Next Article in Journal
Intrinsically Disordered Synthetic Polymers in Biomedical Applications
Next Article in Special Issue
Evaluation of Self-Degradation and Plugging Performance of Temperature-Controlled Degradable Polymer Temporary Plugging Agent
Previous Article in Journal
Poly(3-hydroxybutyrate) (PHB) and Polycaprolactone (PCL) Based Blends for Tissue Engineering and Bone Medical Applications Processed by FDM 3D Printing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dental Materials Applied to 3D and 4D Printing Technologies: A Review

1
Department and Research Institute of Dental Biomaterials and Bioengineering, Yonsei University College of Dentistry, Seoul 03722, Republic of Korea
2
The CONVERSATIONALIST Club, School of Stomatology, Shandong First Medical University, Jinan 250117, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Polymers 2023, 15(10), 2405; https://doi.org/10.3390/polym15102405
Submission received: 23 April 2023 / Revised: 9 May 2023 / Accepted: 17 May 2023 / Published: 22 May 2023
(This article belongs to the Special Issue Preparation and Applications of Biodegradable Polymer Materials)

Abstract

:
As computer-aided design and computer-aided manufacturing (CAD/CAM) technologies have matured, three-dimensional (3D) printing materials suitable for dentistry have attracted considerable research interest, owing to their high efficiency and low cost for clinical treatment. Three-dimensional printing technology, also known as additive manufacturing, has developed rapidly over the last forty years, with gradual application in various fields from industry to dental sciences. Four-dimensional (4D) printing, defined as the fabrication of complex spontaneous structures that change over time in response to external stimuli in expected ways, includes the increasingly popular bioprinting. Existing 3D printing materials have varied characteristics and scopes of application; therefore, categorization is required. This review aims to classify, summarize, and discuss dental materials for 3D printing and 4D printing from a clinical perspective. Based on these, this review describes four major materials, i.e., polymers, metals, ceramics, and biomaterials. The manufacturing process of 3D printing and 4D printing materials, their characteristics, applicable printing technologies, and clinical application scope are described in detail. Furthermore, the development of composite materials for 3D printing is the main focus of future research, as combining multiple materials can improve the materials’ properties. Updates in material sciences play important roles in dentistry; hence, the emergence of newer materials are expected to promote further innovations in dentistry.

1. Introduction

The development and application of dental materials have a long history. Gold and silver, which were used by the Romans in 700 B.C. to repair cavities, are the first materials used for dental restoration. In the 18th century, wax was introduced to obtain impressions, which were then filled with a mixture of plaster and cementum to fabricate oral models [1]. Ceramics were then introduced into dentistry to manufacture porcelain restorations [2]. In the 20th century, composite resins, cement, titanium, stainless steel, and other materials were introduced into dental sciences [3]. Since the 21st century, computer-aided design and computer-aided manufacturing (CAD/CAM) has gradually gained popularity. Three-dimensional (3D) printing materials are more widely used today.
Three-dimensional printing technology, also known as additive manufacturing (AM), first appeared in the late 1970s. Three-dimensional printing gradually builds models by depositing materials layer by layer, which can be used for rapid prototyping and developing personalized 3D products using standardized materials based on CAD models. It uses CAD software to form a 3D digital model and transfer it to a 3D printer, which then converts the computerized digital model into layer-by-layer two-dimensional (2D) parts, generating solid layers to build the desired object [4,5,6,7,8,9,10]. At present, 3D printing has been widely used in many fields of life and has made significant progress [11]. Additive manufacturing technologies for 3D printing can be divided into 7 categories, based on how materials or inks are deposited: stereolithography (SLA), fused deposition modeling (FDM), digital light projection (DLP), selective laser sintering (SLS), computed axial lithography, photopolymer jetting, and powder binder printing [4,12].
In 2013, four-dimensional (4D) printing was proposed for the first time. Shape changing with time is the fourth dimension, besides the three dimensions of length, width, and height [13,14]. In addition, the shape, function, or properties of 4D printed objects can change under various stimuli, such as heat, light, electricity, and magnetic field [15]. Four-dimensional printing has advantages in terms of material adaptability, facilitating the precise configuration of material responsiveness. Further, the bioprinting included in it is an emerging technology, and its biggest advantage is the ability to create 3D structures of organisms (such as tissues, organs, nutrients, and cells), and the manufactured biological structures can change their functions [12]. Currently, 4D printing plays a major role in the medical field, mainly for tissue generation and transplantation. Four-dimensional printing, defined as the fabrication of complex spontaneous structures that change over time in response to external stimuli in expected ways, includes the increasingly popular 4D bioprinting [12,16]. Four-dimensional printing technology is roughly the same as 3D printing. Furthermore, 4D printing technology also involves inkjet bioprinting, extrusion-based bioprinting (EBB), SLA-based bioprinting, and laser-based bioprinting (LAB). Among them, inkjet printing and DIW are preferred, which are in their infancy and facing enormous challenges [17,18,19,20,21,22].
Commonly used materials for dental 3D printing include polymers, ceramics, and metals, most of which possess high accuracies, excellent biocompatibilities, and good mechanical properties. Materials used for 4D printing mainly include Shape Memory Polymers (SMPs) and hydrogels. SMPs include thermally induced SMP, photochromic SMP, and chemical induction SMP [12]. Four-dimensional printing materials are summarized as synthetic polymers (polylactic acid (PLA), acrylonitrile butadiene styrene (ABS), polyvinyl alcohol (PVA), polycaprolactone (PCL), polymethyl methacrylate (PMMA), etc.) and natural biopolymers (hyaluronic acid (HA), chitosan, alginate, etc.), and related cells and growth factors work with them.
However, depending on their specific principles, 3D/4D printing technologies can individually adapt to different materials. Given the rapid advances in digital technologies and specific clinical requirements, an overview of printing materials could be helpful. This review aims to list contemporary 3D printing and 4D printing materials and describe their properties, current stage of development, features, applicable 3D/4D printing technologies, and corresponding indications and applications in dentistry. Emerging 3D-printing dental materials with noteworthy potential for future developments are also discussed. Table 1 lists the differences between 3D printing and 4D printing in terms of technology, materials, applications, and so on.

2. Three-dimensional Printing Materials

2.1. Polymers

Polymers are the most popular materials used for dental and maxillofacial surgical 3D printing. They can be used to manufacture surgical guides, custom trays, working casts, and temporary restorations, supported by implants, crowns, and bridges [23,24]. Polycaprolactone (PCL), polymethyl methacrylate (PMMA), polylactic acid (PLA), poly (lactic-co-glycolic acid) (PLGA), and ultraviolet (UV) resins are some of the representative polymers used for 3D printing in dentistry.

2.1.1. PCL

As an aliphatic polyester, PCL has superior biocompatibility, adjustable degradability, and wide applicability in the biomedical field. A lower melting point (59–64 °C) and excellent compatibility with other substances that are easy to manufacture make it a suitable substitute for materials processed by FDM [25]. PCL is stable in vivo, owing to its high hydrophilicity and solvent solubility. During the process of melt extrusion in 3D printing, its crystallinity and crystal directionality are developed [26].
PCL and its composites are widely used in tissue engineering [27]. The mechanical properties of such scaffolds are affected by many factors, such as composition and porosity [28]. PCL functions as a membrane for guided bone regeneration, and its biocompatibility can be improved by modification. Compared with other printing materials, PCL, with a lower melting point, is considered to be more biocompatible, for it is less likely to cause inflammatory reactions [29,30,31]. Bone defects implanted with membrane-loaded human recombinant bone morphogenetic protein-2 (rhBMP-2) have been shown to achieve calvaria defect healing in 8 weeks [32]. The porosities of the membrane and scaffolds are important influencing factors of bone-healing [33,34]. The designed pores provide space for cell proliferation. A mixture containing bioactive materials, such as hydroxyapatite, β-tricalcium phosphate, and rhBMP-2, enhances biocompatibility and promotes new bone formation in in vivo experiments [35,36]. Three-dimensionally printed PCL scaffolds have been used in bone tissue regeneration (Figure 1A) [37].

2.1.2. PMMA

PMMA, which was first discovered in the 1930s [41], is a commonly used polymer in dentistry. Given its advantages, such as ease of manufacturing, lower cost, and stability in the oral environment, PMMA is the most popular base material for dentures. It also functions as bone cement for screw fixation in the bone, can fill bone defects in cavities and the skull, and can be used to stabilize vertebrae [42]. Other applications include temporary crowns and bridges, obturators, and retainers in orthodontics [43]. Even though PMMA has been applied to prosthetic restorations with predictable clinical performance, the interactions between the oral environment and PMMA are not fully studied by scientists, and potential side impacts on cells have been discovered [44]. Compared with polyamide-12, PMMA can maintain its color as a denture base material [45]. Nanodiamonds in low concentrations can be added as reinforcement to improve the overall properties of PMMA-based fixed interim prostheses [46]. PMMA can be manufactured by FDM [47]. One study also found that SLA technology could be used to print PMMA composites with three different reinforcements: aluminum nitride, titanium oxide, and barium titanate [48]. However, the limited number of studies on 3D printed PMMA compositions and biocompatibility require further investigations. Even though the mechanical properties of 3D printed PMMA are worse than those of PMMA machined by traditional methods, the optimized parameters, as well as infiltration of epoxy and smaller particles, may influence the qualities of the final products [49]. As shown in Figure 1B, 3D printed PMMA functions as a provisional crown after accurate design [38].

2.1.3. PLA

Polylactic acid (PLA) is also a hydrophilic and aliphatic polyester, which is biodegradable and environmentally friendly, making it an ideal material for 3D printing in the future. PLA can also be manufactured by FDM. The process of 3D printing leads to lower molecular weight and degradation temperature, without changes in the semicrystalline polymer structures [50].
However, deformation will occur when the environmental temperature exceeds 50 °C, which may limit its application in 3D printing. Therefore, the idea of forming hybrid materials with other polymers arises. When combined with other bioactive substances, PLA-integrated guided tissue and guided bone regeneration (GTR/GBR) membrane reinforced by magnesium has the best charge capacity, corrosion resistance, and cell adsorption [51]. It has been found that PLA scaffolds combined with nanohydroxyapatite can function as carriers for cells in clinical dental treatment [52]. Deng et al. found that maxillary complete dentures made by FDM with PLA fulfill the need for accuracy [53]. Three-dimensionally printed PLA biodegradable scaffolds were used to produce cell-derived decellularized matrices; this has increased the applicability of 3D printing technology in dental regenerative medicine [54]. Three-dimensionally printed PLA functions as a drilling guide for oral implantology [55]. In general, PLA is an ideal material, with good clinical prospects. PLA can also be considered as a raw material for 3D printed bio-scaffolds for hard tissue, with apatite consisting of calcium and phosphate formed on its surface, after being immersed in simulated body fluid solution for a certain period (Figure 1C) [39].

2.1.4. PLGA

PLGA, which is a copolymer of PLA and polyglycolic acid (PGA), is considered an ideal material, owing to its biodegradability and biocompatibility. Fabricated using a variety of methods, PLGA-based composites combined with metal reinforcements have vast dental application prospects. The composites allow diverse properties, including enhanced cell proliferation and improved antimicrobial properties [56]. The physicochemical properties of PLGA allow manufacturing by FDM. The optimized 3D printing parameters vary for copolymers formed with various ratios of PLA and PGA, which influence the molecular weight, end cap of the copolymer, and physiochemical properties, such as viscosity and heat flow under various temperatures, and a higher proportion of PLA in PLGA is recommended for 3D printing [57].
Three-dimensionally printed PLGA scaffolds have been introduced in tissue engineering. Mironov et al. investigated novel bioresorbable scaffolds by 3D printing based on the PLGA, and it was observed that these scaffolds were non-cytotoxic, demonstrated cell proliferation in the chosen stem cells, and had excellent viscosity [58]. When repairing bone defects, a novel porous scaffold containing PLGA, tricalcium phosphate (TCP), and Mg powder was manufactured using low-temperature rapid prototyping technology, and it has been demonstrated that this newly designed scaffold has obvious biosecurity, promotes new bone and vessel formation, and improves bone tissue quality in rabbit models [59]. The in vivo characteristics of 3D printed PLGA scaffolds have also been assessed. Excellent osteoconductivity and biocompatibility have been observed in the periosteum and in animals with iliac defects [60]. Biological membranes can thus be manufactured by 3D printed PLGA (Figure 1D) [40].

2.1.5. UV Resin

UV resins consist of a polymer monomer, a prepolymer, an active diluent, a photoinitiator, and a photosensitizer [61]. Under UV light of 250–300 nm wavelength, polymerization reaction commences immediately, and the scattered polymers combine into crosslinked polymers, turning from a liquid polymer resin to a solid 3D structure [62]. The conversion degree is affected by the monomer type, photoinitiator type, temperature, and light intensity [63]. UV resins are basic and essential materials for DLP and SLA [64]. The applications of UV resin include crowns, bridges, surgical guides, and other prostheses [65,66].
UV resins have high curing efficiencies, low energy consumption, low solvent emissions, low cost, and moderate curing conditions. Titanium dioxide and tartrazine lake affect the viscosity and curing thickness of the resin [67]. However, sunlight may change the shape and color of the product printed with SLA, because of the polymers’ sensitivity to UV light.
Table 2 lists the above-mentioned polymers currently used for 3D printing and their corresponding characteristics, applicable technologies, and clinical applications.

2.2. Metals

With their ideal mechanical and biological properties, metals have always been a good choice for dental restoration materials, mainly titanium (Ti) and cobalt–chromium (Co–Cr) alloys. However, their esthetic properties and higher cost render them less than optimal for use [68,69].

2.2.1. Ti and Its Alloys

Ti and its alloys are chemically active. Chemically stable oxide films, which are dense and strongly adherent, can be formed on their surfaces in air and other media to protect the substrates from corrosion. Their excellent mechanical properties, outstanding biocompatibility, and low density make them ideal materials for 3D printing [70]. Over the past 20 years, SLM, SLS, EBM, and DMLS have been introduced as promising manufacturing technologies for Ti and its alloys [71,72,73,74,75,76,77]. Owing to their excellent reproducibility, flexibility in design, relatively good resolution, and low-cost effectiveness, these 3D printing methods are widely used to manufacture implantable devices (i.e., dental implants and scaffolds) and prosthetic devices (i.e., dental crowns and denture frameworks) based on Ti and its alloys [69,78,79,80]. Studies have shown that Ti and Ti alloy samples printed by optimized 3D printing technologies have comparable biological properties to machine-milled products [72,81,82]. Yang et al. prepared 3D printed porous Ti6Al4V dental implants and applied a chitosan-based composite coating, which is non-toxic, is favorable for cell proliferation, and has good mechanical properties, thus contributing to the growth of new bone to support damaged bone [81]. Despite their unique advantages, Ti and its alloys can still corrode after implantation because of chewing forces, acidic environments, and so on [83]. Furthermore, 3D printed Ti alloys have shortcomings, such as high cost, poor wear resistance, easy embrittlement, and potential toxicity [4].
At present, medical devices based on Ti and its alloys have been used to fabricate personalized mandibular implants for maxillofacial surgery (Figure 2A) [84].

2.2.2. Co–Cr Alloys

The good mechanical properties and corrosion resistance of Co–Cr alloy allow it to be used to create products with high strength and high-temperature resistance [86]. Co–Cr alloy is hence considered an ideal material for fabricating prostheses and non-precious metal frameworks [82,87,88,89,90,91,92,93].
Co–Cr alloys can be printed by SLA, SLM, and DMLS, allowing rapid and accurate melting of the metal powders into layers using high-powered laser beams, and, consequently, the direct printing of products [69,73,86,91,94,95,96,97,98]. Numerous studies have shown that 3D printed Co–Cr alloy restoration not only meets the needs of dental practices (e.g., acceptable marginal clearance, higher biocompatibility, excellent microstructural homogeneity, and metal-ceramic bond strength) [69,95,99,100], but also eliminates the challenges associated with the milling process of Co–Cr alloys, such as shrinkage of the materials during the casting process. Owing to their improved properties, Co–Cr alloys are promising alternatives for the construction of dental restorations [24,96,101,102,103,104]. Furthermore, the post-production heat treatment of Co–Cr alloys made by SLM allows the effective release of the residual stresses, resulting in a homogeneous microstructure and improved mechanical properties [105]. Unfortunately, the printing parameters, melting temperature, powder adhesion, slagging, thermal stress accumulation, residual stress, and other constraints inherent to SLM technology limit the performance and hence the quality of the 3D printed products [69,95,106]. In prosthodontics, DMLS technology can be applied to manufacture Co–Cr alloy prostheses (Figure 2B) [85].

2.2.3. Others

Stainless-steel and magnesium alloys, both of which have good physical properties and biocompatibility, can be used as laser-selective melt-forming metals. Stainless steel has good sterilization capabilities and can be used for rapid prototyping of the dental implants and orthodontic components by SLM [107,108]. Magnesium (Mg) and its alloys may have important functional roles in the physiological system, due to their close mechanical properties to human bone tissue, their natural ion content, and their in vivo biodegradation characteristics in body fluids. They are usually used as an implant for orthopedic and dental treatment [109]. They also have good mechanical properties and biocompatibility and are promising for oral implantology. Julia et al. successfully modified magnesium implants through a high-pressure anodic oxidation process to obtain biodegradable magnesium-based implants. They are a new type of biomaterial with potential for dental applications and have the potential to make dental metal medical devices [110]. Zhang et al. evaluated magnesium alloys created using SLM technology to implant the fabricating materials for manufacturing porous structural implants, and the mechanical properties of the composite porous implants met the requirements of living for edentulous patients [111].
In addition, new materials, such as nanomaterials, which are less than 100 nm in diameter, have been introduced in dentistry. Adding nanometals and their oxides to other nanomaterials can enhance the antimicrobial, mechanical, and regenerative properties. Metal elements, such as Ag, Cu, Au, Ti, and Zn are antibacterial. Nanometals, such as Ag, Au, ZrO2, and TiO improve orthodontic adhesive’s compressive, tensile, and shear bond strengths. The question of whether nanometals have adverse effects on the oral system still needs further research [112,113,114].
The above-mentioned commonly used metals and their relevant information are listed in Table 3.

2.3. Ceramics

Ceramic materials have become ideal products for dental restorations because of their excellent mechanical properties, biocompatibility, good abrasion and corrosion resistance, and esthetic characteristics, similar to natural teeth [115,116,117]. The commonly used 3D printed ceramic materials in dentistry can be divided into glass, zirconia, and alumina ceramics, based on composition [118,119].

2.3.1. Glass Ceramics

Glass ceramics are a class of composite materials made via high-temperature sintering, molding, and heat treatment, combining crystalline phases. Dental glass ceramics were first proposed as a mica-based material by Malament and Grossman in the mid-1980s [120]. In the 1990s, scientists developed a stronger and more reliable microcrystalline glass for better dental restorations [121,122].
Glass ceramics have high mechanical strengths, low electrical conductivities, high dielectric constants, high chemical resistances, high thermal stabilities, and other superior properties [123]. Depending on their application, dental glass ceramics can be divided into bioactive dental glass ceramics (BDGCs), which have dental bonding capabilities and stimulate specific biological responses at the material/tissue interfaces, and restorative dental glass ceramics (RDGCs). BDGCs are suitable for hypersensitivity therapy, implant coatings, bone regeneration, and periodontal healing [76], while RDGCs can be applied to manufacturing inlays (Figure 3A), onlays, full crowns, partial crowns, bridges, and veneers [124]. Recently, lithography-based AM [125] and SLA [126] have been widely used to fabricate glass–ceramic restorations. Experiments have shown that glass ceramics with excellent mechanical properties can be effectively obtained using the SLA 3D printing process with appropriate parameters [127,128]. The exposure intensity in the SLA process affects the curing width and curing depth, as well as the quality of the final product [127].

2.3.2. Zirconia

Zirconia occupies an important position among oxide ceramics and is the current focus in dental materials, with a wide range of applications [131]. From a materials-science perspective, zirconia crystals are divided into three structures: monoclinic phase (m), tetragonal phase (t), and cubic phase (c), among which m is the most common form of zirconia at room temperature. However, when the temperature changes, these three structures can be interconverted [132,133]. Zirconia ceramics commonly used in dentistry are 3 mol% yttria-stabilized tetragonal zirconia polycrystal (3Y-TZP) [134], zirconia-toughened ceramics, partially stabilized zirconia ceramics, and nano-zirconia and alumina-composite ceramics. Each of these four types of zirconia ceramics contains a stable tetragonal phase and is toughened by the martensitic phase transformation of the t-phase with varying microstructure, resulting in different properties and processing techniques.
Zirconia is biocompatible and osteoconductive, and accommodates the surrounding soft tissues to facilitate bone formation [135]. Specifically, zirconium can reduce inflammatory responses, plaque accumulation, and bacterial population, thereby altering fibroblast adhesion and proliferation [136]. In terms of mechanical properties, zirconia is known as “ceramic steel”, as it has good toughness, ideal hardness, and high strength [137,138]. Zirconia plays an important role in prosthetic dentistry and is the raw material for root-canal piles, crown and bridge restorations, and implant abutments [139]. Zirconia, especially Y-TZP, has high toughness and strength, and has become the most durable among dental ceramics, but its esthetics are suboptimal, and the current Y-TZP zirconia suitable for anterior prostheses is constantly developing and improving [140,141].
SLA/SLS was introduced to fabricate zirconia, reducing the time of the restorative system fabrication and improving the treatment efficiency [142]. High internal stresses, cracks after sintering, and high-volume shrinkage may affect the mechanical properties and clinical suitability of 3D printed materials. Zirconia crown restoration manufactured by 3D printing is shown in Figure 3B [129].

2.3.3. Alumina

Aluminum oxide, or alumina (Al2O3), was first introduced in the 1970s as a ceramic material obtained by calcining aluminum hydroxide, which is often referred to as alumina trihydrate (ATH) [4,143,144]. Alumina includes monocrystalline and polycrystalline alumina in several forms: α, χ, η, δ, κ, θ, γ, and ρ. Alumina has high mechanical strength and excellent chemical stability. The higher purity usually results in higher strength [145]. To reduce the possibility of fracture, zirconia particles are often added to the matrix to obtain zirconia-toughened alumina (ZTA) ceramics with optimal strengths, high hardness, and good stabilities [146]. Cut or milled ceramics are preferred for obtaining flat surfaces. With the development of AM, 3D printed ceramics are gradually meeting complex shape requirements [147,148]. Alumina ceramics can be produced using inkjet printing technology. It is worth noting that the powder’s dimensions are critical, owing to the small size of the nozzle. Recently, it has been found that alumina ceramics can be used in FDM technology, while printing the smallest overhangs or cavities is not possible with milling technology [143,147,149]. One study reported that the physical and mechanical properties of alumina ceramic samples manufactured using SLA technology are comparable to those fabricated using conventional methods [150]. Another study used SLA technology to create precise alumina crown frameworks and have found that it can reduce the marginal gap between the crown framework and the preparation of the crown by refining the SLA manufacturing process. Figure 3C shows an alumina crown frame printed by SLA technology, placed on an all-ceramic crown preparation [130].
Some basic information on above-mentioned ceramics is presented in Table 4.

3. Four-dimensional Printing Materials

Four-dimensional printing materials refer to cell preparations containing biomaterials and bioactive ingredients, which create conditions for the applications of 4D printing in regenerative dentistry [151]. The process of 4D printing can be divided into two categories, according to whether scaffolds are needed, including scaffolds based on cells and the scaffold-free cell-based approach, directly printing cells to form new tissues [152,153,154]. Four-dimensional printing materials are composed of biological materials and living cells. After a certain period, the scaffold degrades and the cells grow in a predetermined way, which shows the requirement that bioinks have certain biocompatibility and self-preservation capacity. They can also be used in tissue regeneration engineering, such as bone and tooth tissue regeneration, with rapid gelation kinetics, tunable mechanical properties, good biocompatibility, and tissue adhesion [155]. Four-dimensional printing materials are summarized as synthetic polymers and natural biopolymers, and related cells and growth factors work with them.

3.1. Polymers

3.1.1. Synthetic Polymers

Synthetic polymers have been widely used in biological 3D printing [156], including PCL, PLA, and PLGA, as mentioned in the previous section [151,157]. In recent years, as the most widely used aliphatic polymer, PCL has gained an important position in the medical field and has potentially become a candidate material for craniofacial repair. Pluronic and poly (ethylene glycol) (PEG) is also a promising material in 4D printing. Pluronic is a block copolymer, consisting of two hydrophobic groups with a hydrophilic group between. Pluronic is capable of forming self-assembled gels at room temperature and is mobile at 10 °C [158]. Forming bioinks with other different polymers, PEG increases the mechanical properties of the origin structures [152,159].
Synthetic polymers are low-cost, mass-producible, chemically stable, and have appropriate degradation rates and photo-crosslinking capabilities [151,160,161]. Although synthetic polymers are not as biocompatible as natural polymers, they are tunable, have a lower gel temperature, and can withstand temperature and pH changes [151,152,162]. Even though, compared with natural polymers and bioceramics, aliphatic ethers have much lower absorption rates and less promotion on cell adhesion [163]. Adding synthetic polymers to natural ones gives more stable structures with tunable 4D printing properties, such as suitable porosity, surface area, and mechanical strength [152,163]. Moreover, synthetic polymers can now mimic the amphiphilic characteristics of natural polymers and introduce antimicrobial properties [164].

3.1.2. Natural Biopolymers

Biopolymer-based composite materials are biocompatible and biodegradable [165]. Natural polymers are usually important materials for fabricating bioinks, the advantages of which are that they can mimic the structure, self-assembly, and biocompatibility of the natural extracellular matrix (ECM). Common natural polymers, including hyaluronic acid (HA), collagen, agarose, chitosan, alginate, etc. have been introduce to 4D printing.
HA, a natural linear polymer, belongs to the ECM, and can usually be isolated from human or animal cartilage or connective tissue [166,167]. Presently, there are many HA-based bioinks, such as HA-based hydrogel bioinks. A recent study has shown that the printed products have high mechanical properties and stability [168]. Hydrogel has excellent biocompatibility, mechanical strength, biological and chemical properties, which made it the most common biological ink material [169]. Hydrogel simulates the microenvironment of the extracellular matrix, and is conducive to cell attachment, proliferation, and differentiation [170]. The modification of hydrogel by methacrylate improves the osteogenesis [168]. Photo-crosslinking HA bioink, chemical-crosslinking HA bioink, and HA-based double-crosslinking bioink have also been applied to 4D printing, and their rheological properties, mechanical properties, and the amount of cell adherence are improved after chemical modification and the addition of cell adhesion oligopeptides [171,172]. Associated with other polymers, such as gelatin and carboxymethylcellulose gel and thiol, HA-based bioinks have enhanced cell viability and biological stability [173,174]. HA promotes tissue repair and wound healing, facilitates recovery from dental surgery, and is also effective in periodontitis and gingivitis therapy [151,175].
Collagen, which also belongs to the ECM, can be used to make bioink, both separately or in combination [176]. Because its crosslinking or gel requires specific temperature and time, the effect of mixing collagen and other materials for printing is conducive to printing collagen separately [152,177]. For example, the combination of collagen and alginate is conducive to cell attachment, improves the mechanical properties of printing products, and is conducive to the application in cartilage tissue engineering [178]. Collagen, combined with gelatin, can improve biological activity and rheological properties [179].
Agarose is a marine polysaccharide whose main chain is composed of disaccharides. Although it has good gel properties, mechanical properties, and biocompatibility, it has limited ability to support cell growth. In order to enhance its performance, biological materials, such as collagen, fibrinogen, and sodium alginate, are often added to agarose gel as agar-based bioink [178,180]. The results show that these materials can promote cell growth and enhance the mechanical properties of bone tissue. In order to maintain the stability of the printing structure and enhance the various properties of cells, chemical treatment, such as carboxylation, is required [178,181,182,183].
Chitosan, a natural biopolymer, plays an important role in dentistry and has been widely used in prosthodontics, oral implantology, and endodontics [184]. Chitosan is not easily soluble in water, but is soluble in acids with pH lower than 6.2. Although it has good biocompatibility and renewable ability, its low mechanical strength limits its application in hard tissue regeneration, and its stability and mechanical properties can be enhanced by adding calcium ion (Ca2+) [185].
In dentistry, alginate, which is a natural polysaccharide, is generally used to obtain conventional impressions. The alginate biopolymers are suitable for making bioinks, because they can adsorb other molecules and water, which diffuse outwards [186]. Alginate-based bioinks can print 3D structures, both with or without chondrocytes, at the same time. Microchannels printed with sodium alginate-based hydrogel materials have higher strength. Polymers such as PCL, gelatin, and poloxamer can also be mixed with sodium alginate for printing [187,188,189,190].
Cellulose, derived from plant fibers, can be used to make tissue engineering scaffolds and wound dressings, and plays a significant role in drug delivery [191]. It has hydrophilic, as well as good insulating and anti-electrostatic, properties, making it the most abundant, renewable, and stable natural linear polymer, widely used in industrial and biomedical fields [180,192,193]. Because of its biocompatibility, different types of cellulosic materials, such as nanocrystals (CNCs), cellulose nanofibers, and nanofibers (CNFs), are often added to the hydrogels to improve the viscosity of bioinks and ensure higher printing accuracy [173,194,195]. The combination of cellulose with alginate, hyaluronic acid, PLA, and other polymers to produce printable hydrogels is the key to the use of cellulose in tissue engineering to print 3D scaffolds and other structures [195]. Adding lignocellulosic fibers to other biopolymers can significantly improve the mechanical strength of biopolymers [180].
The basic information about the above-mentioned natural biopolymers is listed in Table 5.

3.2. Cells

Cells are essential parts in bioinks for guiding tissue and organ regeneration. Mesenchymal stem cells (MSCs) with immune adjustability properties and osteoblasts are ideal cells for 4D printing. Cells can be printed with or without base biomaterials that provide scaffolds. Cell aggregates and spheroids can be directly introduced as 4D printing materials. Cell aggregates refer to the bioinks based merely on cells, without any additional biomaterials.
Multiple cells have been applied to 4D printing in dental and maxilla-facial surgery (Table 6 [201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222]), including both dental and non-dental cells. Dental pulp stem cells (DPSCs), stem cells from the apical papillae (SCAPs), periodontal ligament stem cells (PDLSCs), dental follicle cells (DFCs), gingival fibroblasts (GFs), and stem cells from human exfoliated deciduous teeth (SHEDs) have been applied to 4D printing in tooth regeneration. Non-dental cells include bone marrow stem cells, pre-osteoblasts MC3T3-E1 cells, and MSCs from gingival tissue. The adjustment of differentiation of DPSCs can be achieved by bioactive ingredients, such as concentrated growth factors [223,224]. Similar to DPSCs, SCAPs show a strong mineralization capacity, with a high expression of dentin sialophosphoprotein, which reflects the odontogenic differentiation [225]. The 4D printing of periodontal ligament cells can improve biocompatibility, which can be applied to periodontal therapy, especially pulp revascularization [226,227]. DFCs, which surround the tooth germ, are derived from mesenchymal connective tissue, and have the best proliferation capacity among all the mentioned cells and possible differentiation to osteoblasts and cementoblasts [211]. GFs, with different phenotypes, have therapeutic potential for regenerative medicine. The functionalized base biomaterials trigger the differentiation and proliferation of the GFs. In a clinical study, it has been found that culture GFs significantly reduce vertical pocket depth, compared with β-calcium triphosphate and collagen membrane in the treatment of intrabony periodontal defects [228]. The biocompatibility between hybrid polymers membrane and GFs has shown that GFs are promising cell materials for bioinks [229]. Being extracted from the lost deciduous tooth, SHEDs can promote bone formation and produce dentin in in vivo experiments and are able to improve the expression of HA compared with human fibroblasts, which demonstrated that SHEDs could provide a new strategy for wound healing [230,231].
Four-dimensional printing technology applied in dental and cranial facial surgery consists of the bone regeneration, periodontal complex, dentin-pulp complex (Figure 4), pulps, neural regeneration, whole tooth regeneration, and cartilage, which is mainly applied in the temporomandibular joint [232,233]. However, the segmentation during 3D printing would be a problem to be solved, resulting in the instability of the printing process, leading to the clog of the nozzle and the printability of the products [234]. Currently, researchers have studied the immunological properties of different biomarkers on the surface of stem cells’ membranes. The correlation between biomaterials and stem cells is, today, a research hotspot, and the surrounding factors, such as local pH and metal ions’ impacts on the stem cells, are also required for deeper investigation.

3.3. Growth Factors

Growth factor is a soluble signaling molecule that binds stably to ECM and controls a variety of cellular responses, such as cell growth, value addition, and differentiation. The selective release of growth factors is essential for the bioactivity and tissue regenerative capacity of bioinks. Suitable materials, such as natural and synthetic polymers, are usually selected to incorporate growth factors into sponges, micro/nanoparticles, and hydrogels [235,236]. Recently, the incorporation of cells and growth factors into bioinks to promote bone tissue regeneration has been achieved.
Stromal-derived factor 1 (SDF-1) has a role in regulating cell migration and cell growth, and promotes vascular regeneration and bone regeneration [237]. It can be incorporated into hydrogels for preservation. The addition of SDF-1 to PLA promotes the migration of bone regeneration cells into the printed scaffold and promotes osteogenesis [238]. Bone morphogenetic proteins (BMP) have the ability to induce bone regeneration and bone healing. Bone scaffolds printed with bioinks, and supplemented with BMP-2 and transforming growth factor-beta1 (TGF-β1), stimulated bioactivity and significantly improved bone regeneration [239]. The co-assembly of hydrogels, SDF-1 and BMP-2, achieved a controlled release of both growth factors and promoted the regeneration of periodontal bone tissue [240]. BMP-7 was added to PCL and β-TCP to print bone scaffolds to repair bone defects in animal models, and the results showed better good tissue ingrowth and bone regeneration [241]. BMP-7 was added to PLA to print bone scaffolds for repairing defects and analyzed for bone regeneration. The results showed that BMP-7 was more effective in bone regeneration treatment and the newly generated bone had a porous structure [238]. Vascular endothelial growth factor (VEGF) regulates cell migration, angiogenesis, and bone regeneration [239]. Growth factors (e.g., VEGF and BMP-2) can also bind to nanoparticles to control the rate of growth factor release through nanoparticle movement and integration [242].
Above-mentioned growth factors are listed in Table 7.

4. Futural Prospects and Conclusions Marks

In summary, with digital technology’s development, 3D printing and 4D printing have been adopted in oral clinical medicine. Due to the uniqueness in individual personalization, 3D printing has gained wide application in dentistry. Traditionally applied materials include polymers, metals, and ceramics. Three-dimensional bioprinting, also known as 4D printing, has been developed to improve the therapeutic effects in dento-maxillo-facial surgery. Four-dimensional printing introduced both synthetic and bio-natural materials as raw materials and stem cells, which accelerates the healing period after surgery and achieves bioactive repairing after surgeries.
Futural studies will continue to focus on the regulation of both biological and physical properties of printed products, with less possibility of inflammation after surgeries and induction of early bio-reaction. Presently, to improve the properties of dental materials, researchers are now studying the correlation between each material and working on the investigation of composite materials with desired properties and characteristics. Hydroxyapatite has been corporate into polymers to improve osteoinductivity. Different surface modifications have been studied on the metal to improve the antibacterial properties of titanium and its alloy, and protective coatings were manufactured on the surface of magnesium alloy to adjust the degradation rate. Also, to optimize the bioactivities of the 4D printed products, the biological actions, including stem cell proliferation and differentiation regulation, and the interaction of biomaterials with stem cells have been extensively studied.
It is believed that, with the rapid development of material investigations, multiple ideal novel materials and techniques, with better biocompatibility and bioactivity, would be applied clinically to contribute to the prosperity of dental practices.

Author Contributions

Conceptualization, H.C. and H.-B.J.; software, X.X. and X.L.; validation, H.C. and J.-S.K.; investigation, M.Z. and Q.J.; writing—original draft preparation, H.C., X.X. and X.L.; writing—review and editing, H.C., X.X., X.L., M.Z. and Q.J.; supervision, J.-S.K. and H.-B.J.; project administration, J.-S.K. and H.-B.J.; 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.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bertrand, B.; Colard, T.; Lacoche, C.; Salome, J.F.; Vatteoni, S. An original case of tin dental fillings from 18th century northern France. J. Dent. Res. 2009, 88, 198–200. [Google Scholar] [CrossRef]
  2. Mutter, J.; Naumann, J.; Walach, H.; Daschner, F. Amalgam risk assessment with coverage of references up to 2005. Gesundheitswesen 2005, 67, 204–216. [Google Scholar] [CrossRef]
  3. Lin, L.; Fang, Y.; Liao, Y.; Chen, G.; Gao, C.; Zhu, P. 3D Printing and Digital Processing Techniques in Dentistry: A Review of Literature. Adv. Eng. Mater. 2019, 21, 1801013. [Google Scholar] [CrossRef]
  4. Khorsandi, D.; Fahimipour, A.; Abasian, P.; Saber, S.S.; Seyedi, M.; Ghanavati, S.; Ahmad, A.; De Stephanis, A.A.; Taghavinezhaddilami, F.; Leonova, A.; 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]
  5. Carlotti, M.; Mattoli, V. Functional Materials for Two-Photon Polymerization in Microfabrication. Small 2019, 15, e1902687. [Google Scholar] [CrossRef]
  6. Tian, Y.; Chen, C.; Xu, X.; Wang, J.; Hou, X.; Li, K.; Lu, X.; Shi, H.; Lee, E.S.; Jiang, H.B. A Review of 3D Printing in Dentistry: Technologies, Affecting Factors, and Applications. Scanning 2021, 2021, 9950131. [Google Scholar] [CrossRef]
  7. Vukicevic, M.; Mosadegh, B.; Min, J.K.; Little, S.H. Cardiac 3D Printing and its Future Directions. JACC Cardiovasc. Imaging 2017, 10, 171–184. [Google Scholar] [CrossRef]
  8. Alharbi, N.; Alharbi, S.; Cuijpers, V.; Osman, R.B.; Wismeijer, D. Three-dimensional evaluation of marginal and internal fit of 3D-printed interim restorations fabricated on different finish line designs. J. Prosthodont. Res. 2018, 62, 218–226. [Google Scholar] [CrossRef]
  9. Grossin, D.; Montón, A.; Navarrete-Segado, P.; Özmen, E.; Urruth, G.; Maury, F.; Maury, D.; Frances, C.; Tourbin, M.; Lenormand, P.; et al. A review of additive manufacturing of ceramics by powder bed selective laser processing (sintering/melting): Calcium phosphate, silicon carbide, zirconia, alumina, and their composites. Open Ceram. 2021, 5, 100073. [Google Scholar] [CrossRef]
  10. Vaz, V.M.; Kumar, L. 3D Printing as a Promising Tool in Personalized Medicine. AAPS PharmSciTech 2021, 22, 49. [Google Scholar] [CrossRef]
  11. Zhu, W.; Ma, X.; Gou, M.; Mei, D.; Zhang, K.; Chen, S. 3D printing of functional biomaterials for tissue engineering. Curr. Opin. Biotechnol. 2016, 40, 103–112. [Google Scholar] [CrossRef] [PubMed]
  12. Chu, H.; Yang, W.; Sun, L.; Cai, S.; Yang, R.; Liang, W.; Yu, H.; Liu, L. 4D Printing: A Review on Recent Progresses. Micromachines 2020, 11, 796. [Google Scholar] [CrossRef]
  13. Mallakpour, S.; Tabesh, F.; Hussain, C.M. 3D and 4D printing: From innovation to evolution. Adv. Colloid. Interface Sci. 2021, 294, 102482. [Google Scholar] [CrossRef]
  14. Arif, Z.U.; Khalid, M.Y.; Ahmed, W.; Arshad, H. A review on four-dimensional (4D) bioprinting in pursuit of advanced tissue engineering applications. Bioprinting 2022, 27, e00203. [Google Scholar] [CrossRef]
  15. Wan, X.; Luo, L.; Liu, Y.; Leng, J. Direct Ink Writing Based 4D Printing of Materials and Their Applications. Adv. Sci. 2020, 7, 2001000. [Google Scholar] [CrossRef] [PubMed]
  16. Ashammakhi, N.; Ahadian, S.; Zengjie, F.; Suthiwanich, K.; Lorestani, F.; Orive, G.; Ostrovidov, S.; Khademhosseini, A. Advances and Future Perspectives in 4D Bioprinting. Biotechnol. J. 2018, 13, e1800148. [Google Scholar] [CrossRef] [PubMed]
  17. Murphy, S.V.; Atala, A. 3D bioprinting of tissues and organs. Nat. Biotechnol. 2014, 32, 773–785. [Google Scholar] [CrossRef]
  18. Dey, M.; Ozbolat, I.T. 3D bioprinting of cells, tissues and organs. Sci. Rep. 2020, 10, 14023. [Google Scholar] [CrossRef]
  19. Mandrycky, C.; Wang, Z.; Kim, K.; Kim, D.H. 3D bioprinting for engineering complex tissues. Biotechnol. Adv. 2016, 34, 422–434. [Google Scholar] [CrossRef]
  20. Ostrovidov, S.; Salehi, S.; Costantini, M.; Suthiwanich, K.; Ebrahimi, M.; Sadeghian, R.B.; Fujie, T.; Shi, X.; Cannata, S.; Gargioli, C.; et al. 3D Bioprinting in Skeletal Muscle Tissue Engineering. Small 2019, 15, e1805530. [Google Scholar] [CrossRef]
  21. Ozbolat, I.T.; Hospodiuk, M. Current advances and future perspectives in extrusion-based bioprinting. Biomaterials 2016, 76, 321–343. [Google Scholar] [CrossRef] [PubMed]
  22. Derakhshanfar, S.; Mbeleck, R.; Xu, K.; Zhang, X.; Zhong, W.; Xing, M. 3D bioprinting for biomedical devices and tissue engineering: A review of recent trends and advances. Bioact. Mater. 2018, 3, 144–156. [Google Scholar] [CrossRef] [PubMed]
  23. Revilla-León, M.; Sadeghpour, M.; Özcan, M. An update on applications of 3D printing technologies used for processing polymers used in implant dentistry. Odontology 2020, 108, 331–338. [Google Scholar] [CrossRef] [PubMed]
  24. Pillai, S.; Upadhyay, A.; Khayambashi, P.; Farooq, I.; Sabri, H.; Tarar, M.; Lee, K.T.; Harb, I.; Zhou, S.; Wang, Y.; et al. Dental 3D-Printing: Transferring Art from the Laboratories to the Clinics. Polymers 2021, 13, 157. [Google Scholar] [CrossRef]
  25. Woodruff, M.A.; Hutmacher, D.W. The return of a forgotten polymer—Polycaprolactone in the 21st century. Prog. Polym. Sci. 2010, 35, 1217–1256. [Google Scholar] [CrossRef]
  26. Liu, F.; Vyas, C.; Poologasundarampillai, G.; Pape, I.; Hinduja, S.; Mirihanage, W.; Bartolo, P. Structural Evolution of PCL during Melt Extrusion 3D Printing. Macromol. Mater. Eng. 2018, 303, 1700494. [Google Scholar] [CrossRef]
  27. Malikmammadov, E.; Tanir, T.E.; Kiziltay, A.; Hasirci, V.; Hasirci, N. PCL and PCL-based materials in biomedical applications. J. Biomater. Sci. Polym. Ed. 2018, 29, 863–893. [Google Scholar] [CrossRef]
  28. Bruyas, A.; Lou, F.; Stahl, A.M.; Gardner, M.; Maloney, W.; Goodman, S.; Yang, Y.P. Systematic characterization of 3D-printed PCL/β-TCP scaffolds for biomedical devices and bone tissue engineering: Influence of composition and porosity. J. Mater. Res. 2018, 33, 1948–1959. [Google Scholar] [CrossRef]
  29. Siddiqui, N.; Asawa, S.; Birru, B.; Baadhe, R.; Rao, S. PCL-Based Composite Scaffold Matrices for Tissue Engineering Applications. Mol. Biotechnol. 2018, 60, 506–532. [Google Scholar] [CrossRef]
  30. Arif, Z.U.; Khalid, M.Y.; Noroozi, R.; Hossain, M.; Shi, H.H.; Tariq, A.; Ramakrishna, S.; Umer, R. Additive manufacturing of sustainable biomaterials for biomedical applications. Asian J. Pharm. Sci. 2023, 100812, in press. [Google Scholar] [CrossRef]
  31. Arif, Z.U.; Khalid, M.Y.; Noroozi, R.; Sadeghianmaryan, A.; Jalalvand, M.; Hossain, M. Recent advances in 3D-printed polylactide and polycaprolactone-based biomaterials for tissue engineering applications. Int. J. Biol. Macromol. 2022, 218, 930–968. [Google Scholar] [CrossRef] [PubMed]
  32. Shim, J.-H.; Yoon, M.-C.; Jeong, C.-M.; Jang, J.; Jeong, S.-I.; Cho, D.-W.; Huh, J.-B. Efficacy of rhBMP-2 loaded PCL/PLGA/ β -TCP guided bone regeneration membrane fabricated by 3D printing technology for reconstruction of calvaria defects in rabbit. Biomed. Mater. 2014, 9, 065006. [Google Scholar] [CrossRef] [PubMed]
  33. Shim, J.-H.; Jeong, J.-H.; Won, J.-Y.; Bae, J.-H.; Ahn, G.; Jeon, H.; Yun, W.-S.; Bae, E.-B.; Choi, J.-W.; Lee, S.-H.; et al. Porosity effect of 3D-printed polycaprolactone membranes on calvarial defect model for guided bone regeneration. Biomed. Mater. 2017, 13, 015014. [Google Scholar] [CrossRef] [PubMed]
  34. Berner, A.; Woodruff, M.A.; Lam, C.X.F.; Arafat, M.T.; Saifzadeh, S.; Steck, R.; Ren, J.; Nerlich, M.; Ekaputra, A.K.; Gibson, I.; et al. Effects of scaffold architecture on cranial bone healing. Int. J. Oral Maxillofac. Surg. 2014, 43, 506–513. [Google Scholar] [CrossRef] [PubMed]
  35. Bae, E.-B.; Park, K.-H.; Shim, J.-H.; Chung, H.-Y.; Choi, J.-W.; Lee, J.-J.; Kim, C.-H.; Jeon, H.-J.; Kang, S.-S.; Huh, J.-B. Efficacy of rhBMP-2 Loaded PCL/β-TCP/bdECM Scaffold Fabricated by 3D Printing Technology on Bone Regeneration. BioMed Res. Int. 2018, 2018, 2876135. [Google Scholar] [CrossRef]
  36. Jiao, Z.; Luo, B.; Xiang, S.; Ma, H.; Yu, Y.; Yang, W. 3D printing of HA/PCL composite tissue engineering scaffolds. Adv. Ind. Eng. Polym. Res. 2019, 2, 196–202. [Google Scholar] [CrossRef]
  37. 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 beta-TCP Powder for Alveolar Bone Augmentation in a Beagle Defect Model. Materials 2018, 11, 238. [Google Scholar] [CrossRef]
  38. Giannetti, L.; Apponi, R.; Mordini, L.; Presti, S.; Breschi, L.; Mintrone, F. The occlusal precision of milled versus printed provisional crowns. J. Dent. 2022, 117, 103924. [Google Scholar] [CrossRef]
  39. Maia-Pinto, M.O.C.; Brochado, A.C.B.; Teixeira, B.N.; Sartoretto, S.C.; Uzeda, M.J.; Alves, A.; Alves, G.G.; Calasans-Maia, M.D.; Thire, R. Biomimetic Mineralization on 3D Printed PLA Scaffolds: On the Response of Human Primary Osteoblasts Spheroids and In Vivo Implantation. Polymers 2020, 13, 74. [Google Scholar] [CrossRef]
  40. Naseri, E.; Butler, H.; MacNevin, W.; Ahmed, M.; Ahmadi, A. Low-temperature solvent-based 3D printing of PLGA: A parametric printability study. Drug. Dev. Ind. Pharm. 2020, 46, 173–178. [Google Scholar] [CrossRef]
  41. Rickman, L.J.; Padipatvuthikul, P.; Satterthwaite, J.D. Contemporary denture base resins: Part 1. Dent. Update 2012, 39, 25–38. [Google Scholar] [CrossRef] [PubMed]
  42. Frazer, R.Q.; Byron, R.T.; Osborne, P.B.; West, K.P. PMMA: An Essential Material in Medicine and Dentistry. J. Long-Term Eff. Med. Implant. 2005, 15, 629–639. [Google Scholar] [CrossRef] [PubMed]
  43. Zafar, M.S. Prosthodontic Applications of Polymethyl Methacrylate (PMMA): An Update. Polymers 2020, 12, 2299. [Google Scholar] [CrossRef] [PubMed]
  44. Pituru, S.M.; Greabu, M.; Totan, A.; Imre, M.; Pantea, M.; Spinu, T.; Tancu, A.M.C.; Popoviciu, N.O.; Stanescu, I.-I.; Ionescu, E. A Review on the Biocompatibility of PMMA-Based Dental Materials for Interim Prosthetic Restorations with a Glimpse into Their Modern Manufacturing Techniques. Materials 2020, 13, 2894. [Google Scholar] [CrossRef]
  45. Wieckiewicz, M.; Opitz, V.; Richter, G.; Boening, K.W. Physical Properties of Polyamide-12 versus PMMA Denture Base Material. BioMed Res. Int. 2014, 2014, 150298. [Google Scholar] [CrossRef]
  46. Protopapa, P.; Kontonasaki, E.; Bikiaris, D.; Paraskevopoulos, K.M.; Koidis, P. Reinforcement of a PMMA resin for fixed interim prostheses with nanodiamonds. Dent. Mater. J. 2011, 30, 222–231. [Google Scholar] [CrossRef]
  47. Yang, J.; Li, W.; Mu, B.; Xu, H.; Hou, X.; Yang, Y. 3D printing of toughened enantiomeric PLA/PBAT/PMMA quaternary system with complete stereo-complexation: Compatibilizer architecture effects. Polymer 2022, 242, 124590. [Google Scholar] [CrossRef]
  48. Marin, E.; Mukai, M.; Boschetto, F.; Sunthar, T.P.M.; Adachi, T.; Zhu, W.; Rondinella, A.; Lanzutti, A.; Kanamura, N.; Yamamoto, T.; et al. Production of antibacterial PMMA-based composites through stereolithography. Mater. Today Commun. 2022, 32, 103943. [Google Scholar] [CrossRef]
  49. Polzin, C.; Spath, S.; Seitz, H. Characterization and evaluation of a PMMA-based 3D printing process. Rapid Prototyp. J. 2013, 19, 37–43. [Google Scholar] [CrossRef]
  50. Grémare, A.; Guduric, V.; Bareille, R.; Heroguez, V.; Latour, S.; L’heureux, N.; Fricain, J.-C.; Catros, S.; Le Nihouannen, D. Characterization of printed PLA scaffolds for bone tissue engineering. J. Biomed. Mater. Res. Part A 2018, 106, 887–894. [Google Scholar] [CrossRef]
  51. Du, X.; Song, Y.; Xuan, X.; Chen, S.; Wu, X.; Jiang, H.B.; Lee, E.-S.; Wang, X. Characterization of a Bioresorbable Magnesium-Reinforced PLA-Integrated GTR/GBR Membrane as Dental Applications. Scanning 2020, 2020, 6743195. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, H.-C.; E, L.-L.; Wang, D.-S.; Su, F.; Wu, X.; Shi, Z.-P.; Lv, Y.; Wang, J.-Z. Reconstruction of Alveolar Bone Defects Using Bone Morphogenetic Protein 2 Mediated Rabbit Dental Pulp Stem Cells Seeded on Nano-Hydroxyapatite/Collagen/Poly(L-lactide). Tissue Eng. Part A 2011, 17, 2417–2433. [Google Scholar] [CrossRef] [PubMed]
  53. Deng, K.; Chen, H.; Zhao, Y.; Zhou, Y.; Wang, Y.; Sun, Y. Evaluation of adaptation of the polylactic acid pattern of maxillary complete dentures fabricated by fused deposition modelling technology: A pilot study. PLoS ONE 2018, 13, e0201777. [Google Scholar] [CrossRef] [PubMed]
  54. Hoshiba, T.; Gong, J. Fabrication of cell-derived decellularized matrices on three-dimensional (3D)-printed biodegradable polymer scaffolds. Microsyst. Technol. 2018, 24, 613–617. [Google Scholar] [CrossRef]
  55. Aguado-Maestro, I.; De Frutos-Serna, M.; Gonzalez-Nava, A.; Merino-De Santos, A.B.; Garcia-Alonso, M. Are the common sterilization methods completely effective for our in-house 3D printed biomodels and surgical guides? Injury 2021, 52, 1341–1345. [Google Scholar] [CrossRef]
  56. Jager, A.; Donato, R.K.; Perchacz, M.; Donato, K.Z.; Stary, Z.; Konefal, R.; Serkis-Rodzen, M.; Raucci, M.G.; Fuentefria, A.M.; Jager, E. Human metabolite-derived alkylsuccinate/dilinoleate copolymers: From synthesis to application. J. Mater. Chem. B. 2020, 8, 9980–9996. [Google Scholar] [CrossRef]
  57. Guo, T.; Holzberg, T.R.; Lim, C.G.; Gao, F.; Gargava, A.; Trachtenberg, J.E.; Mikos, A.G.; Fisher, J.P. 3D printing PLGA: A quantitative examination of the effects of polymer composition and printing parameters on print resolution. Biofabrication 2017, 9, 024101. [Google Scholar] [CrossRef]
  58. Mironov, A.V.; Grigoryev, A.M.; Krotova, L.I.; Skaletsky, N.N.; Popov, V.K.; Sevastianov, V.I. 3D printing of PLGA scaffolds for tissue engineering. J. Biomed. Mater. Res. Part A 2017, 105, 104–109. [Google Scholar] [CrossRef]
  59. Lai, Y.; Li, Y.; Cao, H.; Long, J.; Wang, X.; Li, L.; Li, C.; Jia, Q.; Teng, B.; Tang, T.; et al. Osteogenic magnesium incorporated into PLGA/TCP porous scaffold by 3D printing for repairing challenging bone defect. Biomaterials 2019, 197, 207–219. [Google Scholar] [CrossRef]
  60. Ge, Z.; Tian, X.; Heng, B.C.; Fan, V.; Yeo, J.F.; Cao, T. Histological evaluation of osteogenesis of 3D-printed poly-lactic-co-glycolic acid (PLGA) scaffolds in a rabbit model. Biomed. Mater. 2009, 4, 021001. [Google Scholar] [CrossRef]
  61. Albuquerque, P.P.; Moreira, A.D.; Moraes, R.R.; Cavalcante, L.M.; Schneider, L.F. Color stability, conversion, water sorption and solubility of dental composites formulated with different photoinitiator systems. J. Dent. 2013, 41 (Suppl. S3), e67–e72. [Google Scholar] [CrossRef] [PubMed]
  62. Zhang, B.; Li, S.; Hingorani, H.; Serjouei, A.; Larush, L.; Pawar, A.A.; Goh, W.H.; Sakhaei, A.H.; Hashimoto, M.; Kowsari, K.; et al. Highly stretchable hydrogels for UV curing based high-resolution multimaterial 3D printing. J. Mater. Chem. B. 2018, 6, 3246–3253. [Google Scholar] [CrossRef] [PubMed]
  63. Kim, G.T.; Go, H.B.; Yu, J.H.; Yang, S.Y.; Kim, K.M.; Choi, S.H.; Kwon, J.S. Cytotoxicity, Colour Stability and Dimensional Accuracy of 3D Printing Resin with Three Different Photoinitiators. Polymers 2022, 14, 979. [Google Scholar] [CrossRef] [PubMed]
  64. Ishida, Y.; Miyasaka, T. Dimensional accuracy of dental casting patterns created by 3D printers. Dent. Mater. J. 2016, 35, 250–256. [Google Scholar] [CrossRef]
  65. Gok, T.; Capar, I.D.; Akcay, I.; Keles, A. Evaluation of Different Techniques for Filling Simulated C-shaped Canals of 3-dimensional Printed Resin Teeth. J. Endod. 2017, 43, 1559–1564. [Google Scholar] [CrossRef] [PubMed]
  66. Vyas, A.; Garg, V.; Ghosh, S.B.; Bandyopadhyay-Ghosh, S. Photopolymerizable resin-based 3D printed biomedical composites: Factors affecting resin viscosity. Mater. Today 2022, 62, 1435–1439. [Google Scholar] [CrossRef]
  67. Chang, W.-T.; Hsu, H.-J.; Jiang, C.-P.; Lee, S.-Y.; Lin, Y.-M. Effects of titanium dioxide and tartrazine lake on Z-axis resolution and physical properties of resins printed by visible-light 3D printers. Rapid Prototyp. J. 2018, 24, 160–165. [Google Scholar] [CrossRef]
  68. Sulaiman, T.A. Materials in digital dentistry-A review. J. Esthet. Restor. Dent. 2020, 32, 171–181. [Google Scholar] [CrossRef]
  69. Revilla-Leon, M.; Sadeghpour, M.; Ozcan, M. A Review of the Applications of Additive Manufacturing Technologies Used to Fabricate Metals in Implant Dentistry. J. Prosthodont. 2020, 29, 579–593. [Google Scholar] [CrossRef]
  70. Hatamleh, M.M.; Wu, X.; Alnazzawi, A.; Watson, J.; Watts, D. Surface characteristics and biocompatibility of cranioplasty titanium implants following different surface treatments. Dent. Mater. 2018, 34, 676–683. [Google Scholar] [CrossRef]
  71. Cordeiro, J.M.; Beline, T.; Ribeiro, A.L.R.; Rangel, E.C.; da Cruz, N.C.; Landers, R.; Faverani, L.P.; Vaz, L.G.; Fais, L.M.G.; Vicente, F.B.; et al. Development of binary and ternary titanium alloys for dental implants. Dent. Mater. 2017, 33, 1244–1257. [Google Scholar] [CrossRef] [PubMed]
  72. Harada, Y.; Ishida, Y.; Miura, D.; Watanabe, S.; Aoki, H.; Miyasaka, T.; Shinya, A. Mechanical Properties of Selective Laser Sintering Pure Titanium and Ti-6Al-4V, and Its Anisotropy. Materials 2020, 13, 5081. [Google Scholar] [CrossRef] [PubMed]
  73. Bae, E.J.; Kim, H.Y.; Kim, W.C.; Kim, J.H. In vitro evaluation of the bond strength between various ceramics and cobalt-chromium alloy fabricated by selective laser sintering. J. Adv. Prosthodont. 2015, 7, 312–316. [Google Scholar] [CrossRef]
  74. da Costa Valente, M.L.; de Oliveira, T.T.; Kreve, S.; Batalha, R.L.; de Oliveira, D.P.; Pauly, S.; Bolfarini, C.; Bachmann, L.; Dos Reis, A.C. Analysis of the mechanical and physicochemical properties of Ti-6Al-4 V discs obtained by selective laser melting and subtractive manufacturing method. J. Biomed. Mater. Res. B Appl. Biomater. 2021, 109, 420–427. [Google Scholar] [CrossRef]
  75. Kristensen, F.K.; Sverre, J.M.; Bustad, S. Pdb21 a Cost-Utility Analysis of Insulin Glargine (Lantus®) in the Treatment of Patients with Type 1 Diabetes. Value Health 2003, 6, 682. [Google Scholar] [CrossRef]
  76. Holand, W.; Rheinberger, V.; Apel, E.; van ‘t Hoen, C.; Holand, M.; Dommann, A.; Obrecht, M.; Mauth, C.; Graf-Hausner, U. Clinical applications of glass-ceramics in dentistry. J. Mater. Sci. Mater. Med. 2006, 17, 1037–1042. [Google Scholar] [CrossRef] [PubMed]
  77. Ramakrishnaiah, R.; Al Kheraif, A.A.; Mohammad, A.; Divakar, D.D.; Kotha, S.B.; Celur, S.L.; Hashem, M.I.; Vallittu, P.K.; Rehman, I.U. Preliminary fabrication and characterization of electron beam melted Ti-6Al-4V customized dental implant. Saudi J. Biol. Sci. 2017, 24, 787–796. [Google Scholar] [CrossRef] [PubMed]
  78. Sidambe, A.T. Biocompatibility of Advanced Manufactured Titanium Implants—A Review. Materials 2014, 7, 8168–8188. [Google Scholar] [CrossRef]
  79. Kalman, L.; Hosein, Y.; Chimel, T. Workflow Development of a 3D Printed Novel Implant Abutment. 3D Print. Addit. Manuf. 2019, 6, 235–237. [Google Scholar] [CrossRef]
  80. Santos-Coquillat, A.; Mohedano, M.; Martinez-Campos, E.; Arrabal, R.; Pardo, A.; Matykina, E. Bioactive multi-elemental PEO-coatings on titanium for dental implant applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 97, 738–752. [Google Scholar] [CrossRef]
  81. Kim, J.H.; Kim, M.Y.; Knowles, J.C.; Choi, S.; Kang, H.; Park, S.H.; Park, S.M.; Kim, H.W.; Park, J.T.; Lee, J.H.; et al. Mechanophysical and biological properties of a 3D-printed titanium alloy for dental applications. Dent. Mater. 2020, 36, 945–958. [Google Scholar] [CrossRef] [PubMed]
  82. Revilla-León, M.; Meyer, M.J.; Özcan, M. Metal additive manufacturing technologies: Literature review of current status and prosthodontic applications. Int. J. Comput. Dent. 2019, 22, 55–67. [Google Scholar] [PubMed]
  83. Sarao, T.P.S.; Singh, H.; Singh, H. Enhancing Biocompatibility and Corrosion Resistance of Ti-6Al-4V Alloy by Surface Modification Route. J. Therm. Spray Technol. 2018, 27, 1388–1400. [Google Scholar] [CrossRef]
  84. Popov, V.V., Jr.; Muller-Kamskii, G.; Kovalevsky, A.; Dzhenzhera, G.; Strokin, E.; Kolomiets, A.; Ramon, J. Design and 3D-printing of titanium bone implants: Brief review of approach and clinical cases. Biomed. Eng. Lett. 2018, 8, 337–344. [Google Scholar] [CrossRef]
  85. Available online: https://cn.3dsystems.com/materials/cocr-cobalt-chrome (accessed on 8 May 2023).
  86. Konieczny, B.; Szczesio-Wlodarczyk, A.; Sokolowski, J.; Bociong, K. Challenges of Co-Cr Alloy Additive Manufacturing Methods in Dentistry-The Current State of Knowledge (Systematic Review). Materials 2020, 13, 3524. [Google Scholar] [CrossRef]
  87. Oilo, M.; Nesse, H.; Lundberg, O.J.; Gjerdet, N.R. Mechanical properties of cobalt-chromium 3-unit fixed dental prostheses fabricated by casting, milling, and additive manufacturing. J. Prosthet. Dent. 2018, 120, 156.e1–156.e7. [Google Scholar] [CrossRef]
  88. Li, J.; Chen, C.; Liao, J.; Liu, L.; Ye, X.; Lin, S.; Ye, J. Bond strengths of porcelain to cobalt-chromium alloys made by casting, milling, and selective laser melting. J. Prosthet. Dent. 2017, 118, 69–75. [Google Scholar] [CrossRef]
  89. Svanborg, P.; Hjalmarsson, L. A systematic review on the accuracy of manufacturing techniques for cobalt chromium fixed dental prostheses. Biomater. Investig. Dent. 2020, 7, 31–40. [Google Scholar] [CrossRef]
  90. Hong, J.K.; Kim, S.K.; Heo, S.J.; Koak, J.Y. Mechanical Properties and Metal-Ceramic Bond Strength of Co-Cr Alloy Manufactured by Selective Laser Melting. Materials 2020, 13, 5745. [Google Scholar] [CrossRef]
  91. Bilgin, M.S.; Erdem, A.; Dilber, E.; Ersoy, I. Comparison of fracture resistance between cast, CAD/CAM milling, and direct metal laser sintering metal post systems. J. Prosthodont. Res. 2016, 60, 23–28. [Google Scholar] [CrossRef]
  92. Tamac, E.; Toksavul, S.; Toman, M. Clinical marginal and internal adaptation of CAD/CAM milling, laser sintering, and cast metal ceramic crowns. J. Prosthet. Dent. 2014, 112, 909–913. [Google Scholar] [CrossRef] [PubMed]
  93. Bibb, R.; Eggbeer, D.; Williams, R. Rapid manufacture of removable partial denture frameworks. Rapid Prototyp. J. 2006, 12, 95–99. [Google Scholar] [CrossRef]
  94. Zhou, Y.; Li, N.; Yan, J.; Zeng, Q. Comparative analysis of the microstructures and mechanical properties of Co-Cr dental alloys fabricated by different methods. J. Prosthet. Dent. 2018, 120, 617–623. [Google Scholar] [CrossRef] [PubMed]
  95. Presotto, A.G.C.; Cordeiro, J.M.; Presotto, J.G.C.; Rangel, E.C.; da Cruz, N.C.; Landers, R.; Barão, V.A.R.; Mesquita, M.F. Feasibility of 3D printed Co–Cr alloy for dental prostheses applications. J. Alloys Compd. 2021, 862, 158171. [Google Scholar] [CrossRef]
  96. Huang, Z.; Zhang, L.; Zhu, J.; Zhang, X. Clinical marginal and internal fit of metal ceramic crowns fabricated with a selective laser melting technology. J. Prosthet. Dent. 2015, 113, 623–627. [Google Scholar] [CrossRef]
  97. Alageel, O.; Abdallah, M.N.; Alsheghri, A.; Song, J.; Caron, E.; Tamimi, F. Removable partial denture alloys processed by laser-sintering technique. J. Biomed. Mater. Res. B Appl. Biomater. 2018, 106, 1174–1185. [Google Scholar] [CrossRef]
  98. Schweiger, J.; Guth, J.F.; Erdelt, K.J.; Edelhoff, D.; Schubert, O. Internal porosities, retentive force, and survival of cobalt-chromium alloy clasps fabricated by selective laser-sintering. J. Prosthodont. Res. 2020, 64, 210–216. [Google Scholar] [CrossRef]
  99. Sing, S.L.; Huang, S.; Yeong, W.Y. Effect of solution heat treatment on microstructure and mechanical properties of laser powder bed fusion produced cobalt-28chromium-6molybdenum. Mater. Sci. Eng. A 2020, 769, 138511. [Google Scholar] [CrossRef]
  100. Wang, H.; Lim, J.Y. Metal-ceramic bond strength of a cobalt chromium alloy for dental prosthetic restorations with a porous structure using metal 3D printing. Comput. Biol. Med. 2019, 112, 103364. [Google Scholar] [CrossRef]
  101. Zeng, L.; Zhang, Y.; Liu, Z.; Wei, B. Effects of repeated firing on the marginal accuracy of Co-Cr copings fabricated by selective laser melting. J. Prosthet. Dent. 2015, 113, 135–139. [Google Scholar] [CrossRef]
  102. Padrós, R.; Punset, M.; Molmeneu, M.; Velasco, A.B.; Herrero-Climent, M.; Rupérez, E.; Gil, F.J. Mechanical Properties of CoCr Dental-Prosthesis Restorations Made by Three Manufacturing Processes. Influence of the Microstructure and Topography. Metals 2020, 10, 788. [Google Scholar] [CrossRef]
  103. Kim, D.Y.; Kim, C.M.; Kim, J.H.; Kim, H.Y.; Kim, W.C. Evaluation of marginal and internal gaps of Ni-Cr and Co-Cr alloy copings manufactured by microstereolithography. J. Adv. Prosthodont. 2017, 9, 176–181. [Google Scholar] [CrossRef] [PubMed]
  104. Kim, D.Y.; Kim, J.H.; Kim, H.Y.; Kim, W.C. Comparison and evaluation of marginal and internal gaps in cobalt-chromium alloy copings fabricated using subtractive and additive manufacturing. J. Prosthodont. Res. 2018, 62, 56–64. [Google Scholar] [CrossRef]
  105. Takaichi, A.; Kajima, Y.; Kittikundecha, N.; Htat, H.L.; Wai Cho, H.H.; Hanawa, T.; Yoneyama, T.; Wakabayashi, N. Effect of heat treatment on the anisotropic microstructural and mechanical properties of Co-Cr-Mo alloys produced by selective laser melting. J. Mech. Behav. Biomed. Mater. 2020, 102, 103496. [Google Scholar] [CrossRef]
  106. Alharbi, N.; Wismeijer, D.; Osman, R.B. Additive Manufacturing Techniques in Prosthodontics: Where Do We Currently Stand? A Critical Review. Int. J. Prosthodont. 2017, 30, 474–484. [Google Scholar] [CrossRef]
  107. Kalman, L. 3D printing of a novel dental implant abutment. J. Dent. Res. Dent. Clin. Dent. Prospect. 2018, 12, 299–303. [Google Scholar] [CrossRef] [PubMed]
  108. Fotovat, F.; Shishehian, A.; Alijani, S.; Alafchi, B.; Parchami, P. Comparison of shear bond strength of orthodontic stainless-steel brackets on temporary crowns fabricated by three different methods: An in vitro study. Int. Orthod. 2022, 20, 100641. [Google Scholar] [CrossRef]
  109. Yazdimamaghani, M.; Razavi, M.; Vashaee, D.; Moharamzadeh, K.; Boccaccini, A.R.; Tayebi, L. Porous magnesium-based scaffolds for tissue engineering. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 71, 1253–1266. [Google Scholar] [CrossRef]
  110. Shen, B.; Lu, J.; Wang, Y.; Chen, D.; Han, J.; Zhang, Y.; Zhao, Z. Multimodal-based weld reinforcement monitoring system for wire arc additive manufacturing. J. Mater. Res. Technol. 2022, 20, 561–571. [Google Scholar] [CrossRef]
  111. Zhang, X.; Mao, J.; Zhou, Y.; Ji, F.; Chen, X. Mechanical properties and osteoblast proliferation of complex porous dental implants filled with magnesium alloy based on 3D printing. J. Biomater. Appl. 2021, 35, 1275–1283. [Google Scholar] [CrossRef]
  112. Agnihotri, R.; Gaur, S.; Albin, S. Nanometals in Dentistry: Applications and Toxicological Implications-a Systematic Review. Biol. Trace Elem. Res. 2020, 197, 70–88. [Google Scholar] [CrossRef]
  113. Foong, L.K.; Foroughi, M.M.; Mirhosseini, A.F.; Safaei, M.; Jahani, S.; Mostafavi, M.; Ebrahimpoor, N.; Sharifi, M.; Varma, R.S.; Khatami, M. Applications of nano-materials in diverse dentistry regimes. RSC Adv. 2020, 10, 15430–15460. [Google Scholar] [CrossRef] [PubMed]
  114. Felemban, N.H.; Ebrahim, M.I. The influence of adding modified zirconium oxide-titanium dioxide nano-particles on mechanical properties of orthodontic adhesive: An in vitro study. BMC Oral Health 2017, 17, 43. [Google Scholar] [CrossRef] [PubMed]
  115. Zhang, Y.; Kelly, J.R. Dental Ceramics for Restoration and Metal Veneering. Dent. Clin. North. Am. 2017, 61, 797–819. [Google Scholar] [CrossRef] [PubMed]
  116. Kuriyama, S.; Terui, Y.; Higuchi, D.; Goto, D.; Hotta, Y.; Manabe, A.; Miyazaki, T. Novel fabrication method for zirconia restorations: Bonding strength of machinable ceramic to zirconia with resin cements. Dent. Mater. J. 2011, 30, 419–424. [Google Scholar] [CrossRef]
  117. Oilo, M.; Hardang, A.D.; Ulsund, A.H.; Gjerdet, N.R. Fractographic features of glass-ceramic and zirconia-based dental restorations fractured during clinical function. Eur. J. Oral Sci. 2014, 122, 238–244. [Google Scholar] [CrossRef] [PubMed]
  118. Silva, L.H.D.; Lima, E.; Miranda, R.B.P.; Favero, S.S.; Lohbauer, U.; Cesar, P.F. Dental ceramics: A review of new materials and processing methods. Braz. Oral. Res. 2017, 31, e58. [Google Scholar] [CrossRef] [PubMed]
  119. Raghavendra, S.S.; Jadhav, G.R.; Gathani, K.M.; Kotadia, P. Bioceramics in endodontics—A review. J. Istanb. Univ. Fac. Dent. 2017, 51, S128–S137. [Google Scholar] [CrossRef]
  120. Malament, K.A.; Grossman, D.G. The cast glass-ceramic restoration. J. Prosthet. Dent. 1987, 57, 674–683. [Google Scholar] [CrossRef]
  121. Lang, S.A.; Starr, C.B. Castable glass ceramics for veneer restorations. J. Prosthet. Dent. 1992, 67, 590–594. [Google Scholar] [CrossRef]
  122. Wildgoose, D.G.; Johnson, A.; Winstanley, R.B. Glass/ceramic/refractory techniques, their development and introduction into dentistry: A historical literature review. J. Prosthet. Dent. 2004, 91, 136–143. [Google Scholar] [CrossRef] [PubMed]
  123. Fatahi Bafghi, M.; Mehrabi, H.A.; Eftekhari Yekta, B. Crystallization Behavior, Mechanical Properties, and Chemical Resistance of Leucite-Fluoroapatite Glass-Ceramic Glazes. Int. J. Appl. Glass Sci. 2013, 4, 266–273. [Google Scholar] [CrossRef]
  124. Montazerian, M.; Zanotto, E.D. Bioactive and inert dental glass-ceramics. J. Biomed. Mater. Res. A 2017, 105, 619–639. [Google Scholar] [CrossRef] [PubMed]
  125. Lakhdar, Y.; Tuck, C.; Binner, J.; Terry, A.; Goodridge, R. Additive manufacturing of advanced ceramic materials. Progress. Mater. Sci. 2021, 116, 100736. [Google Scholar] [CrossRef]
  126. Baumgartner, S.; Gmeiner, R.; Schonherr, J.A.; Stampfl, J. Stereolithography-based additive manufacturing of lithium disilicate glass ceramic for dental applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 116, 111180. [Google Scholar] [CrossRef]
  127. Yang, B.; Wang, S.; Wang, G.; Yang, X. Mechanical properties and wear behaviors analysis of fluorapatite glass-ceramics based on stereolithography 3D printing. J. Mech. Behav. Biomed. Mater. 2021, 124, 104859. [Google Scholar] [CrossRef]
  128. Schonherr, J.A.; Baumgartner, S.; Hartmann, M.; Stampfl, J. Stereolithographic Additive Manufacturing of High Precision Glass Ceramic Parts. Materials 2020, 13, 1492. [Google Scholar] [CrossRef]
  129. Wang, W.; Yu, H.; Liu, Y.; Jiang, X.; Gao, B. Trueness analysis of zirconia crowns fabricated with 3-dimensional printing. J. Prosthet. Dent. 2019, 121, 285–291. [Google Scholar] [CrossRef]
  130. Dehurtevent, M.; Robberecht, L.; Hornez, J.-C.; Thuault, A.; Deveaux, E.; Béhin, P. Stereolithography: A new method for processing dental ceramics by additive computer-aided manufacturing. Dent. Mater. 2017, 33, 477–485. [Google Scholar] [CrossRef]
  131. Quigley, N.P.; Loo, D.S.S.; Choy, C.; Ha, W.N. Clinical efficacy of methods for bonding to zirconia: A systematic review. J. Prosthet. Dent. 2021, 125, 231–240. [Google Scholar] [CrossRef]
  132. Turp, V.; Tuncelli, B.; Sen, D.; Goller, G. Evaluation of hardness and fracture toughness, coupled with microstructural analysis, of zirconia ceramics stored in environments with different pH values. Dent. Mater. J. 2012, 31, 891–902. [Google Scholar] [CrossRef]
  133. Denry, I.; Kelly, J.R. Emerging ceramic-based materials for dentistry. J. Dent. Res. 2014, 93, 1235–1242. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, Y.; Lawn, B.R. Evaluating dental zirconia. Dent. Mater. 2019, 35, 15–23. [Google Scholar] [CrossRef]
  135. Sanon, C.; Chevalier, J.; Douillard, T.; Cattani-Lorente, M.; Scherrer, S.S.; Gremillard, L. A new testing protocol for zirconia dental implants. Dent. Mater. 2015, 31, 15–25. [Google Scholar] [CrossRef] [PubMed]
  136. Rodriguez, A.E.; Monzavi, M.; Yokoyama, C.L.; Nowzari, H. Zirconia dental implants: A clinical and radiographic evaluation. J. Esthet. Restor. Dent. 2018, 30, 538–544. [Google Scholar] [CrossRef] [PubMed]
  137. Ghodsi, S.; Jafarian, Z. A Review on Translucent Zirconia. Eur. J. Prosthodont. Restor. Dent. 2018, 26, 62–74. [Google Scholar] [CrossRef]
  138. Sadowsky, S.J. Has zirconia made a material difference in implant prosthodontics? A review. Dent. Mater. 2020, 36, 1–8. [Google Scholar] [CrossRef] [PubMed]
  139. Sivaraman, K.; Chopra, A.; Narayan, A.I.; Balakrishnan, D. Is zirconia a viable alternative to titanium for oral implant? A critical review. J. Prosthodont. Res. 2018, 62, 121–133. [Google Scholar] [CrossRef]
  140. Hanawa, T. Zirconia versus titanium in dentistry: A review. Dent. Mater. J. 2020, 39, 24–36. [Google Scholar] [CrossRef]
  141. Zhang, Y.; Lawn, B.R. Novel Zirconia Materials in Dentistry. J. Dent. Res. 2017, 97, 140–147. [Google Scholar] [CrossRef]
  142. Carrillo, M.H.; Lee, G.; Maniere, C.; Olevsky, E.A. Additive manufacturing of powder components based on subtractive sintering approach. Rapid Prototyp. J. 2021, 27, 1731–1736. [Google Scholar] [CrossRef]
  143. Galante, R.; Figueiredo-Pina, C.G.; Serro, A.P. Additive manufacturing of ceramics for dental applications: A review. Dent. Mater. 2019, 35, 825–846. [Google Scholar] [CrossRef] [PubMed]
  144. Jayaswal, G.P.; Dange, S.P.; Khalikar, A.N. Bioceramic in dental implants: A review. J. Indian. Prosthodont. Soc. 2010, 10, 8–12. [Google Scholar] [CrossRef] [PubMed]
  145. Conrad, H.J.; Seong, W.-J.; Pesun, I.J. Current ceramic materials and systems with clinical recommendations: A systematic review. J. Prosthet. Dent. 2007, 98, 389–404. [Google Scholar] [CrossRef]
  146. Sarker, S.; Mumu, H.T.; Al-Amin, M.; Zahangir Alam, M.; Gafur, M.A. Impacts of inclusion of additives on physical, microstructural, and mechanical properties of Alumina and Zirconia toughened alumina (ZTA) ceramic composite: A review. Mater. Today 2022, 62, 2892–2918. [Google Scholar] [CrossRef]
  147. Eickenscheidt, M.; Langenmair, M.; Dbouk, A.; Notzel, D.; Hanemann, T.; Stieglitz, T. 3D-Printed Hermetic Alumina Housings. Materials 2021, 14, 200. [Google Scholar] [CrossRef]
  148. Raheem, A.A.; Hameed, P.; Whenish, R.; Elsen, R.S.; G, A.; Jaiswal, A.K.; Prashanth, K.G.; Manivasagam, G. A Review on Development of Bio-Inspired Implants Using 3D Printing. Biomimetics 2021, 6, 65. [Google Scholar] [CrossRef]
  149. Notzel, D.; Eickhoff, R.; Hanemann, T. Fused Filament Fabrication of Small Ceramic Components. Materials 2018, 11, 1483. [Google Scholar] [CrossRef]
  150. Chugunov, S.; Adams, N.A.; Akhatov, I. Evolution of SLA-Based Al2O3 Microstructure During Additive Manufacturing Process. Materials 2020, 13, 3928. [Google Scholar] [CrossRef]
  151. Mohd, N.; Razali, M.; Ghazali, M.J.; Abu Kasim, N.H. Current Advances of Three-Dimensional Bioprinting Application in Dentistry: A Scoping Review. Materials 2022, 15, 6398. [Google Scholar] [CrossRef]
  152. Gopinathan, J.; Noh, I. Recent trends in bioinks for 3D printing. Biomater. Res. 2018, 22, 11. [Google Scholar] [CrossRef]
  153. Kaushik, S.N.; Kim, B.; Walma, A.M.; Choi, S.C.; Wu, H.; Mao, J.J.; Jun, H.W.; Cheon, K. Biomimetic microenvironments for regenerative endodontics. Biomater. Res. 2016, 20, 14. [Google Scholar] [CrossRef] [PubMed]
  154. Hospodiuk, M.; Dey, M.; Sosnoski, D.; Ozbolat, I.T. The bioink: A comprehensive review on bioprintable materials. Biotechnol. Adv. 2017, 35, 217–239. [Google Scholar] [CrossRef]
  155. Park, J.Y.; Choi, Y.J.; Shim, J.H.; Park, J.H.; Cho, D.W. Development of a 3D cell printed structure as an alternative to autologs cartilage for auricular reconstruction. J. Biomed. Mater. Res. B Appl. Biomater. 2017, 105, 1016–1028. [Google Scholar] [CrossRef] [PubMed]
  156. Sharma, S.; Srivastava, D.; Grover, S.; Sharma, V. Biomaterials in tooth tissue engineering: A review. J. Clin. Diagn. Res. 2014, 8, 309–315. [Google Scholar] [CrossRef]
  157. 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]
  158. Kang, H.W.; Lee, S.J.; Ko, I.K.; Kengla, C.; Yoo, J.J.; Atala, A. A 3D bioprinting system to produce human-scale tissue constructs with structural integrity. Nat. Biotechnol. 2016, 34, 312–319. [Google Scholar] [CrossRef] [PubMed]
  159. Hockaday, L.A.; Kang, K.H.; Colangelo, N.W.; Cheung, P.Y.; Duan, B.; Malone, E.; Wu, J.; Girardi, L.N.; Bonassar, L.J.; Lipson, H.; et al. Rapid 3D printing of anatomically accurate and mechanically heterogeneous aortic valve hydrogel scaffolds. Biofabrication 2012, 4, 035005. [Google Scholar] [CrossRef]
  160. Rodriguez-Salvador, M.; Ruiz-Cantu, L. Revealing emerging science and technology research for dentistry applications of 3D bioprinting. Int. J. Bioprint 2019, 5, 170. [Google Scholar] [CrossRef]
  161. Asa’ad, F.; Pagni, G.; Pilipchuk, S.P.; Gianni, A.B.; Giannobile, W.V.; Rasperini, G. 3D-Printed Scaffolds and Biomaterials: Review of Alveolar Bone Augmentation and Periodontal Regeneration Applications. Int. J. Dent. 2016, 2016, 1239842. [Google Scholar] [CrossRef]
  162. Stepanovska, J.; Supova, M.; Hanzalek, K.; Broz, A.; Matejka, R. Collagen Bioinks for Bioprinting: A Systematic Review of Hydrogel Properties, Bioprinting Parameters, Protocols, and Bioprinted Structure Characteristics. Biomedicines 2021, 9, 1137. [Google Scholar] [CrossRef] [PubMed]
  163. Guvendiren, M.; Burdick, J.A. Engineering synthetic hydrogel microenvironments to instruct stem cells. Curr. Opin. Biotechnol. 2013, 24, 841–846. [Google Scholar] [CrossRef] [PubMed]
  164. Muthukrishnan, L. Imminent antimicrobial bioink deploying cellulose, alginate, EPS and synthetic polymers for 3D bioprinting of tissue constructs. Carbohydr. Polym. 2021, 260, 117774. [Google Scholar] [CrossRef] [PubMed]
  165. Park, S.; Oh, K.K.; Lee, S.H. Biopolymer-Based Composite Materials Prepared Using Ionic Liquids. Adv. Biochem. Eng. Biotechnol. 2019, 168, 133–176. [Google Scholar] [CrossRef] [PubMed]
  166. Sudha, P.N.; Rose, M.H. Beneficial effects of hyaluronic acid. Adv. Food Nutr. Res. 2014, 72, 137–176. [Google Scholar] [CrossRef] [PubMed]
  167. Yoo, H.S.; Lee, E.A.; Yoon, J.J.; Park, T.G. Hyaluronic acid modified biodegradable scaffolds for cartilage tissue engineering. Biomaterials 2005, 26, 1925–1933. [Google Scholar] [CrossRef] [PubMed]
  168. Poldervaart, M.T.; Goversen, B.; de Ruijter, M.; Abbadessa, A.; Melchels, F.P.W.; Oner, F.C.; Dhert, W.J.A.; Vermonden, T.; Alblas, J. 3D bioprinting of methacrylated hyaluronic acid (MeHA) hydrogel with intrinsic osteogenicity. PLoS ONE 2017, 12, e0177628. [Google Scholar] [CrossRef]
  169. Annabi, N.; Tamayol, A.; Uquillas, J.A.; Akbari, M.; Bertassoni, L.E.; Cha, C.; Camci-Unal, G.; Dokmeci, M.R.; Peppas, N.A.; Khademhosseini, A. 25th anniversary article: Rational design and applications of hydrogels in regenerative medicine. Adv. Mater. 2014, 26, 85–123. [Google Scholar] [CrossRef]
  170. Drury, J.L.; Mooney, D.J. Hydrogels for tissue engineering: Scaffold design variables and applications. Biomaterials 2003, 24, 4337–4351. [Google Scholar] [CrossRef]
  171. Highley, C.B.; Prestwich, G.D.; Burdick, J.A. Recent advances in hyaluronic acid hydrogels for biomedical applications. Curr. Opin. Biotechnol. 2016, 40, 35–40. [Google Scholar] [CrossRef]
  172. Ouyang, L.; Highley, C.B.; Rodell, C.B.; Sun, W.; Burdick, J.A. 3D Printing of Shear-Thinning Hyaluronic Acid Hydrogels with Secondary Cross-Linking. ACS Biomater. Sci. Eng. 2016, 2, 1743–1751. [Google Scholar] [CrossRef] [PubMed]
  173. Law, N.; Doney, B.; Glover, H.; Qin, Y.; Aman, Z.M.; Sercombe, T.B.; Liew, L.J.; Dilley, R.J.; Doyle, B.J. Characterisation of hyaluronic acid methylcellulose hydrogels for 3D bioprinting. J. Mech. Behav. Biomed. Mater. 2018, 77, 389–399. [Google Scholar] [CrossRef] [PubMed]
  174. Stichler, S.; Bock, T.; Paxton, N.; Bertlein, S.; Levato, R.; Schill, V.; Smolan, W.; Malda, J.; Tessmar, J.; Blunk, T.; et al. Double printing of hyaluronic acid/poly(glycidol) hybrid hydrogels with poly(epsilon-caprolactone) for MSC chondrogenesis. Biofabrication 2017, 9, 044108. [Google Scholar] [CrossRef] [PubMed]
  175. Casale, M.; Moffa, A.; Vella, P.; Sabatino, L.; Capuano, F.; Salvinelli, B.; Lopez, M.A.; Carinci, F.; Salvinelli, F. Hyaluronic acid: Perspectives in dentistry. A systematic review. Int. J. Immunopathol. Pharm. 2016, 29, 572–582. [Google Scholar] [CrossRef]
  176. van Uden, S.; Silva-Correia, J.; Oliveira, J.M.; Reis, R.L. Current strategies for treatment of intervertebral disc degeneration: Substitution and regeneration possibilities. Biomater. Res. 2017, 21, 22. [Google Scholar] [CrossRef]
  177. Smith, C.M.; Stone, A.; Parkhill, R.L.; Stewart, R.L.; Simpkins, M.W.; Kachurin, A.M.; Warren, W.L.; Williams, S.K. Three-dimensional bioassembly tool for generating viable tissue-engineered constructs. Tissue Eng. 2004, 10, 1566–1576. [Google Scholar] [CrossRef]
  178. Yang, X.; Lu, Z.; Wu, H.; Li, W.; Zheng, L.; Zhao, J. Collagen-alginate as bioink for three-dimensional (3D) cell printing based cartilage tissue engineering. Mater. Sci. Eng. C Mater. Biol. Appl. 2018, 83, 195–201. [Google Scholar] [CrossRef]
  179. Stratesteffen, H.; Kopf, M.; Kreimendahl, F.; Blaeser, A.; Jockenhoevel, S.; Fischer, H. GelMA-collagen blends enable drop-on-demand 3D printablility and promote angiogenesis. Biofabrication 2017, 9, 045002. [Google Scholar] [CrossRef]
  180. Arif, Z.U.; Khalid, M.Y.; Sheikh, M.F.; Zolfagharian, A.; Bodaghi, M. Biopolymeric sustainable materials and their emerging applications. J. Environ. Chem. Eng. 2022, 10, 108159. [Google Scholar] [CrossRef]
  181. Kreimendahl, F.; Kopf, M.; Thiebes, A.L.; Duarte Campos, D.F.; Blaeser, A.; Schmitz-Rode, T.; Apel, C.; Jockenhoevel, S.; Fischer, H. Three-Dimensional Printing and Angiogenesis: Tailored Agarose-Type I Collagen Blends Comprise Three-Dimensional Printability and Angiogenesis Potential for Tissue-Engineered Substitutes. Tissue Eng. Part. C Methods 2017, 23, 604–615. [Google Scholar] [CrossRef]
  182. Forget, A.; Blaeser, A.; Miessmer, F.; Kopf, M.; Campos, D.F.D.; Voelcker, N.H.; Blencowe, A.; Fischer, H.; Shastri, V.P. Mechanically Tunable Bioink for 3D Bioprinting of Human Cells. Adv. Healthc. Mater. 2017, 6, 1700255. [Google Scholar] [CrossRef] [PubMed]
  183. Xiong, J.Y.; Narayanan, J.; Liu, X.Y.; Chong, T.K.; Chen, S.B.; Chung, T.S. Topology evolution and gelation mechanism of agarose gel. J. Phys. Chem. B 2005, 109, 5638–5643. [Google Scholar] [CrossRef] [PubMed]
  184. Fakhri, E.; Eslami, H.; Maroufi, P.; Pakdel, F.; Taghizadeh, S.; Ganbarov, K.; Yousefi, M.; Tanomand, A.; Yousefi, B.; Mahmoudi, S.; et al. Chitosan biomaterials application in dentistry. Int. J. Biol. Macromol. 2020, 162, 956–974. [Google Scholar] [CrossRef] [PubMed]
  185. Benwood, C.; Chrenek, J.; Kirsch, R.L.; Masri, N.Z.; Richards, H.; Teetzen, K.; Willerth, S.M. Natural Biomaterials and Their Use as Bioinks for Printing Tissues. Bioengineering 2021, 8, 27. [Google Scholar] [CrossRef] [PubMed]
  186. Axpe, E.; Oyen, M.L. Applications of Alginate-Based Bioinks in 3D Bioprinting. Int. J. Mol. Sci. 2016, 17, 1976. [Google Scholar] [CrossRef] [PubMed]
  187. Zhang, Y.; Yu, Y.; Chen, H.; Ozbolat, I.T. Characterization of printable cellular micro-fluidic channels for tissue engineering. Biofabrication 2013, 5, 025004. [Google Scholar] [CrossRef]
  188. Gao, Q.; He, Y.; Fu, J.Z.; Liu, A.; Ma, L. Coaxial nozzle-assisted 3D bioprinting with built-in microchannels for nutrients delivery. Biomaterials 2015, 61, 203–215. [Google Scholar] [CrossRef]
  189. Armstrong, J.P.; Burke, M.; Carter, B.M.; Davis, S.A.; Perriman, A.W. 3D Bioprinting Using a Templated Porous Bioink. Adv. Heal. Mater. 2016, 5, 1724–1730. [Google Scholar] [CrossRef]
  190. Wang, X.; Tolba, E.; Schroder, H.C.; Neufurth, M.; Feng, Q.; Diehl-Seifert, B.; Muller, W.E. Effect of bioglass on growth and biomineralization of SaOS-2 cells in hydrogel after 3D cell bioprinting. PLoS ONE 2014, 9, e112497. [Google Scholar] [CrossRef]
  191. Cai, Y.; Chang, S.Y.; Gan, S.W.; Ma, S.; Lu, W.F.; Yen, C.-C. Nanocomposite bioinks for 3D bioprinting. Acta Biomater. 2022, 151, 45–69. [Google Scholar] [CrossRef]
  192. Chmolowska, D.; Hamda, N.; Laskowski, R. Cellulose decomposed faster in fallow soil than in meadow soil due to a shorter lag time. J. Soils Sediments 2016, 17, 299–305. [Google Scholar] [CrossRef]
  193. Liu, M.; Ma, C.; Zhou, D.; Chen, S.; Zou, L.; Wang, H.; Wu, J. Hydrophobic, breathable cellulose nonwoven fabrics for disposable hygiene applications. Carbohydr. Polym. 2022, 288, 119367. [Google Scholar] [CrossRef] [PubMed]
  194. Schwab, A.; Levato, R.; D’Este, M.; Piluso, S.; Eglin, D.; Malda, J. Printability and Shape Fidelity of Bioinks in 3D Bioprinting. Chem. Rev. 2020, 120, 11028–11055. [Google Scholar] [CrossRef] [PubMed]
  195. Shavandi, A.; Hosseini, S.; Okoro, O.V.; Nie, L.; Eghbali Babadi, F.; Melchels, F. 3D Bioprinting of Lignocellulosic Biomaterials. Adv. Heal. Mater. 2020, 9, e2001472. [Google Scholar] [CrossRef]
  196. Bayer, I.S. Hyaluronic Acid and Controlled Release: A Review. Molecules 2020, 25, 2649. [Google Scholar] [CrossRef]
  197. Bloebaum, R.D.; Wilson, A.S.; Martin, W.N. A Review of the Collagen Orientation in the Articular Cartilage. Cartilage 2021, 13, 367s–374s. [Google Scholar] [CrossRef]
  198. Razavi, M.; Qiao, Y.; Thakor, A.S. Three-dimensional cryogels for biomedical applications. J. Biomed. Mater. Res. A 2019, 107, 2736–2755. [Google Scholar] [CrossRef]
  199. Emami, Z.; Ehsani, M.; Zandi, M.; Foudazi, R. Controlling alginate oxidation conditions for making alginate-gelatin hydrogels. Carbohydr. Polym. 2018, 198, 509–517. [Google Scholar] [CrossRef]
  200. Bodin, A.; Gustafsson, L.; Gatenholm, P. Surface-engineered bacterial cellulose as template for crystallization of calcium phosphate. J. Biomater. Sci. Polym. Ed. 2006, 17, 435–447. [Google Scholar] [CrossRef]
  201. Collart-Dutilleul, P.-Y.; Panayotov, I.; Secret, E.; Cunin, F.; Gergely, C.; Cuisinier, F.; Martin, M. Initial stem cell adhesion on porous silicon surface: Molecular architecture of actin cytoskeleton and filopodial growth. Nanoscale Res. Lett. 2014, 9, 564. [Google Scholar] [CrossRef]
  202. Nuti, N.; Corallo, C.; Chan, B.M.F.; Ferrari, M.; Gerami-Naini, B. Multipotent Differentiation of Human Dental Pulp Stem Cells: A Literature Review. Stem Cell. Rev. Rep. 2016, 12, 511–523. [Google Scholar] [CrossRef]
  203. Nada, O.A.; El Backly, R.M. Stem Cells From the Apical Papilla (SCAP) as a Tool for Endogenous Tissue Regeneration. Front. Bioeng. Biotechnol. 2018, 6, 103. [Google Scholar] [CrossRef] [PubMed]
  204. Huang, G.T.J.; Garcia-Godoy, F. 13—Stem cells and dental tissue reconstruction. In Material-Tissue Interfacial Phenomena; Spencer, P., Misra, A., Eds.; Woodhead Publishing: Sawston, UK, 2017; pp. 325–353. [Google Scholar] [CrossRef]
  205. Simon, S.; Cooper, P.; Isaac, J.; Berdal, A. 18—Tissue engineering and endodontics. In Preprosthetic and Maxillofacial Surgery; Ferri, J., Hunziker, E.B., Eds.; Woodhead Publishing: Sawston, UK, 2011; pp. 336–362. [Google Scholar] [CrossRef]
  206. Kang, J.; Fan, W.; Deng, Q.; He, H.; Huang, F. Stem Cells from the Apical Papilla: A Promising Source for Stem Cell-Based Therapy. BioMed Res. Int. 2019, 2019, 6104738. [Google Scholar] [CrossRef] [PubMed]
  207. Kérourédan, O.; Bourget, J.-M.; Rémy, M.; Crauste-Manciet, S.; Kalisky, J.; Catros, S.; Thébaud, N.B.; Devillard, R. Micropatterning of endothelial cells to create a capillary-like network with defined architecture by laser-assisted bioprinting. J. Mater. Sci. 2019, 30, 28. [Google Scholar] [CrossRef] [PubMed]
  208. Gao, H.; Li, B.; Zhao, L.; Jin, Y. Influence of nanotopography on periodontal ligament stem cell functions and cell sheet based periodontal regeneration. Int. J. Nanomed. 2015, 10, 4009–4027. [Google Scholar] [CrossRef]
  209. Salar Amoli, M.; EzEldeen, M.; Jacobs, R.; Bloemen, V. Materials for Dentoalveolar Bioprinting: Current State of the Art. Biomedicines 2022, 10, 71. [Google Scholar] [CrossRef]
  210. Mastrangelo, F.; Nargi, E.; Carone, L.; Dolci, M.; Caciagli, F.; Ciccarelli, R.; Lutiis, M.A.D.; Karapanou, V.; Shaik, B.Y.; Conti, P.; et al. Tridimensional Response of human Dental Follicular Stem Cells onto a Synthetic Hydroxyapatite Scaffold. J. Health Sci. 2008, 54, 154–161. [Google Scholar] [CrossRef]
  211. Zhou, T.; Pan, J.; Wu, P.; Huang, R.; Du, W.; Zhou, Y.; Wan, M.; Fan, Y.; Xu, X.; Zhou, X.; et al. Dental Follicle Cells: Roles in Development and Beyond. Stem Cells Int. 2019, 2019, 9159605. [Google Scholar] [CrossRef]
  212. Dave, J.R.; Chandekar, S.S.; Behera, S.; Desai, K.U.; Salve, P.M.; Sapkal, N.B.; Mhaske, S.T.; Dewle, A.M.; Pokare, P.S.; Page, M.; et al. Human gingival mesenchymal stem cells retain their growth and immunomodulatory characteristics independent of donor age. Sci. Adv. 2022, 8, eabm6504. [Google Scholar] [CrossRef]
  213. Zhang, Q.; Nguyen, P.D.; Shi, S.; Burrell, J.C.; Cullen, D.K.; Le, A.D. 3D bio-printed scaffold-free nerve constructs with human gingiva-derived mesenchymal stem cells promote rat facial nerve regeneration. Sci. Rep. 2018, 8, 6634. [Google Scholar] [CrossRef]
  214. Pizzicannella, J.; Diomede, F.; Gugliandolo, A.; Chiricosta, L.; Bramanti, P.; Merciaro, I.; Orsini, T.; Mazzon, E.; Trubiani, O. 3D Printing PLA/Gingival Stem Cells/ EVs Upregulate miR-2861 and -210 during Osteoangiogenesis Commitment. Int. J. Mol. Sci. 2019, 20, 3256. [Google Scholar] [CrossRef] [PubMed]
  215. Collado-González, M.; García-Bernal, D.; Oñate-Sánchez, R.E.; Ortolani-Seltenerich, P.S.; Álvarez-Muro, T.; Lozano, A.; Forner, L.; Llena, C.; Moraleda, J.M.; Rodríguez-Lozano, F.J. Cytotoxicity and bioactivity of various pulpotomy materials on stem cells from human exfoliated primary teeth. Int. Endod. J. 2017, 50, e19–e30. [Google Scholar] [CrossRef]
  216. Gonzalez-Fernandez, T.; Rathan, S.; Hobbs, C.; Pitacco, P.; Freeman, F.E.; Cunniffe, G.M.; Dunne, N.J.; McCarthy, H.O.; Nicolosi, V.; O’Brien, F.J.; et al. Pore-forming bioinks to enable spatio-temporally defined gene delivery in bioprinted tissues. J. Control. Release 2019, 301, 13–27. [Google Scholar] [CrossRef] [PubMed]
  217. Kuo, C.; Tuan, R. Tissue Engineering with Mesenchymal Stem Cells. IEEE Eng. Med. Biol. Mag. 2003, 22, 51–56. [Google Scholar] [CrossRef] [PubMed]
  218. Attalla, R.; Puersten, E.; Jain, N.; Selvaganapathy, P.R. 3D bioprinting of heterogeneous bi- and tri-layered hollow channels within gel scaffolds using scalable multi-axial microfluidic extrusion nozzle. Biofabrication 2019, 11, 015012. [Google Scholar] [CrossRef] [PubMed]
  219. van den Beucken, J.J.J.P.; Walboomers, X.F.; Nillesen, S.T.M.; Vos, M.R.J.; Sommerdijk, N.A.J.M.; van Kuppevelt, T.H.; Nolte, R.J.M.; Jansen, J.A. In Vitro and In Vivo Effects of Deoxyribonucleic Acid–Based Coatings Funtionalized with Vascular Endothelial Growth Factor. Tissue Eng. 2007, 13, 711–720. [Google Scholar] [CrossRef]
  220. Liu, X.; Gaihre, B.; George, M.N.; Miller II, A.L.; Xu, H.; Waletzki, B.E.; Lu, L. 3D bioprinting of oligo(poly[ethylene glycol] fumarate) for bone and nerve tissue engineering. J. Biomed. Mater. Res. Part A 2021, 109, 6–17. [Google Scholar] [CrossRef]
  221. Su, B.; Peng, X.; Jiang, D.; Wu, J.; Qiao, B.; Li, W.; Qi, X. In Vitro and In Vivo Evaluations of Nano-Hydroxyapatite/Polyamide 66/Glass Fibre (n-HA/PA66/GF) as a Novel Bioactive Bone Screw. PLoS ONE 2013, 8, e68342. [Google Scholar] [CrossRef]
  222. Taguchi, T.; Yanagi, Y.; Yoshimaru, K.; Zhang, X.-Y.; Matsuura, T.; Nakayama, K.; Kobayashi, E.; Yamaza, H.; Nonaka, K.; Ohga, S.; et al. Regenerative medicine using stem cells from human exfoliated deciduous teeth (SHED): A promising new treatment in pediatric surgery. Surg. Today 2019, 49, 316–322. [Google Scholar] [CrossRef]
  223. Jin, R.; Song, G.; Chai, J.; Gou, X.; Yuan, G.; Chen, Z. Effects of concentrated growth factor on proliferation, migration, and differentiation of human dental pulp stem cells in vitro. J. Tissue Eng. 2018, 9, 2041731418817505. [Google Scholar] [CrossRef]
  224. Park, J.H.; Gillispie, G.J.; Copus, J.S.; Zhang, W.; Atala, A.; Yoo, J.J.; Yelick, P.C.; Lee, S.J. The effect of BMP-mimetic peptide tethering bioinks on the differentiation of dental pulp stem cells (DPSCs) in 3D bioprinted dental constructs. Biofabrication 2020, 12, 035029. [Google Scholar] [CrossRef] [PubMed]
  225. Bakopoulou, A.; Leyhausen, G.; Volk, J.; Tsiftsoglou, A.; Garefis, P.; Koidis, P.; Geurtsen, W. Comparative analysis of in vitro osteo/odontogenic differentiation potential of human dental pulp stem cells (DPSCs) and stem cells from the apical papilla (SCAP). Arch. Oral Biol. 2011, 56, 709–721. [Google Scholar] [CrossRef] [PubMed]
  226. Thattaruparambil Raveendran, N.; Vaquette, C.; Meinert, C.; Samuel Ipe, D.; Ivanovski, S. Optimization of 3D bioprinting of periodontal ligament cells. Dent. Mater. 2019, 35, 1683–1694. [Google Scholar] [CrossRef] [PubMed]
  227. Rombouts, C.; Giraud, T.; Jeanneau, C.; About, I. Pulp Vascularization during Tooth Development, Regeneration, and Therapy. J. Dent. Res. 2017, 96, 137–144. [Google Scholar] [CrossRef] [PubMed]
  228. Abdal-Wahab, M.; Abdel Ghaffar, K.A.; Ezzatt, O.M.; Hassan, A.A.A.; El Ansary, M.M.S.; Gamal, A.Y. Regenerative potential of cultured gingival fibroblasts in treatment of periodontal intrabony defects (randomized clinical and biochemical trial). J. Periodontal Res. 2020, 55, 441–452. [Google Scholar] [CrossRef] [PubMed]
  229. Guo, B.; Tang, C.; Wang, M.; Zhao, Z.; Shokoohi-Tabrizi, H.A.; Shi, B.; Andrukhov, O.; Rausch-Fan, X. In vitro biocompatibility of biohybrid polymers membrane evaluated in human gingival fibroblasts. J. Biomed. Mater. Res. Part B 2020, 108, 2590–2598. [Google Scholar] [CrossRef]
  230. Nishino, Y.; Yamada, Y.; Ebisawa, K.; Nakamura, S.; Okabe, K.; Umemura, E.; Hara, K.; Ueda, M. Stem cells from human exfoliated deciduous teeth (SHED) enhance wound healing and the possibility of novel cell therapy. Cytotherapy 2011, 13, 598–605. [Google Scholar] [CrossRef]
  231. Miura, M.; Gronthos, S.; Zhao, M.; Lu, B.; Fisher, L.W.; Robey, P.G.; Shi, S. SHED: Stem cells from human exfoliated deciduous teeth. Proc. Natl. Acad. Sci. USA 2003, 100, 5807–5812. [Google Scholar] [CrossRef]
  232. Han, J.; Kim, D.S.; Jang, H.; Kim, H.-R.; Kang, H.-W. Bioprinting of three-dimensional dentin–pulp complex with local differentiation of human dental pulp stem cells. J. Tissue Eng. 2019, 10, 2041731419845849. [Google Scholar] [CrossRef]
  233. 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]
  234. Xu, H.; Liu, J.; Zhang, Z.; Xu, C. Cell sedimentation during 3D bioprinting: A mini review. Bio-Des. Manuf. 2022, 5, 617–626. [Google Scholar] [CrossRef]
  235. Thrivikraman, G.; Athirasala, A.; Twohig, C.; Boda, S.K.; Bertassoni, L.E. Biomaterials for Craniofacial Bone Regeneration. Dent. Clin. North. Am. 2017, 61, 835–856. [Google Scholar] [CrossRef] [PubMed]
  236. Lee, K.; Silva, E.A.; Mooney, D.J. Growth factor delivery-based tissue engineering: General approaches and a review of recent developments. J. R. Soc. Interface 2011, 8, 153–170. [Google Scholar] [CrossRef] [PubMed]
  237. Cheng, M.; Qin, G. Chapter 11—Progenitor Cell Mobilization and Recruitment: SDF-1, CXCR4, α4-integrin, and c-kit. In Progress in Molecular Biology and Translational Science; Tang, Y., Ed.; Academic Press: Cambridge, MA, USA, 2012; Volume 111, pp. 243–264. [Google Scholar]
  238. Lauer, A.; Wolf, P.; Mehler, D.; Gotz, H.; Ruzgar, M.; Baranowski, A.; Henrich, D.; Rommens, P.M.; Ritz, U. Biofabrication of SDF-1 Functionalized 3D-Printed Cell-Free Scaffolds for Bone Tissue Regeneration. Int. J. Mol. Sci. 2020, 21, 2175. [Google Scholar] [CrossRef]
  239. Kim, J.; Choi, H.S.; Kim, Y.M.; Song, S.C. Thermo-Responsive Nanocomposite Bioink with Growth-Factor Holding and its Application to Bone Regeneration. Small 2023, 19, e2203464. [Google Scholar] [CrossRef]
  240. Tan, J.; Zhang, M.; Hai, Z.; Wu, C.; Lin, J.; Kuang, W.; Tang, H.; Huang, Y.; Chen, X.; Liang, G. Sustained Release of Two Bioactive Factors from Supramolecular Hydrogel Promotes Periodontal Bone Regeneration. ACS Nano 2019, 13, 5616–5622. [Google Scholar] [CrossRef]
  241. Konopnicki, S.; Sharaf, B.; Resnick, C.; Patenaude, A.; Pogal-Sussman, T.; Hwang, K.G.; Abukawa, H.; Troulis, M.J. Tissue-engineered bone with 3-dimensionally printed beta-tricalcium phosphate and polycaprolactone scaffolds and early implantation: An in vivo pilot study in a porcine mandible model. J. Oral. Maxillofac. Surg. 2015, 73, 1016.e1–1016.e11. [Google Scholar] [CrossRef]
  242. Freeman, F.E.; Pitacco, P.; van Dommelen, L.H.A.; Nulty, J.; Browe, D.C.; Shin, J.Y.; Alsberg, E.; Kelly, D.J. 3D bioprinting spatiotemporally defined patterns of growth factors to tightly control tissue regeneration. Sci. Adv. 2020, 6, eabb5093. [Google Scholar] [CrossRef]
  243. Li, Z.; Lin, H.; Shi, S.; Su, K.; Zheng, G.; Gao, S.; Zeng, X.; Ning, H.; Yu, M.; Li, X.; et al. Controlled and Sequential Delivery of Stromal Derived Factor-1 alpha (SDF-1alpha) and Magnesium Ions from Bifunctional Hydrogel for Bone Regeneration. Polymers 2022, 14, 2872. [Google Scholar] [CrossRef]
  244. Zhang, Y.; Zhao, Y.; Xie, Z.; Li, M.; Liu, Y.; Tu, X. Activating Wnt/beta-Catenin Signaling in Osteocytes Promotes Osteogenic Differentiation of BMSCs through BMP-7. Int. J. Mol. Sci. 2022, 23, 16045. [Google Scholar] [CrossRef]
  245. Kimura, M.; Miyahara, K.; Yamasaki, M.; Uchida, N. Comparison of vascular endothelial growth factor/vascular endothelial growth factor receptor 2 expression and its relationship to tumor cell proliferation in canine epithelial and mesenchymal tumors. J. Vet. Med. Sci. 2022, 84, 133–141. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Polymer applications in 3D printing dentistry. (A) Bioengineering scaffold manufactured from PCL in a dog’s mandible. Reprinted with permission from Ref. [37]. (B) Printed PMMA for the provisional crown. Reprinted with permission from Ref. [38]. Copyright 2022 Elsevier. (C) PLA 3D printed scaffold. Reprinted with permission from Ref. [39]. (D) PLGA low-temperature solvent-based 3D printed material for biological membrane. Reprinted with permission from Ref. [40]. Copyright 2020 Taylor & Francis.
Figure 1. Polymer applications in 3D printing dentistry. (A) Bioengineering scaffold manufactured from PCL in a dog’s mandible. Reprinted with permission from Ref. [37]. (B) Printed PMMA for the provisional crown. Reprinted with permission from Ref. [38]. Copyright 2022 Elsevier. (C) PLA 3D printed scaffold. Reprinted with permission from Ref. [39]. (D) PLGA low-temperature solvent-based 3D printed material for biological membrane. Reprinted with permission from Ref. [40]. Copyright 2020 Taylor & Francis.
Polymers 15 02405 g001
Figure 2. 3D-printed metals in dentistry. (A) EBM-manufactured Ti mandibular implant. Reprinted with permission from Ref. [84]. Copyright 2018 Springer Nature. (B) Co–Cr prosthesis manufactured by DMLS. Reprinted with permission from Ref. [85].
Figure 2. 3D-printed metals in dentistry. (A) EBM-manufactured Ti mandibular implant. Reprinted with permission from Ref. [84]. Copyright 2018 Springer Nature. (B) Co–Cr prosthesis manufactured by DMLS. Reprinted with permission from Ref. [85].
Polymers 15 02405 g002
Figure 3. Ceramics for prosthetic dentistry. (A) Bioactive glass inlay. Reprinted with permission from Ref. [124]. Copyright 2016 Wiley. (B) Zirconia crown restoration [129]. (C) SLA-printed alumina crown. Reprinted with permission from Ref. [130]. Copyright 2017 Elsevier.
Figure 3. Ceramics for prosthetic dentistry. (A) Bioactive glass inlay. Reprinted with permission from Ref. [124]. Copyright 2016 Wiley. (B) Zirconia crown restoration [129]. (C) SLA-printed alumina crown. Reprinted with permission from Ref. [130]. Copyright 2017 Elsevier.
Polymers 15 02405 g003
Figure 4. Bioprinting process of patient-specific shaped 3D dentin–pulp complexes. The illustrations show schematic drawings of the (A) 3D bioprinter and (B) printing process to produce patient-specific shaped 3D dentin–pulp complexes. The complex was constructed by serial printing of PCL and two bioinks for dentin and pulp tissue in a layer-by-layer manner. Reprinted with permission from Ref. [232]. Copyright 2019 SAGE Publications.
Figure 4. Bioprinting process of patient-specific shaped 3D dentin–pulp complexes. The illustrations show schematic drawings of the (A) 3D bioprinter and (B) printing process to produce patient-specific shaped 3D dentin–pulp complexes. The complex was constructed by serial printing of PCL and two bioinks for dentin and pulp tissue in a layer-by-layer manner. Reprinted with permission from Ref. [232]. Copyright 2019 SAGE Publications.
Polymers 15 02405 g004
Table 1. The differences between 3D printing and 4D printing.
Table 1. The differences between 3D printing and 4D printing.
3D Printing4D Printing
TechnologySLA, DLP, FDM, SLS, photopolymer jetting, powder binder printer, and computed axial lithography.FDM, SLA, DLP, direct ink writing, inkjet [12].
Suitable materialsThermoplastics, metals, ceramics, biomaterials, or nanomaterials [13].Self-assembled materials, multi-materials, designed materials [13].
ApplicationsJewelry, toys, fashion, entertainment, automobile, aerospace, defense, biomedical devices, etc.Soft robots, grippers, drug delivery, stent and tissue engineering, etc. [12]
AdvantagesHigh material utilization and the ability to manufacture a single complex geometry [6].The precise configuration of material responsiveness [12].
DisadvantagesTime-consuming post-processing, static microstructure, limited layer-by-layer printing speed [12].Slow response rate and low efficiency [12].
Table 2. Polymers for 3D printing in dentistry.
Table 2. Polymers for 3D printing in dentistry.
AbbreviationFull Name of PolymerApplicable
3D Printing Technologies
CharacteristicsClinical Applications
PCLPolycaprolactoneFDMSuperior biocompatibility and adjustable degradabilityTissue engineering scaffolds [27,29,30]
PMMAPolymethyl methacrylateFDM/SLAEasy to manufacture, lower cost, and stable in the oral environmentBone cement and screw fixation, temporary crowns and bridges, obturators, retainers, and denture base material [42,43,44]
PLAPolylactic acidFDMBiodegradable and environmentally friendlyAbsorbable fracture internal fixation material, guided bone/tissue regeneration barrier membrane, and biological scaffold [51,54]
PLGAPoly(lactic-co-glycolic acid)FDMBiodegradability and biocompatibilityTissue engineering scaffolds, guided bone regeneration membrane, and drug-delivery carrier [58,59,60]
UV resinUltraviolet resinSLA/DLPHigh curing efficiency, low energy consumption, and low costProtheses in dental applications, retainers, dentures, and retainers [65,66]
Table 3. Metals for 3D printing in dentistry.
Table 3. Metals for 3D printing in dentistry.
AbbreviationMaterialApplicable
3D Printing
Technologies
CharacteristicsClinical Applications
TiTitanium and its alloysSLS/SLM/EBM/DMLS Extremely chemically stable oxide film, excellent mechanical properties, and outstanding biocompatibility, high cost, limited abrasion resistance, and potential toxicity [70] Dental implants and scaffolds, dental crowns, and denture frameworks [69,78,79,80]
Co–CrCobalt–chromium alloySLS/SLM/DMLS Excellent mechanical properties, corrosion resistance, and good porcelain bonding properties, potentially causes allergic reactions [69,95,99,100]RPD framework, 3-unit FPD framework, crowns, cast post and core [24,96,101,102,103,104]
SSStainless-steelSLM Superior physical properties, biocompatibility, excellent bactericidal ability, and low mechanical properties [107,108]Implants and orthodontic components [109]
-Magnesium alloySLM Favorable mechanical properties and biocompatibility, difficult powder preparation [107,108]Implants [111]
Table 4. Ceramics for 3D printing in dentistry.
Table 4. Ceramics for 3D printing in dentistry.
Ceramic MaterialMolecular
Formula
Applicable
3D Printing Technologies
CharacteristicsClinical Applications
Glass-SLA/SLSHigh mechanical strength, low electrical conductivity, high dielectric constant, good mechanical processing properties, chemical resistance, and thermal stability [123]Hypersensitivity therapy, implant coatings, and bone regeneration in periodontal treatment [76]
ZirconiaZrO2SLA/SLSBiocompatible, osteoconductive, high strength, reduced inflammatory response, high internal stress, easily suffers from cracks after sintering and high-volume shrinkage [135,137,138,129]Preformed as root-canal piles, crowns and bridge restorations, and implant abutments [139]
AluminaAl2O3FDM/SLA/SLSHigh mechanical strength, excellent chemical stability, good electrical insulation properties, and easy to fracture [145]Implants, crowns, bridges, veneers, orthodontic brackets, dental composites, and bone cement materials
Table 5. Natural biopolymers for 4D printing in dentistry.
Table 5. Natural biopolymers for 4D printing in dentistry.
Natural
Biopolymers
Representative SEM FiguresApplicable
4D Printing
Technologies
CharacteristicsClinical Applications
Hyaluronic acid (HA)Polymers 15 02405 i001 Reprinted with permission from Ref. [196]Inkjet bioprinting/
Extrusion-based bioprinting (EBB)/
SLA-based bioprinting/
Laser-based bioprinting (LAB)
Biocompatibility, biodegradability, high mechanical properties, and high stability [169] Printing products, promote tissue repair and wound healing, and effectively treat periodontitis and gingivitis [151,175]
CollagenPolymers 15 02405 i002 Reprinted with permission from Ref. [197]Biocompatible and biodegradableCartilage tissue engineering [178]
AgarosePolymers 15 02405 i003 Reprinted with permission from Ref. [198]. Copyright 2019 Wiley.Good gel properties, mechanical properties, biocompatibility, but inability to support cell growth [178,181,182,183]Promote cell growth and enhance the mechanical properties of bone tissue [178,181,182,183]
ChitosanPolymers 15 02405 i004 Reprinted with permission from Ref. [196]Good biocompatibility, renewable ability, but low mechanical strength [185]Widely used in prosthodontics, oral implantology, and endodontics [184]
Alginate Polymers 15 02405 i005 Reprinted with permission from Ref. [199]. Copyright 2018 Elsevier.Good biocompatibility Obtain conventional impressions and print 3D structures
Cellulose Polymers 15 02405 i006 Reprinted with permission from Ref. [200]. Copyright 2006 Taylor & Francis.Hydrophilic, as good insulating and anti-electrostatic properties [180,192,193]Make tissue engineering scaffolds, wound dressings, and used in drug delivery [191]
Table 6. Cells applied in 4D printing in dentistry applications.
Table 6. Cells applied in 4D printing in dentistry applications.
CellsFull Name and SourceRepresentative SEM FiguresTargeted-Differentiated TissuePotential Applications in Dentistry
DPSCsdental pulp stem cellsPolymers 15 02405 i007 Reprinted with permission from Ref. [201]osteoblasts, cementoblasts, odontoblatsalveolar bone regeneration, periodontal ligament regeneration, dental pulp regeneration [202]
SCAPsstem cells from the apical papillaePolymers 15 02405 i008 Reprinted with permission from Ref. [203]odontoblasts, osteoblasts, neural cells, capillary networkendodontic treatment, pulp-dentin regeneration, bone regeneration, angiogenesis, neural regeneration and repair, periodontal tissue regeneration, bioroot engineering [204,205,206,207]
PDLSCsperiodontal ligament stem cellsPolymers 15 02405 i009 Reprinted with permission from Ref. [208]neurons, fibroblasts, osteoblasts, cementoblastsperiodontal ligament regeneration, periodontal regeneration, mandibular defect repair, root regeneration [209]
DFCsdental follicle cellsPolymers 15 02405 i010 Reprinted with permission from Ref. [210]osteoblasts, cementoblasts, and neuronsbone defects, pulp-dentin regeneration, and neural tissue regeneration [211]
GMSCsgingival mesenchymal stem cellsPolymers 15 02405 i011 Reprinted with permission from Ref. [212]osteoblasts, chondrocytes, neuro-like cellsperiodontal ligament regeneration, facial nerve regeneration, osteogenesis, vascular regeneration [213,214]
SHEDsstem cells from human exfoliated deciduous teethPolymers 15 02405 i012 Reprinted with permission from Ref. [215]. Copyright 2017 Wiley.osteoblasts, odontoblasts, neuro-like blasts,pulp regeneration, bone regeneration [216]
hBMSChuman bone marrow stem cellsPolymers 15 02405 i013 Reprinted with permission from Ref. [217]. Copyright 2003 IEEE.osteoblastsalveolar bone regeneration, bone regeneration, gene delivery [218]
HUVECshuman umbilical vein endothelial cellsPolymers 15 02405 i014 Reprinted with permission from Ref. [219]. Copyright 2007 Mary Ann Liebert, Inc.pulp-dentin, microvessels, small blood vesselscardio-vascularized tissue regeneration [220]
MC3T3-E1osteoblast cell precursor MC3T3-E1Polymers 15 02405 i015 Reprinted with permission from Ref. [221]osteoblasts, cementoblasts, bone tissue, nervealveolar bone regeneration [222]
Table 7. Growth factors for 4D printing in dentistry.
Table 7. Growth factors for 4D printing in dentistry.
AbbreviationFull NameRepresentative SEM FiguresCharacteristicsClinical Applications
SDF-1Stromal-derived factor 1Polymers 15 02405 i016 Reprinted with permission from Ref. [243]Soluble signaling molecules
Combine with printing materials to selectively release to control cell growth, value addition, differentiation, etc. [235,236]
Promote vascular regeneration and bone regeneration [237]
BMP-2Bone morphogenetic proteins 2Polymers 15 02405 i017 Reprinted with permission from Ref. [239]. Copyright 2023 Wiley.Induce bone regeneration and bone healing
BMP-7Bone morphogenetic proteins 7Polymers 15 02405 i018 Reprinted with permission from Ref. [244]
TGF-β1Transforming growth factor-beta1Polymers 15 02405 i019 Reprinted with permission from Ref. [239]. Copyright 2023 Wiley.Stimulate bioactivity and improve bone regeneration [239]
VEGFVascular endothelial growth factorPolymers 15 02405 i020 Reprinted with permission from Ref. [245]Regulate cell migration, angiogenesis, and bone regeneration [238]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cai, H.; Xu, X.; Lu, X.; Zhao, M.; Jia, Q.; Jiang, H.-B.; Kwon, J.-S. Dental Materials Applied to 3D and 4D Printing Technologies: A Review. Polymers 2023, 15, 2405. https://doi.org/10.3390/polym15102405

AMA Style

Cai H, Xu X, Lu X, Zhao M, Jia Q, Jiang H-B, Kwon J-S. Dental Materials Applied to 3D and 4D Printing Technologies: A Review. Polymers. 2023; 15(10):2405. https://doi.org/10.3390/polym15102405

Chicago/Turabian Style

Cai, HongXin, Xiaotong Xu, Xinyue Lu, Menghua Zhao, Qi Jia, Heng-Bo Jiang, and Jae-Sung Kwon. 2023. "Dental Materials Applied to 3D and 4D Printing Technologies: A Review" Polymers 15, no. 10: 2405. https://doi.org/10.3390/polym15102405

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