Eng. 2008, 130, 021017. [Google Scholar] [CrossRef]
  55. Cui, X.F.; Boland, T.; D-Lima, D.D.; Lotz, M.K. Thermal inkjet printing in tissue engineering and regenerative medicine. Recent Pat. Drug Deliv. Formula 2012, 6, 149–155. [Google Scholar] [CrossRef]
  56. Saunders, R.E.; Gough, J.E.; Derby, B. Delivery of human fibroblast cells by piezoelectric drop-on-demand inkjet printing. Biomaterials 2008, 29, 193–203. [Google Scholar] [CrossRef]
  57. Murphy, S.V.; Atala, A. 3D bioprinting of tissues and organs. Nat. Biotechnol. 2014, 32, 773–785. [Google Scholar] [CrossRef] [PubMed]
  58. Onses, M.S.; Sutanto, E.; Ferreira, P.M.; Alleyne, A.G.; Rogers, J.A. Mechanisms, capabilities, and applications of high-resolution electrohydrodynamic jet printing. Small 2015, 11, 4237–4266. [Google Scholar] [CrossRef] [PubMed]
  59. Gasperini, L.; Maniglio, D.; Motta, A.; Migliaresi, C. An electrohydrodynamic bioprinter for alginate hydrogels containing living cells. Tissue Eng. Part C-Methods 2015, 21, 123–132. [Google Scholar] [CrossRef] [PubMed]
  60. Workman, V.L.; Tezera, L.B.; Elkington, P.T.; Jayasinghe, S.N. Controlled generation of microspheres incorporating extracellular matrix fibrils for three-dimensional cell culture. Adv. Funct. Mater. 2014, 24, 2648–2657. [Google Scholar] [CrossRef]
  61. Hölzl, K.; Lin, S.; Tytgat, L.; Van Vlierberghe, S.; Gu, L.; Ovsianikov, A. Bioink properties before, during and after 3D bioprinting. Biofabrication 2016, 8, 032002. [Google Scholar] [CrossRef]
  62. Fan, D.; Li, Y.; Wang, X.; Zhu, T.; Wang, Q.; Cai, H.; Li, W.; Tian, Y.; Liu, Z. Progressive 3D printing technology and its application in medical materials. Front. Pharm. 2020, 11, 122. [Google Scholar] [CrossRef]
  63. Khorsandi, D.; Fahimipour, A.; Abasian, P.; Seyedig, M.; Ghanavatih, S.; Ahmadf, A.; Amoretti, A.; Fatemeh, S.; Makvandi, P.; Tay, F.R.; et al. 3D and 4D printing in dentistry and maxillofacial surgery: Printing techniques, materials, and applications. Acta Biomater. 2021, 122, 26–49. [Google Scholar] [CrossRef]
  64. Wang, Z.; Abdulla, R.; Parker, B.; Samanipour, R.; Ghosh, S.; Kim, K. A simple and high-resolution stereolithography-based 3D bioprinting system using visible light crosslinkable bioinks. Biofabrication 2015, 7, 045009. [Google Scholar] [CrossRef]
  65. Ge, L.; Dong, L.; Wang, D.; Ge, Q.; Gu, G. A digital light processing 3D printer for fast and high-precision fabrication of soft pneumatic actuators. Sens. Actuators A Phys. 2018, 273, 285–292. [Google Scholar] [CrossRef]
  66. Olakanmi, E.O.; Cochrane, R.F.; Dalgarno, K.W. A review on selective laser sintering/melting (SLS/SLM) of aluminium alloy powders: Processing, microstructure, and properties. Prog. Mater. Sci. 2015, 74, 401–477. [Google Scholar] [CrossRef]
  67. Ngo, T.D.; Kashani, A.; Imbalzano, G.; Nguyen, K.T.Q.; Hui, D. Additive manufacturing (3D printing): A review of materials, methods, applications and challenges. Compos. B Eng. 2018, 143, 172–196. [Google Scholar] [CrossRef]
  68. Park, C.H.; Rios, H.F.; Jin, Q.; Sugai, J.V.; Padial-Molina, M.; Taut, A.D.; Flanagan, C.L.; Hollister, S.J.; Giannobile, W.V. Tissue engineering bone-ligament complexes using fiber-guiding scaffolds. Biomaterials 2012, 33, 137–145. [Google Scholar] [CrossRef] [PubMed]
  69. Vaquette, C.; Fan, W.; Xiao, Y.; Hamlet, S.; Hutmacher, D.W.; Ivanovski, S. A biphasic scaffold design combined with cell sheet technology for simultaneous regeneration of alveolar bone/periodontal ligament complex. Biomaterials 2012, 33, 5560–5573. [Google Scholar] [CrossRef] [PubMed]
  70. Lee, C.H.; Hajibandeh, J.; Suzuki, T.; Fan, A.; Shang, P.; Mao, J.J. Three-dimensional printed multiphase scaffolds for regeneration of periodontium complex. Tissue Eng. Part A 2014, 20, 1342–1351. [Google Scholar] [CrossRef]
  71. Mwema, F.M.; Akinlabi, E.T. Basics of Fused Deposition modelling (FDM), in ‘Fused Deposition Modeling’; Springer: Berlin/Heidelberg, Germany, 2020; pp. 1–15. [Google Scholar]
  72. Kade, J.C.; Dalton, P.D. Polymers for melt electrowriting. Adv. Healthc. Mater. 2020, 10, 2001232. [Google Scholar] [CrossRef]
  73. Van Genderen, A.M.; Jansen, K.; Kristen, M.; van Duijn, J.; Li, Y.; Schuurmans, C.C.L.; Malda, J.; Vermonden, T.; Jansen, J.; Masereeuw, R.; et al. Topographic guidance in melt-electrowritten tubular scaffolds enhances engineered kidney tubule performance. BioRxiv 2020, 8, 617364. [Google Scholar] [CrossRef]
  74. Dalton, P.D. Melt electrowriting with additive manu- facturing principles. Curr. Opin. Biomed. Eng. 2017, 2, 49–57. [Google Scholar] [CrossRef]
  75. Chen, H.; de Malheiro, A.; van Blitterswijk, C.; Mota, C.; Wieringa, P.A.; Moroni, L. Direct writing electrospinning of scaffolds with multidimensional fiber architecture for hierarchical tissue engineering. ACS Appl. Mater. Interfaces 2017, 9, 38187–38200. [Google Scholar] [CrossRef]
  76. O’Connell, C.D.; Bridges, O.; Everett, C.; Antill-O’Brien, N.; Onofrillo, C.; Bella, C.D. Electrostatic distortion of melt-electrowritten patterns by 3D objects: Quantification, modeling, and toolpath correction. Adv. Mater. Tech. 2021, 6, 2100345. [Google Scholar] [CrossRef]
  77. Du, X.; Fu, S.; Zhu, Y. 3D printing of ceramic-based scaffolds for bone tissue engineering: An overview. J. Mater. Chem. B 2018, 6, 4397–4412. [Google Scholar] [CrossRef]
  78. Gillispie, G.; Prim, P.; Copus, J.; Fisher, J.; Mikos, A.G.; Yoo, J.J.; Atala, A.; Lee, S.J. Assessment methodologies for extrusion-based bioink printability. Biofabrication 2020, 12, 022003. [Google Scholar] [CrossRef] [PubMed]
  79. Wang, Z.; Huang, X. Elements of 3D bioprinting in periodontal regeneration: Frontiers and prospects. Processes 2021, 9, 1724. [Google Scholar] [CrossRef]
  80. Carrel, J.P.; Wiskott, A.; Moussa, M.; Rieder, P.; Scherrer, S.; Durual, S. A 3D printed TCP/HA structure as a new osteoconductive scaffold for vertical bone augmentation. Clin. Oral Implant Res. 2016, 27, 55–62. [Google Scholar] [CrossRef]
  81. Mangano, C.; Barboni, B.; Valbonetti, L.; Berardinelli, P.; Martelli, A.; Muttini, A.; Bedini, R.; Tetè, S.; Piattelli, A.; Mattioli, M. In Vivo Behavior of a custom-made 3D synthetic bone substitute in sinus augmentation procedures in sheep. J. Oral. Implantol. 2015, 41, 240–250. [Google Scholar] [CrossRef]
  82. Lakkaraju, R.; Guntakandla, V.; Gooty, J.; Palaparthy, R.; Vundela, R.; Bommireddy, V. Three-dimensional printing—A new vista for periodontal regeneration: A review. Int. J. Med. Rev. 2017, 4, 81–85. [Google Scholar] [CrossRef]
  83. Reçica, B.; Popovska, M.; Cana, A.; Bedxeti, L.Z.; Tefiku, U.; Spasovski, S.; Spasovska-Gjorgovska, A.; Kutllovci, T.; Ahmedi, J.F. Use of biomaterials for periodontal regeneration: A review. Open Access Maced J. Med. Sci. 2020, 8, 90–97. [Google Scholar] [CrossRef]
  84. Cho, H.; Tarafder, S.; Fogge, M.; Kao, K.; Lee, C.H. Periodontal ligament stem/progenitor cells with protein-releasing scaffolds for cementum formation and integration on dentin surface. Connect. Tissue Res. 2016, 57, 488–495. [Google Scholar] [CrossRef]
  85. Shim, J.-H.; Won, J.-Y.; Park, J.-H.; Bae, J.-H.; Ahn, G.; Kim, C.-H.; Lim, D.-H.; Cho, D.-W.; Yun, W.-S.; Bae, E.-B.; et al. Effects of 3D-printed polycaprolactone/β-tricalcium phosphate membranes on guided bone regeneration. Int. J. Mol. Sci. 2017, 18, 899. [Google Scholar] [CrossRef]
  86. Dubey, N.; Ferreira, J.A.; Daghrery, A.; Aytac, Z.; Malda, J.; Bhaduri, S.B.; Bottino, M.C. Highly tunable bioactive fiber-reinforced hydrogel for guided bone regeneration. Acta. Biomater. 2020, 113, 164–176. [Google Scholar] [CrossRef]
  87. Hsieh, H.Y.; Yao, C.C.; Hsu, L.F.; Tsai, L.H.; Jeng, J.H.; Young, T.H.; Chen, Y.J. Biological properties of human periodontal ligament cells spheroids cultivated on chitosan and polyvinyl alcohol membranes. J. Formosan. Med. Assoc. 2022. [Google Scholar] [CrossRef]
  88. Bai, L.; Ji, P.; Li, X.; Gao, H.; Li, L.; Wang, C. Mechanical characterization of 3D-printed individualized Ti-Mesh (membrane) for alveolar bone defects. J. Healthc Eng. 2019, 2019, 4231872. [Google Scholar] [CrossRef] [PubMed]
  89. Zhao, D.; Dong, H.; Niu, Y.; Fan, W.; Jiang, M.; Li, K.; Wei, Q.; Palin, W.; Zhang, Z. Electrophoretic deposition of novel semi- permeable coatings on 3D-printed Ti-Nb alloy meshes for guided alveolar bone regeneration. Dent. Mat. 2022, 38, 431–443. [Google Scholar] [CrossRef] [PubMed]
  90. Park, C.H.; Rios, H.F.; Jin, Q.; Bland, M.E.; Flanagan, C.L.; Hollister, S.J.; Giannobile, W.V. Biomimetic hybrid scaffolds for engineering human tooth-ligament interfaces. Biomaterials 2010, 31, 5945–5952. [Google Scholar] [CrossRef] [PubMed]
  91. Costa, P.F.; Vaquette, C.; Zhang, Q.; Reis, R.L.; Ivanovski, S.; Hutmacher, D.W. Advanced tissue engineering scaffold design for regeneration of the complex hierarchical periodontal structure. J. Clin. Periodontol. 2014, 41, 283–294. [Google Scholar] [CrossRef] [PubMed]
  92. Wang, C.Y.; Chiu, Y.C.; Lee, A.K.; Lin, Y.A.; Lin, P.Y.; Shie, M.Y. Biofabrication of gingival fibroblast cell-laden collagen/strontium-doped calcium silicate 3D-printed bi-layered scaffold for osteoporotic periodontal regeneration. Biomedicines 2021, 9, 431. [Google Scholar] [CrossRef] [PubMed]
  93. Carter, S.S.; Costa, P.F.; Vaquette, C.; Ivanovski, S.; Hutmacher, D.W.; Malda, J. Additive biomanufacturing: An advanced approach for periodontal tissue regeneration. Ann. Biomed. Eng. 2017, 45, 12–22. [Google Scholar] [CrossRef]
  94. Huang, R.Y.; Tai, W.C.; Ho, M.H.; Chang, P.C. Combination of a biomolecule-aided biphasic cryogel scaffold with a barrier membrane adhering PDGF-encapsulated nanofibers to promote periodontal regeneration. J. Periodontal. Re. 2020, 55, 529–538. [Google Scholar] [CrossRef]
  95. Daghrery, A.; de Souza, I.J.; Castilho, M.; Malda, J.; Bottino, M.C. Unveiling the potential of melt electrowritting in regenerative dental medicine. Acta Biomater. 2022, in press. [Google Scholar] [CrossRef]
  96. Park, S.A.; Lee, H.-J.; Kim, K.-S.; Lee, S.J.; Lee, J.-T.; Kim, S.-Y.; Chang, N.-H.; Park, S.-Y. In vivo evaluation of 3D-printed polycaprolactone scaffold implantation combined with β-TCP powder for alveolar bone augmentation in a beagle defect model. Materials 2018, 11, 238. [Google Scholar] [CrossRef]
  97. Goh, B.T.; The, L.Y.; Tan, D.B.; Zhang, Z.; Teoh, S.H. Novel 3D polycaprolactone scaffold for ridge preservation—A pilot randomised controlled clinical trial. Clin. Oral. Implants Res. 2014, 26, 271–277. [Google Scholar] [CrossRef]
  98. Kijartorn, P.; Thammarakcharoen, F.; Suwanprateeb, J.; Buranawat, B. The use of three dimensional printed hydroxyapatite granules in alveolar ridge preservation. Key Eng. Mater. 2017, 751, 663–667. [Google Scholar] [CrossRef]
  99. Gul, M.; Arif, A.; Ghafoor, R. Role of three-dimensional printing in periodontal regeneration and repair: Literature review. J. Indian Soc. Periodontol. 2019, 23, 504–510. [Google Scholar] [CrossRef] [PubMed]
  100. Torres, J.; Tamimi, F.; Alkhraisat, M.H.; Prados-Frutos, J.C.; Rastikerdar, E.; Gbureck, U.; Barralet, J.E.; López-Cabarcos, E. Vertical bone augmentation with 3D-synthetic monetite blocks in the rabbit calvaria. J. Clin. Periodontol. 2011, 38, 1147–1153. [Google Scholar] [CrossRef] [PubMed]
  101. Yen, H.H.; Stathopoulou, P.G. CAD/CAM and 3D-printing applications for alveolar ridge augmentation. Curr. Oral. Health Rep. 2018, 5, 127–132. [Google Scholar] [CrossRef]
  102. Reis, E.C.C.; Borges, A.P.; Araújo, M.V.; Mendes, V.C.; Guan, L.; Davies, J.E. Periodontal regeneration using a bilayered PLGA/calcium phosphate construct. Biomaterials 2011, 32, 9244–9253. [Google Scholar] [CrossRef]
  103. Obregon, F.; Vaquette, C.; Ivanovski, S.; Hutmacher, D.W.; Bertassoni, L.E. Three-dimensional bioprinting for regenerative dentistry and craniofacial tissue engineering. J. Dent. Res. 2015, 94, 143S–152S. [Google Scholar] [CrossRef]
  104. Park, C.H.; Kim, K.H.; Rios, H.F.; Lee, Y.M.; Giannobile, W.V.; Seol, Y.J. Spatiotemporally controlled microchannels of periodontal mimic scaffolds. J. Dent. Res. 2014, 93, 1304–1312. [Google Scholar] [CrossRef]
  105. Sumida, T.; Otawa, N.; Kamata, Y.U.; Kamakura, S.; Mtsushita, T.; Kitagaki, H.; Mori, S.; Sasaki, K.; Fujibayashi, S.; Takemoto, M.; et al. Custom-made titanium devices as membranes for bone augmentation in implant treatment: Clinical application and the comparison with conventional titanium mesh. J. Craniomaxillofac Surg. 2015, 43, 2183–2188. [Google Scholar] [CrossRef]
  106. Pilipchuk, S.P.; Monje, A.; Jiao, Y.; Hao, J.; Kruger, L.; Flanagan, C.L.; Hollister, S.J.; Giannobile, W.V. Integration of 3D printed and micropatterned polycaprolactone scaffolds for guidance of oriented collagenous tissue formation in vivo. Adv. Healthc Mater 2016, 5, 676–687. [Google Scholar] [CrossRef]
  107. Adel-Khattab, D.; Giacomini, F.; Gildenhaar, R.; Berger, G.; Gomes, C.; Linow, U.; Hardt, M.; Peleska, B.; Günster, J.; Stiller, M.; et al. Development of a synthetic tissue engineered three-dimensional printed bioceramic-based bone graft with homogenously distributed osteoblasts and mineralizing bone matrix in vitro. J. Tissue Eng. Regen. Med. 2018, 12, 44–58. [Google Scholar] [CrossRef]
  108. Lei, L.; Yu, Y.; Ke, T.; Sun, W.; Chen, L. The application of three-dimensional printing model and platelet-rich fibrin technology in guided tissue regeneration surgery for severe bone defects. J. Oral. Implantol. 2019, 45, 35–43. [Google Scholar] [CrossRef] [PubMed]
  109. Bartnikowski, M.; Vaquette, C.; Ivanovski, S. Workflow for highly porous resorbable custom 3D printed scaffolds using medical grade polymer for large volume alveolar bone regeneration. Clin. Oral. Implants Res. 2020, 31, 431–441. [Google Scholar] [CrossRef] [PubMed]
  110. Tamimi, F.; Torres, J.; Gbureck, U.; Lopez-Cabarcos, E.; Bassett, D.C.; Alkhraisat, M.H.; Barralet, J.E. Craniofacial vertical bone augmentation: A comparison between 3D printed monolithic monetite blocks and autologous onlay grafts in the rabbit. Biomaterials 2009, 30, 6318–6326. [Google Scholar] [CrossRef] [PubMed]
  111. Di Giacomo, G.A.; da Silva, J.V.; da Silva, A.M.; Paschoal, G.H.; Cury, P.R.; Szarf, G. Accuracy and complications of computer-designed selective laser sintering surgical guides for flapless dental implant placement and immediate definitive prosthesis installation. J. Periodontol. 2012, 83, 410–419. [Google Scholar] [CrossRef]
  112. Cassetta, M.; Di Mambro, A.; Giansanti, M.; Stefanelli, L.V.; Cavallini, C. The intrinsic error of a stereolithographic surgical template in implant guided surgery. Int. J. Oral. Maxillofac. Surg. 2013, 42, 264–275. [Google Scholar] [CrossRef]
  113. Pozzi, A.; Tallarico, M.; Marchetti, M.; Scarfò, B.; Esposito, M. Computer-guided versus free-hand placement of immediately loaded dental implants: 1-year post-loading results of a multicentre randomised controlled trial. Eur. J. Oral. Implantol. 2014, 7, 229–242. [Google Scholar]
  114. Stübinger, S.; Buitrago-Tellez, C.; Cantelmi, G. Deviations between placed and planned implant positions: An accuracy pilot study of skeletally supported stereolithographic surgical templates. Clin. Implant. Dent. Relat. Res. 2014, 16, 540–551. [Google Scholar] [CrossRef]
  115. Shen, P.; Zhao, J.; Fan, L.; Qiu, H.; Xu, W.; Wang, Y.; Zhang, S.; Kim, Y.-J. Accuracy evaluation of computer-designed surgical guide template in oral implantology. J. Craniomaxillofac Surg. 2015, 43, 2189–2194. [Google Scholar] [CrossRef]
  116. Verhamme, L.M.; Meijer, G.J.; Boumans, T.; de Haan, A.F.; Bergé, S.J.; Maal, T.J. A clinically relevant accuracy study of computer-planned implant placement in the edentulous maxilla using mucosa-supported surgical templates. Clin. Implant. Dent. Relat. Res. 2015, 17, 343–352. [Google Scholar] [CrossRef]
  117. Xu, L.W.; You, J.; Zhang, J.X.; Liu, Y.F.; Peng, W. Impact of surgical template on the accuracy of implant placement. J. Prosthodont. 2016, 25, 641–646. [Google Scholar] [CrossRef]
  118. Bernard, L.; Vercruyssen, M.; Duyck, J.; Jacobs, R.; Teughels, W.; Quirynen, M. A randomized controlled clinical trial comparing guided with nonguided implant placement: A 3-year follow-up of implant-centered outcomes. J. Prosthet. Dent. 2019, 121, 904–910. [Google Scholar] [CrossRef] [PubMed]
  119. Kislitsyn, A.; Savinkov, R.; Novkovic, M.; Onder, L.; Bocharov, G. Computational approach to 3D Modeling of the lymph node geometry. Computation 2015, 3, 222–234. [Google Scholar] [CrossRef]
  120. Rodriguez, M.J.; Brown, J.; Giordano, J.; Lin, S.J.; Omenetto, F.G.; Kaplan, D.L. Silk based bioinks for soft tissue reconstruction using 3-dimensional (3D) printing with in vitro and in vivo assessments. Biomaterials 2017, 117, 105–115. [Google Scholar] [CrossRef] [PubMed]
  121. Miri, A.K.; Mirzaee, I.; Hassan, S.; Mesbah Oskui, S.; Nieto, D.; Khademhosseini, A.; Zhang, Y.S. Effective bioprinting resolution in tissue model fabrication. Lab. Chip. 2019, 19, 2019–2037. [Google Scholar] [CrossRef] [PubMed]
  122. Tasoglu, S.; Demirci, U. Bioprinting for stem cell research. Trends Biotechnol. 2013, 31, 10–19. [Google Scholar] [CrossRef]
  123. Albritton, J.L.; Miller, J.S. 3D bioprinting: Improving in vitro models of metastasis with heterogeneous tumor microenvironments. Dis. Model. Mech. 2017, 10, 3–14. [Google Scholar] [CrossRef]
  124. Chang, R.; Nam, J.; Sun, W. Direct cell writing of 3D microorgan for in vitro pharmacokinetic model. Tissue Eng. Part C Methods 2008, 14, 157–166. [Google Scholar] [CrossRef]
  125. Mandrycky, C.; Wang, Z.; Kim, K.; Kim, D.-H. 3D bioprinting for engineering complex tissues. Biotechnol. Adv. 2016, 34, 422–434. [Google Scholar] [CrossRef]
  126. Melchels, F.P.; Domingos, M.A.; Klein, T.J.; Malda, J.; Bartolo, P.J.; Hutmacher, D.W. Additive manufacturing of tissues and organs. Prog. Polym. Sci. 2012, 37, 1079–1104. [Google Scholar] [CrossRef]
  127. Rodriguez-Salvador, M.; Ruiz-Cantu, L. Revealing emerging science and technology research for dentistry applications of 3D bioprinting. Int. J. Bioprint. 2018, 5, 170. [Google Scholar] [CrossRef]
  128. Ingber, D.E.; Mow, V.C.; Butler, D.; Niklason, L.; Huard, J.; Mao, J.; Yannas, I.; Kaplan, D.; Vunjak-Novakovic, G. Tissue engineering and developmental biology: Going biomimetic. Tissue Eng. 2006, 12, 3265–3283. [Google Scholar] [CrossRef] [PubMed]
  129. Shu, A.F. 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, 44102. [Google Scholar]
  130. Li, J.P.; Chen, M.J.; Fan, X.Q.; Zhou, H.F. Recent advances in bioprinting techniques: Approaches, applications and future prospects. J. Transl. Med. 2016, 14, 271. [Google Scholar] [CrossRef]
  131. Matai, I.; Kaur, G.; Seyedsalehi, A.; McClinton, A.; Laurencin, C.T. Progress in 3D bioprinting technology for tissue/organ regenerative engineering. Biomaterials 2020, 226, 119536. [Google Scholar] [CrossRef]
  132. Kato, B.; Wisser, G.; Agrawal, D.K.; Wood, T.; Thankam, F.G. 3D bioprinting of cardiac tissue: Current challenges and perspectives. J. Mater. Sci. Mater. Med. 2021, 32, 54. [Google Scholar] [CrossRef]
  133. Thattaruparambil Raveendran, N.; Vaquette, C.; Meinert, C.; Ipe, D.S.; Ivanovski, S. Optimization of 3D bioprinting of periodontal ligament cells. Dent. Mater. 2019, 35, 1683–1694. [Google Scholar] [CrossRef] [PubMed]
  134. Ma, Y.; Ji, Y.; Huang, G.; Ling, K.; Zhang, X.; Xu, F. Bioprinting 3D cell-laden hydrogel microarray for screening human periodontal ligament stem cell response to extracellular matrix. Biofabrication 2015, 7, 044105. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, L.; Xu, W.; Chen, Y.; Wang, J. Alveolar bone repair of rhesus monkeys by using BMP-2 gene and mesenchymal stem cells loaded three-dimensional printed bioglass scaffold. Sci. Rep. 2019, 9, 18175. [Google Scholar] [CrossRef]
  136. Lin, H.-H.; Chao, P.-H.G.; Tai, W.-C.; Chang, P.-C. 3D-printed collagen-based waveform microfibrous scaffold for periodontal ligament reconstruction. Int. J. Mol. Sci. 2021, 22, 7725. [Google Scholar] [CrossRef]
  137. Vurat, M.T.; Şeker, Ş.; Lalegül-Ülker, Ö.; Parmaksiz, M.; Elçin, A.E.; Elçin, Y.M. Development of a multicellular 3D-bioprinted microtissue model of human periodontal ligament-alveolar bone biointerface: Towards a pre-clinical model of periodontal diseases and personalized periodontal tissue engineering. Genes Dis. 2020, 9, 1008–1023. [Google Scholar] [CrossRef]
  138. Li, Z.; Li, S.; Yang, J.; Ha, Y.; Zhang, Q.; Zhou, X.; He, C. 3D bioprinted gelatin/gellan gum-based scaffold with double-crosslinking network for vascularized bone regeneration. Carbohydr. Polym. 2022, 290, 119469. [Google Scholar] [CrossRef] [PubMed]
  139. Castilho, M.; de Ruijter, M.; Beirne, S.; Ito, K. Multitechnology biofabrication: A new approach for the manufacturing of functional tissue structures? Trends Biotechnol. 2020, 38, 1316–1328. [Google Scholar] [CrossRef] [PubMed]
  140. De Ruijter, M.; Ribeiro, A.; Dokter, I.; Castilho, M.; Malda, J. Simultaneous micropatterning of fibrous meshes and bioinks for the fabrication of living tissue constructs. Adv. Healthc. Mater. 2019, 8, 1800418. [Google Scholar] [CrossRef] [PubMed]
  141. Diloksumpan, P.; de Ruijter, M.; Castilho, M.; Gbureck, U.; Vermonden, T.; van Weeren, P.R.; Malda, J.; Levato, R. Combining multi-scale 3D printing technologies to engineer reinforced hydrogel-ceramic interfaces. Biofabrication 2020, 12, 025014. [Google Scholar] [CrossRef] [PubMed]
  142. Dos Santos, D.M.; de Annunzio, S.R.; Carmello, J.C.; Pavarina, A.C.; Fontana, C.R.; Correa, D.S. Combining coaxial electrospinning and 3D printing: Design of biodegradable bilayered membranes with dual drug delivery capability for periodontitis treatment. ACS Appl. Bio. Mater 2022, 5, 146–159. [Google Scholar] [CrossRef]
  143. Liu, J.; Zou, Q.; Wang, C.; Lin, M.; Li, Y.; Zhang, R.; Li, Y. Electrospinning and 3D printed hybrid bi-layer scaffold for guided bone regeneration. Mat. Design 2021, 210, 110047. [Google Scholar] [CrossRef]
  144. Gao, B.; Yang, Q.; Zhao, X.; Jin, G.; Ma, Y.; Xu, F. 4D Bioprinting for biomedical applications. Trends Biotechnol. 2016, 34, 746–756. [Google Scholar] [CrossRef]
  145. Betsch, M.; Cristian, C.; Lin, Y.Y.; Blaeser, A.; Schöneberg, J.; Vogt, M.; Buhl, E.M. Incorporating 4D into bioprinting: Real-time magnetically directed collagen fiber alignment for generating complex multi- layered tissues. Adv. Healthc. Mater. 2018, 7, 1800894. [Google Scholar] [CrossRef]
  146. Liu, F.; Wang, W.; Mirihanage, W.; Hinduja, S.; Bartolo, P. A plasma- assisted bioextrusion system for tissue engineering. CIRP Ann. 2018, 67, 229–232. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The evolution of 3D printing technologies.
Figure 1. The evolution of 3D printing technologies.
Membranes 12 00902 g001
Figure 2. Droplet-based 3D printing techniques; DOD: drop-on-demand.
Figure 2. Droplet-based 3D printing techniques; DOD: drop-on-demand.
Membranes 12 00902 g002
Figure 3. Laser-assisted bioprinting.
Figure 3. Laser-assisted bioprinting.
Membranes 12 00902 g003
Figure 4. Stereolithography schematic principle.
Figure 4. Stereolithography schematic principle.
Membranes 12 00902 g004
Figure 5. Digital projection printing principle.
Figure 5. Digital projection printing principle.
Membranes 12 00902 g005
Figure 6. Selective laser sintering principle.
Figure 6. Selective laser sintering principle.
Membranes 12 00902 g006
Figure 7. Thermal extrusion 3D printing: fused deposition modelling (right) and melt electrowritting (left).
Figure 7. Thermal extrusion 3D printing: fused deposition modelling (right) and melt electrowritting (left).
Membranes 12 00902 g007
Figure 8. Non-thermal extrusion 3D printing.
Figure 8. Non-thermal extrusion 3D printing.
Membranes 12 00902 g008
Figure 9. Schematic view of monophasic and multiphasic scaffolds.
Figure 9. Schematic view of monophasic and multiphasic scaffolds.
Membranes 12 00902 g009
Table 1. Main biomaterials used in scaffold 3D printing.
Table 1. Main biomaterials used in scaffold 3D printing.
MaterialAdvantagesDisadvantages
Natural polymers
Collagen
Alginate
Hyaluronic acid
Chitosan
Biocompatible
Good cell affinity
Hydrophilicity
Antibacterial effect
Low mechanical properties
Fast degradation rate
Lack of bioactivity
Synthetic polymers
Polycaprolactone (PCL)
Polylactic acid (PLA)
Polyglycolic acid (PGA)
Polyethylene glycol (PEG)
Poly(lactic-co-glycolic) acid (PLGA)
Highly adjustable physiochemical and mechanical properties
Wide range of degradation and resorption kinetics
Good repeatability
Low bioactivity
Slow degradation rate
Acidic byproducts
Bio-ceramics
Hydroxyapatite (HA)
β-tricalcium phosphate (β-TCP)
Bioactive glass
Bioactive
Biocompatible
Osteoconductive
Potential osteoinductive
Hydrophilicity
Not compatible with cell encapsulation
Stiffness
Brittleness
Low ductility
Low flexibility
Inconsistent cell reactions (variations in surface quality)
Table 2. Applications of 3D printing in periodontology.
Table 2. Applications of 3D printing in periodontology.
ApplicationAuthorsType of StudyMethodMaterial3D Printer
GTRKim et al., 2010 [52]In vivo3D-printed tooth scaffoldPoly-epsilon caprolactone and hydroxyapatiteNot mentioned
Park et al., 2010 [90]In vivo3D-printed scaffoldPCL-PGA3D wax-printing system (ModelMaker II, Solidscape, Inc., Merrimack, NH, USA)
Carlo Reis et al., 2011 [102]In vivo3D-printed scaffoldPLGA/CaP bilayered biomaterialNot mentioned
Park et al., 2012 [68]In vivo3D-printed scaffoldPoly-ε caprolactone solution (PCL)3-D rapid prototyping wax printer (ModelMaker II; Solidscape Inc., Merrimack, NH, USA)
Obregon et al., 2015 [103]In vivo3D-printed scaffoldBilayered biomaterialNot mentioned
Vaquette et al., 2012 [69]In vivoFDM + solution electrospinningPCLFDM, Osteopore Inc. Singapore
In-house solution spinning device
Costa et al., 2014 [91]In vivo3D-printed scaffoldBilayered biomaterialNot mentioned
Park et al., 2014 [104]In vivo3D-printed scaffoldGelatin, chitosanNot mentioned
Lee et al., 2014 [70]In vivoLayer-by-layer depositionPCL + hydroxyapatiteBioplotter, EnvisionTEC
Rasperini et al., 2015 [26]Case report3D-printed Bioresorbable ScaffoldPCLSLS (Formiga P100 System; EOS e-Manufacturing Solutions, Pflugerville, TX, USA))
Sumida et al., 2015 [105]RCT3D-printed scaffoldTitaniumNot mentioned
Pilipchuk et al., 2016 [106]Preclinical study3D-printed scaffoldPCLNot mentioned
Adel-Khattab et al., 2018 [107]In vitro3D-printed scaffoldBioceramicR1Series ExOne (PROMETAL, North Huntingdon, PA, USA)
Lei et al. 2019 [108]Case report3D-printed bone modelNot mentionedNot mentioned
Bartnikowski et al., 2020 [109]RCTLayer-by-layer depositionPCLBioplotter, EnvisionTEC, Dearborn, MI, USA
Socket preservationGoh et al., 2015 [97]Pilot RCT3D-printed bioresorbable scaffoldPCLFDM techniques (FDM 3000; Stratasys, Eden Prairie, MN, USA)
Kijartorn et al., 2017 [98]Prospective study3D-printed scaffoldHydroxyapatite granulesProjet 160, 3D systems
Park et al.,
2018 [96]
In vivo3D-printed bioresorbable scaffoldPCL3D bioprinting system (laboratory -made system in Korea Institute of Machinery and Materials, Daejeon, Korea)
Vertical bone augmentationTamimi et al., 2009 [110]Case report3D-printed monolithic monetite blocksSynthetic calcium phosphates3D-powder
Printing system (Z-Corporation, Burlington, MA, USA)
Torres et al., 2011 [100]In vivo3D-printed monolithic monetite blocksA/b-tricalcium phosphate3D-powder
Printing system (Z-Corporation, Burlington, MA, USA)
Sinus augmentationMangano et al., 2015 [103]In vivo3D synthetic bone substituteCeramicNot mentioned
Guided implant placementDi Giacomo et al., 2005 [111]NRCTSLA surgical guidesPolymerSimplant CSI Materialise, Ann Arbor, MI, USA
Cassetta et al., 2013 [112]Retrospective3D-printed surgical guideAcrylicSLA surgical guide (External Hex Safe1, Materialise Dental, Leuven, Belgium)
Pozzi et al., 2014 [113]Clinical trialSLA surgical guidesAcrylic resinNobel Procera, Nobel Biocare, Zurich, Switzerland
Stübinger et al., 2014 [114]Prospective3D-printed surgical guidePolymerAstra Tech AB, Mölndal, Sweden
Shen et al., 2015 [115]RCTSLA templatesAcrylicGeomagic, version 10.0, Geomagic, Research triangle Park, NC, USA
Verhamme et al., 2015 [116]Prospective3D-printed surgical guideNot mentionedNobelGuide (Nobel Biocare, Gothenburg, Sweden
Xu et al., 2016 [117]In vitroSLA surgical guidesAcrylicConne×350; Objet, Rehovot, Israel
Bernard et al., 2019 [118]RCTSLA surgical guidesAcrylicSimplant; Materialise Dental, Waltham, MA, USA
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sufaru, I.-G.; Macovei, G.; Stoleriu, S.; Martu, M.-A.; Luchian, I.; Kappenberg-Nitescu, D.-C.; Solomon, S.M. 3D Printed and Bioprinted Membranes and Scaffolds for the Periodontal Tissue Regeneration: A Narrative Review. Membranes 2022, 12, 902. https://doi.org/10.3390/membranes12090902

AMA Style

Sufaru I-G, Macovei G, Stoleriu S, Martu M-A, Luchian I, Kappenberg-Nitescu D-C, Solomon SM. 3D Printed and Bioprinted Membranes and Scaffolds for the Periodontal Tissue Regeneration: A Narrative Review. Membranes. 2022; 12(9):902. https://doi.org/10.3390/membranes12090902

Chicago/Turabian Style

Sufaru, Irina-Georgeta, Georgiana Macovei, Simona Stoleriu, Maria-Alexandra Martu, Ionut Luchian, Diana-Cristala Kappenberg-Nitescu, and Sorina Mihaela Solomon. 2022. "3D Printed and Bioprinted Membranes and Scaffolds for the Periodontal Tissue Regeneration: A Narrative Review" Membranes 12, no. 9: 902. https://doi.org/10.3390/membranes12090902

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop