Chapter 2 - With References

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 33

1.

Literature Review
In this chapter, zirconia, the need for porosity, its fabrication techniques, and the
Direct Ink Writing (DIW) method are elaborated with the help of previously done
research. Firstly, a review of dental implants and the various biomaterials currently
used to manufacture dental implants are given (Section 2.1 -2.2), then an overview
of zirconia, its properties and reason for use as a biomaterial is discussed in Section
2.3. The need for porosity and functionally graded zirconia is provided in Section 2.4
followed by the various fabrication techniques for zirconia implants (Section 2.5 and
2.6) In Section 2.7, the DIW of zirconia is explained in greater detail, including ink
composition, rheology, printing parameters and post-processing conditions.

2.1 Overview of Dental Implants


A dental implant is a surgical device inserted into oral tissues beneath the mucosa
and/or periosteum. It may also be inserted inside or through the bone underneath.
The objective of the implant is to fasten and stabilize the prosthesis i.e., the crown,
maintaining its position and providing support [1].

Dental implants have a history spanning almost two millennia, with the earliest
recorded use traced back to the Egyptian civilization around 2500 BC [2]. During this
period, small prostheses were employed and connected to neighboring teeth using
gold wires. The first instance of dental implants themselves is credited to the Mayans
around 600AD [3]. They utilized shell fragments as implants and substitutes for
mandibular teeth.

In recent decades, significant progress has occurred in the field of biomaterials for
dental implants, accompanied by advancements in clinical applications. Dr.
Brånemark, through his studies in Sweden during the 1960s, played a pivotal role in
making modern dental implantology achievable. He was the first to introduce the
concept of osseointegration [4].

Modern dental implants are categorized into four main types: Root form implants,
Subperiosteal implants, Blade implants, and Trans-osseous implants [1]. Root form
implants are currently the most used form of implant. Root form implants consist of
three key components: the fixture, abutment, and prosthesis [5]. The fixture
resembling a cylinder screw, is inserted into the osseous part of the jaw, serving as
the equivalent of a tooth root. The abutment, elevated from the bone surface to
above the mucosal surface by an abutment screw, is connected to the fixture [6].
Finally, the prosthesis is either screwed on or cemented to the abutment, completing
the implant. These implants exhibit variations in shape, including rounded or pointed
apex, thread spacing, body shape, and other features [7]. A schematic of the root
form implant has been given in Figure 1: “Illustration of 2 root-form implants
supporting a 3-tooth prosthesis. Root-form implant components: Prosthesis screw
(Ps), abutment screw (As), abutment (A), fixture (F)”, Figure adopted from Smith et
al, (1993) [8]Figure 1 .

Figure 1: “Illustration of 2 root-form implants supporting a 3-tooth prosthesis. Root-form


implant components: Prosthesis screw (Ps), abutment screw (As), abutment (A), fixture
(F)”, Figure adopted from Smith et al, (1993) [8]

2.2 Implant materials


The ideal implant material necessitates sufficient mechanical properties i.e., strength,
toughness, wear resistance, corrosion resistance and biocompatibility [8,9]. In the
early twentieth century, dentists employed materials like stainless steel as an implant
material. However, these implants frequently encountered failure, primarily due to
infection and subsequent bone loss [10]. Subsequent extensive research has
explored various materials for implants, and contemporary dental implants are
commonly manufactured from metals, ceramics, or polymers [3]. Table 1 provides a
compilation of materials utilized in the fabrication of endosseous dental implants.
Table 1:Materials used for the fabrication of endosseous dental implants

Implant Material Common Name or Abbreviation


I. Metals
Titanium Ti
Ti-6A1-4V extra low interstitial (ELI)
Ti-6Al-4V
Ti-6Al-7Nb
Titanium Alloys Ti-5Al-2.5Fe
Ti-15 Zr-4Nb-2Ta-0.2Pd
Ti-29Nb-13Ta-4.6Zr
Roxolid (83%–87%Ti-13%–17%Zr)
Stainless Steel SS, 316 LSS
Cobalt Chromium Alloy Vitallium, Co-Cr-Mo
Gold Alloys Au Alloys
Tantalum Ta
II. Ceramics
Alumina Al2O3, polycrystalline alumina or single-
Hydroxyapatite crystal sapphire, HA
Beta-Tricalcium phosphate Ca10(PO4)10(OH)2
β-TCP, Ca3(PO4)2
C
Carbon low-temperature isotropic (LTI)
ultra-low-temperature isotropic (ULTI)
Carbon-Silicon C-Si
Bioglass SiO2/CaO/Na2O/P2O5
Zirconia ZrO2
Zirconia-toughened alumina ZTA
III. Polymers
Polymethylmethacrylate PMMA
Polytetrafluoroethylene PTFE
Polyethylene PE
Polysulfone PSF
Polyurethane PU
Polyether ether ketone PEEK
Adopted from: Rehman et al (2015) [11]

Presently, Titanium and its alloys are considered the "gold standard" for implant
manufacturing [12], primarily owing to titanium's superior biocompatibility and
osteogenic potential compared to various biomaterials [13]. Besides titanium,
ceramics are being explored as a potential alternative.

2.2.1 Titanium and its alloys:


There are six distinct types of Titanium designated by ASTM for use as implant
biomaterials [14]. Among these, four are grades of commercially pure titanium, while
the remaining two are titanium alloys. Ti-6Al-4V, an alloy containing 6% aluminum
and 4% vanadium, is the most utilized grade. This preference is attributed to its
relatively low density, high strength, and resistance to corrosion and fatigue. Despite
being stiffer than bone, leading to stress shielding, its modulus of elasticity is closer
to bone than any other implant material, second only to pure titanium [15]. The low
wear and abrasion resistance of the alloy is overcome through modifications and
treatment of the surface. Furthermore, these surface treatments help improve the
osseointegration as osseointegration mainly depends on surface properties such as
its composition, hydrophilicity, and roughness [16].

Titanium however has the following drawbacks: its visible gray color through peri-
implant mucosa [16], stress shielding, nonspecific immunomodulation and
autoimmunity induced by the metal [18], galvanic side effects after contact with
saliva and fluoride [19], and potential cellular sensitization [20].

2.2.1 Ceramics:
Initially, the use of ceramics in dental implants was confined to coatings on metal-
based implants to enhance osseointegration [21]. This was mainly due to the low
fracture toughness of the ceramics; for instance, aluminium oxide exhibited excellent
osseointegration but lacked the mechanical properties for sustained long-term loads
[22]. Zirconia, particularly yttrium-stabilized tetragonal polycrystalline zirconia (Y-
TZP), has emerged as a prominent ceramic for dental implants due to its superior
strength and fracture resistance compared to other ceramics [23].
Zirconia implants are gaining popularity, particularly for addressing aesthetic
concerns associated with Titanium implants. Zirconia abutments closely match the
color of natural teeth, possess excellent mechanical properties, demonstrate a high
potential for osseointegration and have a high resistance to plaque adhesion [24].
However, zirconia has limitations, necessitating surface modification for reliable use
as an implant material; for instance the smooth surface of zirconia implants,
present without surface roughening procedures, negatively impact
osseointegration. The surface modifications are complex, elevate the cost of
zirconia implants, and introduce the possibility of affecting the implant's mechanical
properties by introducing flaws [25]. Other shortcomins of zirconia include a high
susceptibility to aging [26], poor resin bonding after loading, thermocycling and
lower retention into cores compared to metals [27].

2.2.3 Polymers:
High-Performance Polymers (HPPs), particularly polyetheretherketone (PEEK) from
the polyaryletherketone (PAEK) polymer family, are being considered as implant
materials to address challenges associated with titanium or zirconia use [28]. PEEK is
being considered due to its minimal difference in elasticity compared to bone, its
stress-protective attributes, aesthetics, and biocompatibility [29]. The main drawback
of PEEK is it’s low osseointegration properties compared to Titanium or Zirconia.
Therefore, additional surface modifications, such as Hydroxyapatite (HA)
incorporation [30], TiO2 modification [31], or fluorination and sulfonation [32], are
required for enhanced performance.

2.3 Zirconia as an implant material


2.3.1 Chemistry of zirconia:

Zirconium dioxide (ZrO2), commonly known as zirconia, exhibits polymorphism with


martensitic lattice transformations. The transformations, illustrated in Figure 2, occur
diffusionless and reversibly [33]. Unalloyed zirconia adopts three crystallographic
forms at ambient pressure depending on temperature. At room temperature to
1170°C, it assumes a monoclinic lattice. Between 1170°C and 2370°C, a tetragonal
form is observed, and above 2370°C until its melting point, a cubic lattice structure is
present. Notably, the dilatometer curve displays hysteresis, meaning that the
transformation from monoclinic to tetragonal occurs at 1170°C during heating and
the reverse occurs at 950°C during cooling [34]. The transition from monoclinic to
tetragonal lattice accompanies an increase in density, signifying a reduction in crystal
volume [35]. Similarly, a minimal increase in density occurs when transitioning from
tetragonal to cubic lattice structure.
Figure 2:Lattice transformations of zirconium oxide [34]
To prevent the transformation of Zirconia from the tetragonal phase to the
monoclinic phase and the consequent volume increase, tetragonal phase stabilization
is achieved through selective doping with oxides of magnesium, calcium, yttrium,
and cerium [36]. he most used types of Zirconia are Yttrium cation-doped tetragonal
zirconia polycrystals (3Y-TZP), Magnesium cation-doped partially stabilized zirconia
(Mg-PSZ) and Zirconia toughened alumina (ZTA) [37]. 3Y-TZP comprises nearly
100% tetragonal zirconia (t-ZrO 2) stabilized by 2-3 mol% Y 2O3 [38]. Mg-PSZ exhibits
a coarse-grained microstructure, featuring nanometric t-ZrO 2 precipitates within
micrometer-sized cubic zirconia (c-ZrO 2) grains [39]. Lastly, ZTA consists of 5~30wt
% t-ZrO2 in an Al2O3 ceramic matrix [40]. The mechanical properties of these
ceramics have been summarized in Table 2.

Table 2: Mechanical properties of commercially available zirconia ceramics [38,39,40]


Ceramic Stabilizer Young's Hardness Fracture Density Bending Compression
modulus toughness strength strength
(mol%) (GPa) (GPa) (MPa.m1/2) (g/cm3) (MPa) (MPa)
Mg-PSZ 6-9.5% 200 10-14 4.7-15 5.74–6 450–700 2000
MgO
Y-TZP 2-3% 210 10-12 2.5-15 >6 900–1200 2000
Y2O3
ZTA 10-20% 380 ~17.5 6-12 ≥ 3.97 > 500 4100
ZrO2
Ce-TZP 12-15% 215 6-12 15-30 > 6.20 551 2000
CeO2
(a) (b)

(d)
(c)

Representative microstructures of (a) Y-TZP at a scale of 1µm [41] (b) Mg-PSZ at a scale
of 50µm [42] (c) ZTA at a scale of 1µm [r12], where the white particles are zirconia, and
the black grains are alumina and (d) Ce-TZP at a scale of 1µm [43]

2.3.2 Low Temperature Degradation of zirconia:


Transformation toughening of zirconia refers to the enhancement of the material's
toughness and resistance to crack propagation through a phase transformation.
Specifically, this phenomenon occurs when tetragonal zirconia (t-ZrO2) transforms
into monoclinic zirconia (m-ZrO2) under application of external stress [44]. This
transformation is accompanied by a volume increase (~4.5%), inducing surface
compressive stresses that effectively resist crack propagation.

Low-temperature degradation (LTD), also called aging, occurs similarly, with


metastable t-ZrO2 slowly transforming into stable m-ZrO2 in the presence of water
or water vapor. While a controlled surface tetragonal-monoclinic transformation can
improve the fracture toughness of Y-TZP through the discussed mechanism,
excessive transformation leads to a significant drop in mechanical properties [45].

Figure 3:Illustration of low temperature degradation in zirconia. Figure adopted from


Hjerppe, (2015) [46]

In LTD, the transformation initiates within isolated surface-exposed grains in contact


with water. The transformation in each grain creates compressive stress on
neighbouring grains, generating microcracks (Figure 3). These microcracks facilitate
water penetration, crack propagation, and phase destabilization, making the
degradation autocatalytic [46]. According to Guo et al, The mechanism of LTD is
based on the annihilation of Oxygen vacancies [47]. The number of oxygen
vacancies is given by the Equation (1) where Vo’’ represents Oxygen vacancies.
' X ''
Y 2 O3 Zr O2 2Y Zr +3 Oo + V o
↔ (1)

2.3.3 Ceria stabilized zirconia (Ce-TZP):


It is apparent that reducing oxygen vacancies in the ceramic can decrease
susceptibility to aging [47]. The number of oxygen vacancies in Ce-TZP is given in
Equation (2).
X X
CeO Zr O2 Ce Zr +2 Oo (2)

A comparison of equations 1 and 2 reveals that Ce-TZP has a lower number of


oxygen vacancies compared to Y-TZP, indicating higher resistance to low-
temperature degradation (LTD) than Y-TZP (Figure 4). In addition to reducing
susceptibility to aging, Ce-TZP offers the advantage of a larger critical phase change
grain size compared to Y-TZP, enabling implant production without the need for
ultrafine powder.

Figure 4: Low-temperature degradation kinetics of Ce-TZP, Y-TZP, Mg-PSZ, ATZ and ZTA
measured at 1341°C and expected at 37°C [47]
2.4 Porosity in dental implants:
Various surface modifications have been suggested to enhance the properties of
dental implants [48]. Despite the positive impact of surface modification on implant-
to-bone interaction, the precise mechanism underlying this improvement is not
clearly understood [49]. Additionally, surface modifications present certain
drawbacks, including the gradual dissolution of coating materials, potential negative
biological effects on due to the release of coating particles from the surface [50],
and the associated cost of these coatings.

To overcome these issues, porous biomaterials have been proposed as a solution.


These pores cause an improvement in osseointegration, biocompatibility, and a
reduction in the stress shielding effect [51]. Furthermore, the pores can also act as a
drug delivery system.

2.4.1 Effect of porosity on osseointegration:


Prior research has established that the topography of implant surfaces exerts a more
substantial influence on bone cell behaviour than the composition of the implant
material itself [52]. Dental implants facilitate cell attachment by providing a
conducive area for cell adhesion [53]. Enhanced osteointegration is achieved through
rough surfaces which promote improved adhesion of osteoblasts, cell proliferation,
and the formation of extracellular matrix [52,53].

Porosity enhances osseointegration by increasing the implant's surface area which in


turn facilitates improved cell attachment [49]. Pore parameters such as size, shape,
interconnection, and arrangement, exert a significant impact on osseointegration
[54]. Research by Takizawa et al. (2018) indicated that implants with a pore size of
60–80 μm and a porosity of 30 – 40% exhibit superior bone regeneration compared
to non-porous implants [55]. According to Zhang et al (2018) micropores promote
osseointegration by augmenting the surface area thereby enhancing the absorption
of proteins conducive to bone growth, ion exchange (Ca 2+/PO43-), and apatite
production [56]. As per Li et al (2005) open interconnected pores are more
advantageous compared to closed pores, as they promote bone ingrowth [57].
Furthermore, According to Miao et al (2010) each tissue has its own preferred pore
size for ingrowth [58] i.e., 5–15μm for fibroblast ingrowth, 70–120μm for
chondrocyte ingrowth,100–400μm for bone regeneration [59]. A schematic
representation of the osseointegration process of a porous dental implant has been
provided in Figure 5.
Figure 5:Graphical representation of timeline of osseointegration of dental implants,
reproduced from Uzeda et al (2021) [60]

2.4.2 Effect of porosity on stress shielding:


The mismatch of the elastic modulus between implants and bone tissue leads to
stress shielding of the bone and subsequent failure of the implant [61]. The elastic
modulus of the implant is controlled by its porosity, pore morphology and pore
distribution [62]. Luo et al. (1999) proposed Equation (3) [63] to establish a
relationship between the microstructure and mechanical properties of the implant,
where M0 and M represent the Young’s modulus of dense and porous materials, P is
the volume fraction of porosity, and a is related to morphology and the Poisson ratio
of the matrix.

M o∗(1−P) (3) [63]


M=
(1−a∗P)

The value of a is determined by Equation (4) [63], where ν represents the Poisson
ratio.

(13−15 ν )(1−ν) (4) [63]


a=
(14−10 ν )

Figure 6 shows the drop in the elastic modulus of yttria-stabilized zirconia as the
percentage of porosity increases [63]. Therefore, selecting an appropriate pore size
and porosity percentage can mitigate the elastic modulus mismatch between zirconia
(215 GPa for Ce-TZP) and bone (16 GPa).
Figure 6:The effect of porosity content on the elastic moduli with pore size of 2µm,
adapted from Luo et al (1999) [63]

2.5 Conventional fabrication techniques for porous zirconia


implants
To fabricate porous zirconia, there exists various ceramic processing methods such
as partial sintering [64], replica methods [65], sacrificial templates [66], direct
foaming [67] and freeze casting [68]. The size of pores through different processing
techniques have been shown in Figure 7.

Figure 7:Porosity [%] and average pore [µm] size achieved via different processing routes.
Figure reproduced from Studart et al (2006) [69]

2.5.1 Partial sintering:


Partial sintering involves sintering the green body at temperatures below the
complete densification temperature of 1450°C. The size of the starting powder
determines the pore size, while the degree of partial sintering controls the overall
porosity. To achieve the desired pore size, larger particle sizes (two to five times
larger than the pore) are utilized in the starting powders. This method allows for the
production of ceramics with a porosity of less than 60% [70].
Kocjan et al. (2013) achieved the formation of nanoporous zirconia by sintering
zirconia green bodies in the temperature range of 1000°C-1150°C, employing a
heating rate of 5°C/min and a dwell time of 2 hours [71]. Figure 8 illustrates the
presence of inter-particle and intra-particle pores (highlighted in white in the top
right corner of the microstructures). In the initial phase of sintering, intra-particle
coalescence occurred, leading to a reduction in the number of intra-particle pores
and increase in size of inter-particle pores. As the temperature continued to rise,
densification increased during the intermediate sintering stage, resulting in the
elimination of both types of pores. Figure 9 shows the presence of pores in
microstructure of partially sintered zirconia at 1275°C for 5 hours.
Figure 8:Microstructure of the ion-polished surfaces of the centrifuged Meso-TZ samples
partially sintered for 2 hours at (a) 600 °C, (b) 1000 °C, (c) 1050 °C, (d) 1100 °C and (e)
1150 °C at a scale of 100µm. Figure reproduced from Kocjan et al (2013)[71]

Figure 9:Microstructure of a partially sintered zirconia ceramic after 5 hours at 1,275°C.


Scale of 1µm. Figure reproduced from Nettleship et al (2013) [72]

2.5.2 Replica method:


In the replica method, a porous structure such as a polyurethane (PU) sponge
undergoes coating or immersion in the ceramic solution. This sponge is then
pyrolyzed and sintered to create a porous ceramic which mirrors the pore
morphology of the sponge [73]. This technique is primarily used to produce ceramics
with open cell structures in the range of 150 µm to a few millimetres, with diverse
porosities and pore sizes [74]. The large range of the cell size is attributed to
efficiency limitations in infiltration and excess slurry removal [75].

Various templates, including PU, wood [76], coral [77], and polymeric foams [78],
have been utilized to create porous ceramics. The zirconia slurry mixture is prepared
by ultrasonically dispersing zirconia powder in distilled water [79]. In a study by Zhu
et al. (2015), treated PU foam underwent drying at 80°C for 12 hours, followed by
thermal treatment with heating to 800°C for 1 hour and then 1350°C for 5 hours at a
rate of 2°C/min before cooling to room temperature. These foams were then
sintered at 1400°C for 2.5 hours at a heating rate of 10°C/min, followed by cooling
to 800°C at a rate of 5°C/min and furnace cooling to room temperature [80]. The
microstructure of the ceramic formed through this process has been given in Figure
10.
As depicted in Figure 10, the ceramic scaffolds adopt a structure like the
impregnated foam. Thus, by regulating the pore size in the initial foam, the pore size
of the final porous structure can be controlled. Kim et al. (2008) demonstrated this
by utilizing two types of polyurethane foam templates with 45 pores per inch for
large pores and 60 pores per inch for small pores. This resulted in obtaining spherical
pores with sizes of approximately 500–700 µm and 150–200 µm, respectively [81].

Figure 10:Microstructure of sintered zirconia scaffold using replica method. Scale of 500µm
Figure adapted from Zhu et al (2015) [81]

2.5.3 Sacrificial template:


In this technique, appropriate amounts of sacrificial fugitive are mixed with ceramic
powder. These sacrificial fugitives serve as pore-forming agents that are
subsequently removed from the final ceramic through evaporation or burning before
or during sintering [82]. The classification of sacrificial agents and their examples
have been given in Table 3. The porosity is controlled by the quantity of the
sacrificial template, while the pore morphology is influenced by the morphology of
the agents, particularly when the sizes of the pore-forming agents exceed those of
the ceramic powder [83].
Table 3:Classification and examples of sacrificial agents [82]
Type Example
Polymer beads
Synthetic organic
Organic fibers
Starch
Natural organic Cotton fiber
Cellulose
Nickel
Metallic and inorganic Glass
Fly ash
Water
Liquid Gel
Emulsions

Cruz et al. (2008) employed colloidal processing to prepare nanocrystalline zirconia


with well-defined macropores, utilizing polymethylmethacrylate (PMMA) as a
sacrificial fugitive. They achieved isolated individual pores measuring 0.5–1 μm,
immersed in a matrix of grains sized 200–300 nm, and porosity levels of 10–20%
[84]. The microstructure obtained by Cruz et al (2008) is given in Figure 11.
Figure 11:Microstructure of a sintered body fabricated by gel-casting with 2μm
PMMA templates after sintering at 1000 °C for 2 h. Scale of 10µm. Figure adapted
from Cruz et al (2008) [84]

2.5.4 Direct foaming:

In this technique, bubbles are generated inside the ceramic slurry to generate a
foam structure. This structure is subsequently dried and sintered to yield porous
ceramics [85]. These bubbles are generated by solid, liquid or in-situ gas forming
blowing agents. Examples of the blowing agents are listed in Table 4. The quantity of
the blowing agent added during the process controls the total porosity of the
ceramic. The resulting ceramic possesses solid struts, and a broad range of pore
sizes [86]. This technique is characterized by lower cost and higher ease of
production compared to replica methods or sacrificial template techniques. The main
disadvantage of direct foaming is that it produces ceramics with a wide cell size
distribution and an anisotropic structure [87].
Table 4:Examples of blowing agents used in direct foaming technique [88]
Type of foaming agent Example
Solid CaCO3 powder
Liquid Freons
Pentane
In-situ Gas formers CO2 evolution through oxidation of SiC
H2O formation through cross-linking
reactions in silicone resins

Leng et al. (2022) successfully generated porous zirconia ceramic foams through
direct foaming, achieving foams with a cell size ranging from 83 μm to 450 μm and
an average cell window of 18 μm to 125 μm [Figure 12]. Increasing the solid loading
from 20 vol% zirconia to 27 vol% zirconia resulted in a reduction of the cell size to
119 μm to 285 μm and the average cell window to 32 μm to 53 μm [89].
Figure 12:SEM micrograph of 27 vol% solid loading zirconia foam produced through direct
foaming. Figure adapted from Leng et al (2022) [89]

2.5.5 Freeze casting:

Free casting is defined by Deville as “The process of freezing a liquid suspension,


followed by sublimation of the solidified phase from the solid to the gas under
reduced pressure, and subsequent sintering to consolidate and densify the walls”
[90]. This technique produces materials with a microstructure like that of frozen
solvent crystals. The microstructure consists of unidirectional, elongated, and
continuous channels [91] as shown in Figure 13. In this technique the ceramic
powder is first mixed with the solvent, binder, dispersant and functional additives
such as enzymes. The slurry then undergoes casting, solidification via freezing and
sublimation. The frozen solvent is removed in the sublimation step leading to
formation of voids and pores in the ceramic structure. The porous structure is then
sintered to obtain the final ceramic [92].

Hu et al. (2010) utilized freeze casting to produce highly porous and hierarchical
ceramic supports with a 15 vol% solid loading of zirconia. The ceramic slurry,
comprising of zirconia powder, tert-butyl alcohol as the solvent, 0.5 wt% polyvinyl
butyral as the binder, and an alkali solution as the dispersant, was unidirectionally
frozen, exposed to free-drying under vacuum conditions at -50°C, and then sintered
at 1450°C for 2 hours [93]. The resulting green compacts exhibited unidirectionally
aligned pore channels, with a pore channel size around 100μm and wall thickness
ranging from 30 to 100μm [Figure 13] [94].

Figure 13:Microstrcuture of porous YSZ ceramics with unidirectionally aligned pore


channels: (a) perpendicular section to the cooled plate. Scale of 100µm. Figure adapted
from Hu et al (2010) [94]
2.6 Additive Manufacturing techniques for fabrication of
porous zirconia implants:
conventional methods for producing porous zirconia face challenges such as
irregularly shaped pores, insufficient interconnectivity, poor reproducibility, long
production times, high wear rates for machining tools, low machining accuracy, and
the inability to manufacture complex shapes [95,96]. In response to these
limitations, Additive Manufacturing (AM) techniques have emerged as a potential
solution. However, it is important to note that most AM techniques for ceramic
materials also require a long production time and a high initial investment cost [97].

Various AM techniques are available for processing zirconia ceramics, such as


Selective Laser Sintering (SLS) [98], Selective Laser Melting (SLM) [99],
Stereolithography (SLA) [100], Inkjet Printing (IJP) [101], and Fused Deposition
Modelling (FDM) [102]. As SLS and SLM yield fully dense ceramic parts, they are
unsuitable to produce porous zirconia. To obtain porous zirconia structures ink-based
AM techniques, such as SLA, drop-on-demand inkjet printing, and extrusion-based
direct ink writing (DIW) can be combined with the sacrificial fugitive technique [108].
A schematic of SLA, inkjet printing and DIW have been given in Figure 14 -Figure 16.

According to Guo et al. (2022), DIW stands out as the most used technique for
three-dimensionally fabricating porous ceramics due to its cost-effectiveness and
scalability. In niche cases, drop-on-demand inkjet printing and SLA are applied to
manipulate internal and external hierarchical structures and produce complex
ceramic shapes, respectively [103].

Figure 14:Schematic of powder and slurry based selective laser sintering. Figure adapted
from Zocca et al (2015) [104].
Figure 15:Schematic of Stereolithography. Figure adapted from Zocca et al (2015) [104].

Figure 16:Schematic of Direct Ink Writing. Figure adapted from Zocca et al (2015) [104].

2.6.1 Drop-on-demand inkjet printing:


Drop-on-demand inkjet printing is type of material jetting technique where the ink
droplets are generated and ejected from the printing nozzles only when required,
enabling precise control over the material deposition process [105]. The technique
follows a layer-by-layer approach, where a digital model, created through computer-
aided design (CAD) software, guides the selective deposition of ink droplets based on
the desired pattern. After each layer, there may be a drying or curing step to solidify
the deposited material before proceeding to the next layer. The printed part
undergoes post processing such debinding and sintering to obtain final part. A
schematic of the printing setup is shown in Figure 17.

Zhao et al (2002) printed zirconia parts with vertical walls of 340 ± 15µm thickness
and a spacing of 350 ± 24 µm using drop-on-demand inkjet printing [106]. Figure 18
shows the microstructure of green and sintered parts printed using inks with 62.3 wt
% solid loading of zirconia by Fayazar et al (2020) [107].
Figure 17:Schematic of drop-on-demand inkjet printing system. Figure reproduced from
Chung et al (2005) [108].

Figure 18:Microstructure of zirconia parts printed using the drop-on-demand inkjet


printing process before and after sintering; a - green; b – sintered. Scale of 5µm. Figure
adopted from Fayazfar et al (2020) [109]

2.6.2 Stereo lithography (SLA):


Stereolithography (SLA) is a vat-polymerization technique based on the selective
polymerization of a photopolymer slurry containing ceramic particles [110].
Photopolymers are gels comprising of liquid monomers that harden when exposed to
UV light [111]. The typical process of fabricating a part via SLA is listed below and
illustrated in Figure 19:Illustration of the SLA process. Figure adopted from Osman et
al (2017) [110]Figure 19.

1. Mixing:
40-56 vol% of ceramic powder is dispersed within the photopolymer.
2. Design input:
3D model is sliced into 2D layers using CAD/CAM software.
3. Layer-by-layer printing and curing:
Thin layer of mixture is deposited on a moving film. UV laser projects the
shape of the corresponding 2D layer and cures the slurry. The cured part is
attached to the layer below and the entire green body moves up the tube,
creating space for the next layer to attach. Uncured slurry is recycled and
reused later in the process.
4. Debinding
Binder is removed through dissolution in water and evaporation in furnace.
5. Sintering:
This porous structure undergoes sintering to obtain a finished ceramic part.

Figure 19:Illustration of the SLA process. Figure adopted from Osman et al (2017) [110]

2.6.4 Direct Ink Writing (DIW):


Direct Ink Writing (DIW) is a type of material extrusion AM technique [112]. It was
initially established in 1997 by Cesarani et al. at Sandia national laboratories for the
fabrication of concentrated materials such as ceramic pastes with some portion of
organic binder [113]. It has been used for fabrication of a wide range of materials
such as monolithic or composite ceramics, polymers, and alloys [114]. An illustration
of the DIW process for fabricating a scaffold is shown in Figure 20, it involves the
following steps:

1. Mixing:
The ink for DIW is prepared by mixing ceramic, solvent, binder, dispersant
and the sacrificial fugitive.
2. Design input:
3D model is designed and sliced into 2D layers in the CAD/CAM software.
3. Layer-by-layer printing and curing:
A continuous filament of ink is extruded via a nozzle onto the
substrate/previous layers. The nozzle follows the path of the corresponding
2D layer. As each layer is extruded the nozzle moves up to create space for
the new layer.
4. Drying
Some of the water content in the green body is removed via evaporation in
an oven.
5. Debinding and Sintering:
The printed part undergoes debinding and sintering to obtain a finished
ceramic part.
Figure 20:Illustration of DIW process. Figure reproduced from Peng et al (2018)[115]

2.7 Direct Ink Writing of porous zirconia


Direct Ink Writing (DIW) has been used to produce dense zirconia structures in
multiple studies [116,117]. Li et al (2015) produced zirconia scaffolds through DIW,
utilizing a water-based ink with a solid loading of 70 vol% [118]. Additionally, Peng
et al (2018) achieved parts with over 94% of theoretical density through DIW [119].
To produce porous zirconia structures, the conventional technique of sacrificial
fugitive addition will be combined with DIW. The details regarding the inks, printing
parameters, and post-processing parameters for DIW of zirconia are elaborated
below.

2.7.1 Ink for DIW:


The contents and the rheological parameters of the ink are vital in the DIW
technique. The ink consists of fine ceramic powder particles (10μm – 50μm), binder
materials, a solvent and other suitable additive such as surfactants or dispersants
[120]. The ink must be homogenous and without air bubbles to obtain printing
without defect formation [121]. The content of the ink determines its rheological
behaviour which in turn is the major factor affecting the printability of the ink. DIW
requires the ink to be a non-Newtonian fluid which is highly viscous and exhibits
shear-thinning behaviour [122]. When the ink is still, it is expected to exhibit solid
like behaviour and retain its free-standing 3D shape, wheras during the extrusion
process we expect the ink to exhibit fluid like behaviour [109]. Thus, the ink needs
to show viscoelastic behaviour [122]. An illustration of the various rheology
properties important for ink used in DIW has been given in Figure 21.
The elastic and viscous property of the ink is represented by its storage modulus (G′)
and the loss modulus (G″) respectively. G′ depicts the elastic energy during shear
deformation and thus a high G′, in the order of 10 3 Pa according to Liao et al [123],
enables sufficient shape retention of ink upon extrusion. Only if G″ is lower than G’
at lower shear stress will the elastic behaviour dominate when the ink is still. The
yield stress is denoted by the cross-over point of G’ and G’’ i.e., the ink yields to
shear deformation beyond this point. The stiffness and yield stress determine the
capablity of the ink to retain its shape after extusion, thus these values are needed
to be high[124]. According to Peng et al (2018) the yield stress (σ y) of the ink should
be between 102 – 103 Pa [118].

Figure 21:An illustration of rheology properties important for ink used in DIW. Figure
reproduced from Guo et al (2022) [109]

Water is the predominant solvent employed in preparing inks for Direct Ink Writing
(DIW). Although water offers advantages like cost-effectiveness, scalability, and an
environmentally friendly process, it presents drawbacks, including complexities
during evaporation. Alternatively, organic solvents like toluene, stearic acid, and
methylethylketone, while drying rapidly, pose toxicity concerns [125]. Examples of
inks and the carrier used in DIW has been summarized in Table 5.
Table 5: Examples of inks and the carrier used in DIW. Table adapted from Zhang et
al[124]
Solid loading Carriers and additives
< 38 vol% Water, dispersant, binder and plasticizer
58 vol% Epoxy resin and sodium polyacrylate
60 vol% Ammonium polymethacrylate and cold-water-dispersible methylcellulose
47 vol% Cellulose and ammonium polyacrylate
88 wt% Water and Dolapix CE64
39.5 vol% Water, glycerin and Dolapix A 88
70 wt% Pluronic F-127 hydrogel

2.7.2 Printing Process for DIW:


The DIW setup consists of a syringe, an air pressure pipe connected to a pressure
chamber and a movable stage [Figure 22]. Thus, the printing process parameters are
diameter and shape of the nozzle, printing speed, raster pattern and line width over
height ratio of filament extruded [126]. Shahzad et al (2021) states that the
optimum nozzle diameter to avoid clogging in DIW of ceramics is 400–800μm [127].
The printing speed depends on the rheological parameters of the ink, according to
Peng et al (2018) the shear rate corresponding to that print speed must ensure that
the viscosity of the ink is in the range of 10 –100 Pa/s. Furthermore, they also
recommended a line width over height ratio to be ~1.45 to prevent unnecessary
shearing of filaments below the current layer being printed [128].

Figure 22:Illustration of the working of DIW machine. Figure reproduced from Ovhal et al
(2020) [128]

According to Mahili et al (2021), the most important printing parameters in DIW are
the extrusion rate, nozzle travel speed and the distance between the nozzle and the
piece [129]. The extrusion rate affects the viscosity of the ink due to its shear
thinning behaviour. A higher rate causes the ink to become less viscous and forces
the ink to spread sidewise. A higher nozzle travel speed not only decreases accuracy
but also affects the critical height (h c). Nozzle height (h) is defined as the distance
between the sample and the tip of the nozzle. For h < h c the ink is forced sidewise
as it is being squeezed between the nozzle and the sample. For h > h c the ink is
deposited freely, and the cross-sectional geometry depends only on the rheological
parameters of the ink [130]. The following equation describes how h c is calculated:
Vd (5)
h c=
(Dn∗v d ) ❑
Where Vd is the volume of the extruded ink per unit time, D n is the diameter of the
nozzle and vn is the nozzle speed. Figure 23 shows the effect of critical height, the h c
of the system was found to be 0.80 mm which corresponds to sample 4. The
samples 1-3 and 5-6 show various printing defects [131].

Figure 23:Single-wall cylinders (10 layers) extruded with the h of 0.32mm for cylinder 1,
0.48mm for cylinder 2, 0.64mm for cylinder 3, 0.80mm for cylinder 4, 0.96mm for cylinder
5, and 1.28mm for cylinder 6. Figure reproduced from Smay et al (2002) [131]

2.7.3 Post-Printing Treatment for DIW Ceramics:


After printing the green body undergoes drying, debinding and sintering to increase
density and enhance mechanical properties.

The green bodies must be dried homogeneously and slowly to minimize drying
induced stresses. These stresses occur due to shrinkage, moisture transfer, internal
pressure etc. can cause failure [132]. Tabard et al (2021) recommended drying the
green bodies in a climatic chamber with controlled atmosphere, with the conditions
gradually changing from 98%RH at 24 °C to 30%RH at 60°C over four days [133].

Debinding involves removal of both solvent, such as water or Pluronic 127 hydrogel
(PF127), and the sacrificial fugitive in successive stages. There are different
debinding methods such as the thermal, solvent, and catalytic method [134]. Tabard
et al (2021) used a stepwise debinding stage of heating 0.5°C/min up to 350°C to
eliminate the PF127 followed by heating 2 °C/min up to 600 °C and a dwelling at that
temperature for 2 h to burn out the remaining starch [133]. Thermogravimetric
analysis (TGA) can be used to determine the optimal thermomechanical debinding
cycle. In TGA, the changes in weight of the sample are measured based on the
change in the temperature. The obtained curve gives information about the thermal
stability and decomposition kinetics of the volatile components. Weight loss in the
curve shows decomposition or evaporation[135].

Since the technique chosen for obtaining the porous structure is a hybrid technique
of DIW + Sacrificial fugitive technique the sintering temperate is recommended to be
between 1400°C to 1550°C by Stawarczx1yk et al (2013) [67].

2.7.4 Direct Ink Writing of functionally graded zirconia:

As mentioned in section 2.4.1 and 2.4.2 a high porosity and a large pore size in the
sintered structure will result in better osseointegration. However, it also leads a to
decrease in the mechanical properties of the implant. To overcome this issue,
Porosity graded - Functionally Graded Materials can be used. These materials are
composites that gradually vary in their pore density or pore size over their volume
and as a result, have locally tailored properties [136,137]. A schematic
representation of the porous graded zirconia structure has been given in Figure 24.

Figure 24:Schematic diagrams of FGMs with graded porous layers on a dense core. Figure
adopted from Miao et al (2010)[132]

To obtain functionally graded zirconia, the DIW process can be carried out with two
sets of inks. One ink containing starch as the sacrificial fugitive to produce the
porous structure and the other ink containing no starch to obtain fully dense
structures.

2.7 Conclusion:
Direct ink writing (DIW) using inks with sacrificial fugitives shows the potential to
produced porous zirconia structures due to its speed and high dimensional accuracy.
There are various parameters which need to be optimized to obtain parts with
desired dimensions and mechanical properties.

The rheological properties such as viscosity response to shear, storage modulus, loss
modulus and yield stress of the ink play an important role in the printing process.
The inks must exhibit shear thinning behaviour for the ink to flow easily through the
nozzle and the inks must also be able to retain their shape after extrusion. These
properties can be altered and optimized through manipulation of amount of pluronic,
solid content and dispersant used. In this thesis, based on prior work by Tabard et al
(2021) the ink is a hydrogel prepared using Pluronic (PF -127), dispersant (Dolapix
CE) with a combined solid loading of 40 vol% of Ceria stabilized zirconia (CEZ-12)
and starch (potato or corn).

The printing parameters of the DIW process i.e., nozzle height, nozzle diameter,
nozzle travel speed, extrusion rate, layer interpenetration all play an important role
in affecting the shape of the printed part and can also cause printing defects in the
final part.

Post printing process i.e., drying, debinding and sintering are essential in obtaining
parts with high relative density and mechanical properties. Several defects might
arise during these processes such as drying cracks, surface cracks, bloating and
warping. The temperature, duration and the atmosphere are all important
parameters which need to be optimized to avoid such defects.
Bibliography
[1] R. Gupta, N. Gupta, D.K.K. Weber, Dental Implants, 2023.
[2] C.M. Abraham, A Brief Historical Perspective on Dental Implants, Their Surface
Coatings and Treatments, Open Dent. J. 8 (2014) 50–55.
https://doi.org/10.2174/1874210601408010050.
[3] U. Pasqualini, M.E. Pasqualini, Treatise of Implant Dentistry, 2009.
http://www.ncbi.nlm.nih.gov/pubmed/28125196.
[4] C. Pandey, D. Rokaya, B.P. Bhattarai, Contemporary Concepts in Osseointegration
of Dental Implants: A Review, Biomed Res. Int. 2022 (2022) 1–11.
https://doi.org/10.1155/2022/6170452.
[5] D.G.K. Hong, J. Oh, Recent advances in dental implants, Maxillofac. Plast. Reconstr.
Surg. 39 (2017) 33. https://doi.org/10.1186/s40902-017-0132-2.
[6] J.-J. Kim, J.-H. Lee, J.C. Kim, J.-B. Lee, I.-S.L. Yeo, Biological Responses to the
Transitional Area of Dental Implants: Material- and Structure-Dependent
Responses of Peri-Implant Tissue to Abutments, Materials (Basel). 13 (2019) 72.
https://doi.org/10.3390/ma13010072.
[7] S. Jain, S. Soni, S. Lodhi, R. Khan, A. Jain, B. Khare, B.S. Thakur, P.K. Jain,
Contemporary Trends in Dental Implants, Asian J. Dent. Heal. Sci. 2 (2022) 48–54.
https://doi.org/10.22270/ajdhs.v2i4.25.
[8] D.C. Smith, Dental implants: materials and design considerations., Int. J.
Prosthodont. 6 (1993) 106–17.
[9] G.R. Parr, L.K. Gardner, R.W. Toth, Titanium: the mystery metal of implant
dentistry. Dental materials aspects., J. Prosthet. Dent. 54 (1985) 410–4.
https://doi.org/10.1016/0022-3913(85)90562-1.
[10] D.-W. Kang, S.-H. Kim, Y.-H. Choi, Y.-K. Kim, Repeated failure of implants at the
same site: a retrospective clinical study, Maxillofac. Plast. Reconstr. Surg. 41
(2019) 27. https://doi.org/10.1186/s40902-019-0209-1.
[11] D.F. Williams, Implants in dental and maxillofacial surgery., Biomaterials. 2
(1981) 133–46. https://doi.org/10.1016/0142-9612(81)90039-9.
[12] J.E. Lemons, Dental implant biomaterials., J. Am. Dent. Assoc. 121 (1990) 716–9.
https://doi.org/10.14219/jada.archive.1990.0268.
[13] R.L. Sakaguchi, J.L. Ferracane, J.M. Powers, Craig’s restorative dental materials,
n.d.
[14] K.B. Sagomonyants, M.L. Jarman-Smith, J.N. Devine, M.S. Aronow, G.A. Gronowicz,
The in vitro response of human osteoblasts to polyetheretherketone (PEEK)
substrates compared to commercially pure titanium, Biomaterials. 29 (2008)
1563–1572. https://doi.org/10.1016/j.biomaterials.2007.12.001.
[15] S. Berner, M. Dard, J. Gottlow, A. Molenberg, M. Wieland, Titanium-zirconium: A
novel material for dental implants, Eur. Cells Mater. 17 (2009).
[16] J. W. Nicholson, Titanium Alloys for Dental Implants: A Review, Prosthesis. 2
(2020) 100–116. https://doi.org/10.3390/prosthesis2020011.
[17] M.E. Hoque, N.-N. Showva, M. Ahmed, A. Bin Rashid, S.E. Sadique, T. El-Bialy, H.
Xu, Titanium and titanium alloys in dentistry: current trends, recent
developments, and future prospects, Heliyon. 8 (2022) e11300.
https://doi.org/10.1016/j.heliyon.2022.e11300.
[18] ASTM, Standard Specification for Titanium and Titanium Alloy Castings, ASTM
B367- (n.d.). https://doi.org/10.1520/B0367-06.
[19] M. Niinomi, Mechanical properties of biomedical titanium alloys, Mater. Sci. Eng.
A. 243 (1998) 231–236. https://doi.org/10.1016/S0921-5093(97)00806-X.
[20] G.R.M. Matos, Surface Roughness of Dental Implant and Osseointegration, J.
Maxillofac. Oral Surg. 20 (2021) 1–4. https://doi.org/10.1007/s12663-020-
01437-5.
[21] M. Ferrari, M. Carrabba, A. Vichi, C. Goracci, M. Cagidiaco, Influence of Abutment
Color and Mucosal Thickness on Soft Tissue Color, Int. J. Oral Maxillofac. Implants.
32 (2017) 393–399. https://doi.org/10.11607/jomi.4794.
[22] Z. Özkurt, E. Kazazoǧlu, Zirconia dental implants: A literature review, J. Oral
Implantol. 37 (2011) 367–376. https://doi.org/10.1563/AAID-JOI-D-09-00079.
[23] A. Mellado-Valero, A. Muñoz, V. Pina, M. Sola-Ruiz, Electrochemical Behaviour and
Galvanic Effects of Titanium Implants Coupled to Metallic Suprastructures in
Artificial Saliva, Materials (Basel). 11 (2018) 171.
https://doi.org/10.3390/ma11010171.
[24] P.P. Poli, F.V. de Miranda, T.O.B. Polo, J.F. Santiago Júnior, T.J. Lima Neto, B.R. Rios,
W.G. Assunção, E. Ervolino, C. Maiorana, L.P. Faverani, Titanium Allergy Caused
by Dental Implants: A Systematic Literature Review and Case Report, Materials
(Basel). 14 (2021) 5239. https://doi.org/10.3390/ma14185239.
[25] W.R. Lacefield, Current Status of Ceramic Coatings for Dental Implants, Implant
Dent. 7 (1998) 315–322. https://doi.org/10.1097/00008505-199807040-00010.
[26] N. Cionca, D. Hashim, A. Mombelli, Zirconia dental implants: where are we now,
and where are we heading?, Periodontol. 2000. 73 (2017) 241–258.
https://doi.org/10.1111/prd.12180.
[27] R.-J. Kohal, W. Att, M. Bächle, F. Butz, Ceramic abutments and ceramic oral
implants. An update, Periodontol. 2000. 47 (2008) 224–243.
https://doi.org/10.1111/j.1600-0757.2007.00243.x.
[28] F. Zarone, M.I. Di Mauro, P. Ausiello, G. Ruggiero, R. Sorrentino, Current status on
lithium disilicate and zirconia: a narrative review, BMC Oral Health. 19 (2019)
134. https://doi.org/10.1186/s12903-019-0838-x.
[29] F.H. Schünemann, M.E. Galárraga-Vinueza, R. Magini, M. Fredel, F. Silva, J.C.M.
Souza, Y. Zhang, B. Henriques, Zirconia surface modifications for implant
dentistry, Mater. Sci. Eng. C. 98 (2019) 1294–1305.
https://doi.org/10.1016/j.msec.2019.01.062.
[30] C. Sanon, J. Chevalier, T. Douillard, M. Cattani-Lorente, S.S. Scherrer, L. Gremillard,
A new testing protocol for zirconia dental implants, Dent. Mater. 31 (2015) 15–
25. https://doi.org/10.1016/j.dental.2014.09.002.
[31] A.A. Madfa, F.A. Al-Sanabani, N.H. Al-Qudami, J.S. Al-Sanabani, A.G. Amran, Use of
Zirconia in Dentistry: An Overview, Open Biomater. J. 5 (2014) 1–7.
https://doi.org/10.2174/1876502501405010001.
[32] M.G. Wiesli, M. Özcan, High-Performance Polymers and Their Potential
Application as Medical and Oral Implant Materials: A Review, Implant Dent. 24
(2015) 448–457. https://doi.org/10.1097/ID.0000000000000285.
[33] B. Wang, M. Huang, P. Dang, J. Xie, X. Zhang, X. Yan, PEEK in Fixed Dental
Prostheses: Application and Adhesion Improvement, Polymers (Basel). 14 (2022)
2323. https://doi.org/10.3390/polym14122323.
[34] J.R. Dondani, J. Iyer, S.D. Tran, Surface Treatments of PEEK for Osseointegration to
Bone, Biomolecules. 13 (2023) 464. https://doi.org/10.3390/biom13030464.
[35] T. Kizuki, T. Matsushita, T. Kokubo, Apatite-forming PEEK with TiO2 surface layer
coating, J. Mater. Sci. Mater. Med. 26 (2015) 41. https://doi.org/10.1007/s10856-
014-5359-1.
[36] G.M. WOLTEN, Diffusionless Phase Transformations in Zirconia and Hafnia, J. Am.
Ceram. Soc. 46 (1963) 418–422. https://doi.org/10.1111/j.1151-
2916.1963.tb11768.x.
[37] E.C. Subbarao, H.S. Maiti, K.K. Srivastava, Martensitic transformation in zirconia,
Phys. Status Solidi. 21 (1974) 9–40. https://doi.org/10.1002/pssa.2210210102.
[38] J. Chevalier, L. Gremillard, A. V. Virkar, D.R. Clarke, The tetragonal-monoclinic
transformation in zirconia: Lessons learned and future trends, J. Am. Ceram. Soc.
92 (2009) 1901–1920. https://doi.org/10.1111/j.1551-2916.2009.03278.x.
[39] A.P. Bechepeche, O. Treu, E. Longo, C.O. Paiva-Santos, J.A. Varela, Experimental
and theoretical aspects of the stabilization of zirconia, J. Mater. Sci. 34 (1999)
2751–2756. https://doi.org/10.1023/A:1004698026465.
[40] N.R.F.A. Silva, I. Sailer, Y. Zhang, P.G. Coelho, P.C. Guess, A. Zembic, R.J. Kohal,
Performance of Zirconia for Dental Healthcare, Materials (Basel). 3 (2010) 863–
896. https://doi.org/10.3390/ma3020863.
[41] M.F.R.P. Alves, L.Q.B. de Campos, B.G. Simba, C.R.M. da Silva, K. Strecker, C. dos
Santos, Microstructural Characteristics of 3Y-TZP Ceramics and Their Effects on
the Flexural Strength, Ceramics. 5 (2022) 798–813.
https://doi.org/10.3390/ceramics5040058.
[42] F. Meschke, G. De Portu, N. Claussen, Microstructure and thermal stability of fine-
grained (Y, Mg)-PSZ ceramics with alumina additions, J. Eur. Ceram. Soc. 11
(1993) 481–486. https://doi.org/10.1016/0955-2219(93)90025-M.
[43] M.A. Gafur, M. Al-Amin, M.S.R. Sarker, M.Z. Alam, Structural and Mechanical
Properties of Alumina-Zirconia (ZTA) Composites with Unstabilized Zirconia
Modulation, Mater. Sci. Appl. 12 (2021) 542–560.
https://doi.org/10.4236/msa.2021.1211036.
[44] C. Piconi, G. Maccauro, Zirconia as a ceramic biomaterial, Biomaterials. 20 (1999)
1–25. https://doi.org/10.1016/S0142-9612(98)00010-6.
[45] S. Lawson, Environmental degradation of zirconia ceramics, J. Eur. Ceram. Soc. 15
(1995) 485–502. https://doi.org/10.1016/0955-2219(95)00035-S.
[46] J. Hjerppe, THE INFLUENCE OF CERTAIN PROCESSING FACTORS ON THE
DURABILITY OF YTTRIUM STABILIZED ZIRCONIA USED AS DENTAL
BIOMATERIAL, n.d. https://www.researchgate.net/publication/265073509.
[47] X. Guo, Low temperature degradation mechanism of tetragonal zirconia ceramics
in water: role of oxygen vacancies, Solid State Ionics. 112 (1998) 113–116.
https://doi.org/10.1016/S0167-2738(98)00212-4.
[48] Z.J. Wally, W. van Grunsven, F. Claeyssens, R. Goodall, G.C. Reilly, Porous titanium
for dental implant applications, Metals (Basel). 5 (2015) 1902–1920.
https://doi.org/10.3390/met5041902.
[49] G. Kumar, C.K. Tison, K. Chatterjee, P.S. Pine, J.H. McDaniel, M.L. Salit, M.F. Young,
C.G. Simon, The determination of stem cell fate by 3D scaffold structures through
the control of cell shape, Biomaterials. 32 (2011) 9188–9196.
https://doi.org/10.1016/j.biomaterials.2011.08.054.
[50] C.M. Bidan, K.P. Kommareddy, M. Rumpler, P. Kollmannsberger, P. Fratzl, J.W.C.
Dunlop, Geometry as a Factor for Tissue Growth: Towards Shape Optimization of
Tissue Engineering Scaffolds, Adv. Healthc. Mater. 2 (2013) 186–194.
https://doi.org/10.1002/adhm.201200159.
[51] M. Mour, D. Das, T. Winkler, E. Hoenig, G. Mielke, M.M. Morlock, A.F. Schilling,
Advances in Porous Biomaterials for Dental and Orthopaedic Applications,
Materials (Basel). 3 (2010) 2947–2974. https://doi.org/10.3390/ma3052947.
[52] M. de Wild, R. Schumacher, K. Mayer, E. Schkommodau, D. Thoma, M. Bredell, A.
Kruse Gujer, K.W. Grätz, F.E. Weber, Bone Regeneration by the Osteoconductivity
of Porous Titanium Implants Manufactured by Selective Laser Melting: A
Histological and Micro Computed Tomography Study in the Rabbit, Tissue Eng.
Part A. 19 (2013) 2645–2654. https://doi.org/10.1089/ten.tea.2012.0753.
[53] K. Sarna-Boś, K. Skic, J. Sobieszczański, P. Boguta, R. Chałas, Contemporary
Approach to the Porosity of Dental Materials and Methods of Its Measurement,
Int. J. Mol. Sci. 22 (2021) 8903. https://doi.org/10.3390/ijms22168903.
[54] F.M. Klenke, Y. Liu, H. Yuan, E.B. Hunziker, K.A. Siebenrock, W. Hofstetter, Impact
of pore size on the vascularization and osseointegration of ceramic bone
substitutesin vivo, J. Biomed. Mater. Res. Part A. 85A (2008) 777–786.
https://doi.org/10.1002/jbm.a.31559.
[55] F. Bai, Z. Wang, J. Lu, J. Liu, G. Chen, R. Lv, J. Wang, K. Lin, J. Zhang, X. Huang, The
Correlation Between the Internal Structure and Vascularization of Controllable
Porous Bioceramic Materials In Vivo : A Quantitative Study, Tissue Eng. Part A. 16
(2010) 3791–3803. https://doi.org/10.1089/ten.tea.2010.0148.
[56] A. Rodriguez-Contreras, M. Punset, J.A. Calero, F.J. Gil, E. Ruperez, J.M. Manero,
Powder metallurgy with space holder for porous titanium implants: A review, J.
Mater. Sci. Technol. 76 (2021) 129–149.
https://doi.org/10.1016/j.jmst.2020.11.005.
[57] K. Zhang, Y. Fan, N. Dunne, X. Li, Effect of microporosity on scaffolds for bone
tissue engineering, Regen. Biomater. 5 (2018) 115–124.
https://doi.org/10.1093/rb/rby001.
[58] Y. Zhang, Z. Fan, Y. Xing, S. Jia, Z. Mo, H. Gong, Effect of microtopography on
osseointegration of implantable biomaterials and its modification strategies,
Front. Bioeng. Biotechnol. 10 (2022).
https://doi.org/10.3389/fbioe.2022.981062.
[59] V.K. Balla, S. Bodhak, S. Bose, A. Bandyopadhyay, Porous tantalum structures for
bone implants: Fabrication, mechanical and in vitro biological properties, Acta
Biomater. 6 (2010) 3349–3359. https://doi.org/10.1016/j.actbio.2010.01.046.
[60] X. Miao, Y. Hu, J. Liu, B. Tio, P. Cheang, K.A. Khor, Highly interconnected and
functionally graded porous bioceramics, in: Key Eng. Mater., Trans Tech
Publications Ltd, 2003: pp. 595–598.
https://doi.org/10.4028/www.scientific.net/kem.240-242.595.
[61] G.H. Loh, E. Pei, D. Harrison, M.D. Monzón, An overview of functionally graded
additive manufacturing, Addit. Manuf. 23 (2018) 34–44.
https://doi.org/10.1016/j.addma.2018.06.023.
[62] X. Miao, D. Sun, Graded/Gradient Porous Biomaterials, Materials (Basel). 3
(2009) 26–47. https://doi.org/10.3390/ma3010026.
[63] X. Miao, D. Sun, Graded/gradient porous biomaterials, Materials (Basel). 3 (2010)
26–47. https://doi.org/10.3390/ma3010026.
[64] S.H. Oh, I.K. Park, J.M. Kim, J.H. Lee, In vitro and in vivo characteristics of PCL
scaffolds with pore size gradient fabricated by a centrifugation method,
Biomaterials. 28 (2007) 1664–1671.
https://doi.org/10.1016/j.biomaterials.2006.11.024.
[65] J. Grech, E. Antunes, Zirconia in dental prosthetics: A literature review, J. Mater.
Res. Technol. 8 (2019) 4956–4964. https://doi.org/10.1016/j.jmrt.2019.06.043.
[66] G. Davidowitz, P.G. Kotick, The Use of CAD/CAM in Dentistry, Dent. Clin. North
Am. 55 (2011) 559–570. https://doi.org/10.1016/j.cden.2011.02.011.
[67] B. Stawarczyk, M. Özcan, L. Hallmann, A. Ender, A. Mehl, C.H.F. Hämmerlet, The
effect of zirconia sintering temperature on flexural strength, grain size, and
contrast ratio, Clin. Oral Investig. 17 (2013) 269–274.
https://doi.org/10.1007/s00784-012-0692-6.
[68] H. Nakai, M. Inokoshi, K. Nozaki, K. Komatsu, S. Kamijo, H. Liu, M. Shimizubata, S.
Minakuchi, B. Van Meerbeek, J. Vleugels, F. Zhang, Additively manufactured
zirconia for dental applications, Materials (Basel). 14 (2021) 1–9.
https://doi.org/10.3390/ma14133694.
[69] J. Sun, L. Gao, Influence of forming methods on the microstructure of 3Y-TZP
specimens, Ceram. Int. 29 (2003) 971–974. https://doi.org/10.1016/S0272-
8842(03)00050-6.
[70] Arnaud G, The truth about zirconia, Dent Tech. 80 (n.d.) 59–72.
[71] E. Siarampi, E. Kontonasaki, L. Papadopoulou, N. Kantiranis, T. Zorba, K.M.
Paraskevopoulos, P. Koidis, Flexural strength and the probability of failure of cold
isostatic pressed zirconia core ceramics, J. Prosthet. Dent. 108 (2012) 84–95.
https://doi.org/10.1016/S0022-3913(12)60112-7.
[72] R.A. Lyubushkin, V. V. Sirota, O.N. Ivanov, Fabrication and properties of zirconium
ceramic from zirconium dioxide nanopowder, Glas. Ceram. 68 (2011) 61–64.
https://doi.org/10.1007/s10717-011-9322-z.
[73] V.P. Meshalkin, A. V. Belyakov, Methods Used for the Compaction and Molding of
Ceramic Matrix Composites Reinforced with Carbon Nanotubes, Processes. 8
(2020) 1004. https://doi.org/10.3390/pr8081004.
[74] A. V. Muley, S. Aravindan, I.P. Singh, Nano and hybrid aluminum based metal
matrix composites: an overview, Manuf. Rev. 2 (2015) 15.
https://doi.org/10.1051/mfreview/2015018.
[75] M. Mazaheri, A. Simchi, F. Golestani-Fard, Densification and grain growth of
nanocrystalline 3Y-TZP during two-step sintering, J. Eur. Ceram. Soc. 28 (2008)
2933–2939. https://doi.org/10.1016/j.jeurceramsoc.2008.04.030.
[76] I.-W. Chen, X.-H. Wang, Sintering dense nanocrystalline ceramics without final-
stage grain growth, Nature. 404 (2000) 168–171.
https://doi.org/10.1038/35004548.
[77] A. Kocjan, Z. Shen, Colloidal processing and partial sintering of high-performance
porous zirconia nanoceramics with hierarchical heterogeneities, J. Eur. Ceram.
Soc. 33 (2013) 3165–3176.
https://doi.org/10.1016/j.jeurceramsoc.2013.06.004.
[78] F. Chen, Y. Yang, Q. Shen, L. Zhang, Macro/micro structure dependence of
mechanical strength of low temperature sintered silicon carbide ceramic foams,
Ceram. Int. 38 (2012) 5223–5229.
https://doi.org/10.1016/j.ceramint.2012.03.030.
[79] C.-Y. Lee, S. Lee, J.-H. Ha, J. Lee, I.-H. Song, K.-S. Moon, Effect of the Sintering
Temperature on the Compressive Strengths of Reticulated Porous Zirconia, Appl.
Sci. 11 (2021) 5672. https://doi.org/10.3390/app11125672.
[80] Y. Yang, S. Shimai, Y. Sun, M. Dong, H. Kamiya, S. Wang, Fabrication of porous Al 2
O 3 ceramics by rapid gelation and mechanical foaming, J. Mater. Res. 28 (2013)
2012–2016. https://doi.org/10.1557/jmr.2013.170.
[81] Y. Tang, K. Zhao, L. Hu, Z. Wu, Two-step freeze casting fabrication of
hydroxyapatite porous scaffolds with bionic bone graded structure, Ceram. Int.
39 (2013) 9703–9707. https://doi.org/10.1016/j.ceramint.2013.04.038.
[82] A.R. Studart, U.T. Gonzenbach, E. Tervoort, L.J. Gauckler, Processing Routes to
Macroporous Ceramics: A Review, J. Am. Ceram. Soc. 89 (2006) 1771–1789.
https://doi.org/10.1111/j.1551-2916.2006.01044.x.
[83] G. Carotenuto, Dense/Porous Layered Hydroxylapatite Ceramic for Orthopedic
Device Coating Prepared by Tape Casting Technique., Adv. Perform. Mater. 5
(1998) 171–181. https://doi.org/10.1023/A:1008626330162.
[84] I. Nettleship, Improving Ceramic Processing Through the Application of
Microstructure Mining, Metallogr. Microstruct. Anal. 2 (2013) 372–377.
https://doi.org/10.1007/s13632-013-0098-0.
[85] B. V Manoj Kumar, Y.-W. Kim, Processing of polysiloxane-derived porous
ceramics: a review, Sci. Technol. Adv. Mater. 11 (2010) 044303.
https://doi.org/10.1088/1468-6996/11/4/044303.
[86] P. Colombo, Conventional and novel processing methods for cellular ceramics,
Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 364 (2006) 109–124.
https://doi.org/10.1098/rsta.2005.1683.
[87] A. Herzog, R. Klingner, U. Vogt, T. Graule, Wood-Derived Porous SiC Ceramics by
Sol Infiltration and Carbothermal Reduction, J. Am. Ceram. Soc. 87 (2004) 784–
793. https://doi.org/10.1111/j.1551-2916.2004.00784.x.
[88] R.A. White, J.N. Weber, E.W. White, Replamineform: A New Process for Preparing
Porous Ceramic, Metal, and Polymer Prosthetic Materials, Science (80-. ). 176
(1972) 922–924. https://doi.org/10.1126/science.176.4037.922.
[89] M.R. Nangrejo, M.J. Edirisinghe, Porosity and Strength of Silicon Carbide Foams
Prepared Using Pre-ceramic Polymers, J. Porous Mater. 9 (2002) 131–140.
https://doi.org/10.1023/A:1020834509443.
[90] C.I. Resende-Gonçalves, N. Sampaio, J. Moreira, O. Carvalho, J. Caramês, M.C.
Manzanares-Céspedes, F. Silva, B. Henriques, J. Souza, Porous Zirconia Blocks for
Bone Repair: An Integrative Review on Biological and Mechanical Outcomes,
Ceramics. 5 (2022) 161–172. https://doi.org/10.3390/ceramics5010014.
[91] Y. Zhu, R. Zhu, J. Ma, Z. Weng, Y. Wang, X. Shi, Y. Li, X. Yan, Z. Dong, J. Xu, C. Tang, L.
Jin, In vitro cell proliferation evaluation of porous nano-zirconia scaffolds with
different porosity for bone tissue engineering, Biomed. Mater. 10 (2015) 055009.
https://doi.org/10.1088/1748-6041/10/5/055009.
[92] H.-W. Kim, S.-Y. Shin, H.-E. Kim, Y.-M. Lee, C.-P. Chung, H.-H. Lee, I.-C. Rhyu, Bone
Formation on the Apatite-coated Zirconia Porous Scaffolds within a Rabbit
Calvarial Defect, J. Biomater. Appl. 22 (2008) 485–504.
https://doi.org/10.1177/0885328207078075.
[93] O. Lyckfeldt, J.M.F. Ferreira, Processing of porous ceramics by ‘starch
consolidation,’ J. Eur. Ceram. Soc. 18 (1998) 131–140.
https://doi.org/10.1016/S0955-2219(97)00101-5.
[94] A.A. Madfa, F.A. Al-Sanabani, N.H. Al-Qudami, J.S. Al-Sanabani, A.G. Amran, Use of
Zirconia in Dentistry: An Overview, Open Biomater. J. 5 (2014) 1–7.
https://doi.org/10.2174/1876502501405010001.
[95] H. Santa Cruz, J. Spino, G. Grathwohl, Nanocrystalline ZrO2 ceramics with
idealized macropores, J. Eur. Ceram. Soc. 28 (2008) 1783–1791.
https://doi.org/10.1016/j.jeurceramsoc.2007.12.028.
[96] P. Sepulveda, J.G.. Binner, Processing of cellular ceramics by foaming and in situ
polymerisation of organic monomers, J. Eur. Ceram. Soc. 19 (1999) 2059–2066.
https://doi.org/10.1016/S0955-2219(99)00024-2.
[97] X. Mao, S. Shimai, S. Wang, Gelcasting of alumina foams consolidated by epoxy
resin, J. Eur. Ceram. Soc. 28 (2008) 217–222.
https://doi.org/10.1016/j.jeurceramsoc.2007.06.006.
[98] Q. Leng, D. Yao, Y. Xia, H. Liang, Y.-P. Zeng, Microstructure and permeability of
porous zirconia ceramic foams prepared via direct foaming with mixed
surfactants, J. Eur. Ceram. Soc. 42 (2022) 7528–7537.
https://doi.org/10.1016/j.jeurceramsoc.2022.08.060.
[99] S. Deville, Freeze-Casting of Porous Ceramics: A Review of Current Achievements
and Issues, Adv. Eng. Mater. 10 (2008) 155–169.
https://doi.org/10.1002/adem.200700270.
[100] S. Deville, Freeze-Casting of Porous Biomaterials: Structure, Properties and
Opportunities, Materials (Basel). 3 (2010) 1913–1927.
https://doi.org/10.3390/ma3031913.
[101] C. Gaudillere, J.M. Serra, Freeze-casting: Fabrication of highly porous and
hierarchical ceramic supports for energy applications, Boletín La Soc. Española
Cerámica y Vidr. 55 (2016) 45–54. https://doi.org/10.1016/j.bsecv.2016.02.002.
[102] L. Hu, C.-A. Wang, Y. Huang, C. Sun, S. Lu, Z. Hu, Control of pore channel size
during freeze casting of porous YSZ ceramics with unidirectionally aligned
channels using different freezing temperatures, J. Eur. Ceram. Soc. 30 (2010)
3389–3396. https://doi.org/10.1016/j.jeurceramsoc.2010.07.032.
[103] C.X.. Lam, X.. Mo, S.. Teoh, D.. Hutmacher, Scaffold development using 3D printing
with a starch-based polymer, Mater. Sci. Eng. C. 20 (2002) 49–56.
https://doi.org/10.1016/S0928-4931(02)00012-7.
[104] L.N. Khanlar, A. Salazar Rios, A. Tahmaseb, A. Zandinejad, Additive Manufacturing
of Zirconia Ceramic and Its Application in Clinical Dentistry: A Review, Dent. J. 9
(2021) 104. https://doi.org/10.3390/dj9090104.
[105] S. Nachum, J. Vogt, F. Raether, Additive manufacturing of ceramics:
Stereolithography versus binder jetting, CFI Ceram. Forum Int. 93 (2016) E27–
E33.
[106] X. Zhang, X. Wu, J. Shi, Additive manufacturing of zirconia ceramics: a state-of-the-
art review, J. Mater. Res. Technol. 9 (2020) 9029–9048.
https://doi.org/10.1016/j.jmrt.2020.05.131.
[107] A. Zocca, P. Colombo, C.M. Gomes, J. Günster, Additive Manufacturing of Ceramics:
Issues, Potentialities, and Opportunities, J. Am. Ceram. Soc. 98 (2015) 1983–2001.
https://doi.org/10.1111/jace.13700.
[108] J. Deckers, J. Vleugels, J.P. Kruth, Additive manufacturing of ceramics: A review, J.
Ceram. Sci. Technol. 5 (2014) 245–260. https://doi.org/10.4416/JCST2014-
00032.
[109] Z. Guo, L. An, S. Lakshmanan, J.N. Armstrong, S. Ren, C. Zhou, Additive
Manufacturing of Porous Ceramics With Foaming Agent, J. Manuf. Sci. Eng. 144
(2022). https://doi.org/10.1115/1.4051828.
[110] M. Revilla-León, M.J. Meyer, A. Zandinejad, M. Özcan, Additive manufacturing
technologies for processing zirconia in dental applications A literature review on
current status and future perspectives, 2020.
[111] X. Zhao†, J.R.G. Evans, M.J. Edirisinghe, J.-H. Song, Direct Ink-Jet Printing of
Vertical Walls, J. Am. Ceram. Soc. 85 (2002) 2113–2115.
https://doi.org/10.1111/j.1151-2916.2002.tb00414.x.
[112] J. Chung, S. Ko, C.P. Grigoropoulos, N.R. Bieri, C. Dockendorf, D. Poulikakos,
Damage-Free Low Temperature Pulsed Laser Printing of Gold Nanoinks On
Polymers, J. Heat Transfer. 127 (2005) 724–732.
https://doi.org/10.1115/1.1924627.
[113] H. Fayazfar, F. Liravi, U. Ali, E. Toyserkani, Additive manufacturing of high loading
concentration zirconia using high-speed drop-on-demand material jetting, Int. J.
Adv. Manuf. Technol. 109 (2020) 2733–2746. https://doi.org/10.1007/s00170-
020-05829-2.
[114] R.B. Osman, A.J. van der Veen, D. Huiberts, D. Wismeijer, N. Alharbi, 3D-printing
zirconia implants; a dream or a reality? An in-vitro study evaluating the
dimensional accuracy, surface topography and mechanical properties of printed
zirconia implant and discs, J. Mech. Behav. Biomed. Mater. 75 (2017) 521–528.
https://doi.org/10.1016/j.jmbbm.2017.08.018.
[115] A. Shen, D. Caldwell, A.W.K. Ma, S. Dardona, Direct write fabrication of high-
density parallel silver interconnects, Addit. Manuf. 22 (2018) 343–350.
https://doi.org/10.1016/j.addma.2018.05.010.
[116] A. Shahzad, I. Lazoglu, Direct ink writing (DIW) of structural and functional
ceramics: Recent achievements and future challenges, Compos. Part B Eng. 225
(2021) 109249. https://doi.org/10.1016/j.compositesb.2021.109249.
[117] K. Fu, Y. Wang, C. Yan, Y. Yao, Y. Chen, J. Dai, S. Lacey, Y. Wang, J. Wan, T. Li, Z.
Wang, Y. Xu, L. Hu, Graphene Oxide-Based Electrode Inks for 3D-Printed Lithium-
Ion Batteries, Adv. Mater. 28 (2016) 2587–2594.
https://doi.org/10.1002/adma.201505391.
[118] E. Peng, X. Wei, U. Garbe, D. Yu, B. Edouard, A. Liu, J. Ding, Robocasting of dense
yttria-stabilized zirconia structures, J. Mater. Sci. 53 (2018) 247–273.
https://doi.org/10.1007/s10853-017-1491-x.
[119] Y. Li, L. Li, B. Li, Direct write printing of three-dimensional ZrO2 biological
scaffolds, Mater. Des. 72 (2015) 16–20.
https://doi.org/10.1016/j.matdes.2015.02.018.
[120] B.A.E. Ben‐Arfa, A.S. Neto, I.M. Miranda Salvado, R.C. Pullar, J.M.F. Ferreira,
Robocasting: Prediction of ink printability in solgel bioactive glass, J. Am. Ceram.
Soc. (2018) jace.16092. https://doi.org/10.1111/jace.16092.
[121] A. Gaddam, D.S. Brazete, A.S. Neto, B. Nan, H.R. Fernandes, J.M.F. Ferreira,
Robocasting and surface functionalization with highly bioactive glass of ZrO 2
scaffolds for load bearing applications, J. Am. Ceram. Soc. 105 (2022) 1753–1764.
https://doi.org/10.1111/jace.17869.
[122] J.A. Lewis, J.E. Smay, J. Stuecker, J. Cesarano, Direct Ink Writing of Three-
Dimensional Ceramic Structures, J. Am. Ceram. Soc. 89 (2006) 3599–3609.
https://doi.org/10.1111/j.1551-2916.2006.01382.x.
[123] J. Liao, H. Chen, H. Luo, X. Wang, K. Zhou, D. Zhang, Direct ink writing of zirconia
three-dimensional structures, J. Mater. Chem. C. 5 (2017) 5867–5871.
https://doi.org/10.1039/C7TC01545C.
[124] J. Zhang, M. Yarahmadi, L. Cabezas, M. Serra, S. Elizalde, J.M. Cabrera, L. Llanes, G.
Fargas, Robocasting of dense 8Y zirconia parts: Rheology, printing, and
mechanical properties, J. Eur. Ceram. Soc. 43 (2023) 2794–2804.
https://doi.org/10.1016/j.jeurceramsoc.2022.11.042.
[125] E. Peng, D. Zhang, J. Ding, Ceramic Robocasting: Recent Achievements, Potential,
and Future Developments, Adv. Mater. 30 (2018) 1802404.
https://doi.org/10.1002/adma.201802404.
[126] S.B. Balani, S.H. Ghaffar, M. Chougan, E. Pei, E. Şahin, Processes and materials used
for direct writing technologies: A review, Results Eng. 11 (2021) 100257.
https://doi.org/10.1016/j.rineng.2021.100257.
[127] A. Shahzad, I. Lazoglu, Direct ink writing (DIW) of structural and functional
ceramics: Recent achievements and future challenges, Compos. Part B Eng. 225
(2021) 109249. https://doi.org/10.1016/j.compositesb.2021.109249.
[128] M.M. Ovhal, N. Kumar, J.-W. Kang, 3D direct ink writing fabrication of high-
performance all-solid-state micro-supercapacitors, Mol. Cryst. Liq. Cryst. 705
(2020) 105–111. https://doi.org/10.1080/15421406.2020.1743426.
[129] B. Malihi, Direct ink writing of zirconia for biomedical applications, (n.d.).
[130] M.M. Methani, M. Revilla‐León, A. Zandinejad, The potential of additive
manufacturing technologies and their processing parameters for the fabrication
of all‐ceramic crowns: A review, J. Esthet. Restor. Dent. 32 (2020) 182–192.
https://doi.org/10.1111/jerd.12535.
[131] J.E. Smay, J. Cesarano, J.A. Lewis, Colloidal inks for directed assembly of 3-D
periodic structures, Langmuir. 18 (2002) 5429–5437.
https://doi.org/10.1021/la0257135.
[132] Z.-X. Gong, A.S. Mujumdar, Y. Itaya, S. Mori, M. Hasatani, DRYING OF CLAY AND
NONCLAY MEDIA : HEAT AND MASS TRANSFER AND QUALITY ASPECTS, Dry.
Technol. 16 (1998) 1119–1152. https://doi.org/10.1080/07373939808917457.
[133] L. Tabard, V. Garnier, E. Prud’Homme, E.-J. Courtial, S. Meille, J. Adrien, Y. Jorand,
L. Gremillard, Robocasting of highly porous ceramics scaffolds with hierarchized
porosity, Addit. Manuf. 38 (2021) 101776.
https://doi.org/10.1016/j.addma.2020.101776.
[134] J. González-Gutiérrez, G.B. Stringari, I. Emri, 3 Powder Injection Molding of Metal
and Ceramic Parts, (n.d.). www.intechopen.com.
[135] S.D. Jadhav, J. Vleugels, J.P. Kruth, J. Van Humbeeck, K. Vanmeensel, Mechanical
and electrical properties of selective laser-melted parts produced from surface-
oxidized copper powder, Mater. Des. Process. Commun. 2 (2020) 1–8.
https://doi.org/10.1002/mdp2.94.

You might also like