Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Heliyon 10 (2024) e28063

Contents lists available at ScienceDirect

Heliyon
journal homepage: www.cell.com/heliyon

Research article

Numerical investigation of the optimal porosity of titanium foam


for dental implants
Hussein Farroukh a, *, Fouad Kaddah b, Toufic Wehbe a
a
Mechanical Engineering Department, Saint Joseph University of Beirut, Beirut, 17-5208, Lebanon
b
Civil Engineering Department, Saint Joseph University of Beirut, Beirut, 17-5208, Lebanon

A R T I C L E I N F O A B S T R A C T

Keywords: Background: This paper aims to indicate numerically the accurate porosity used for dental im­
Titanium foam plants, following the emphasis on the preference for titanium foam on pure titanium implants. A
Porosity 3D-optimized numerical model is created to demonstrate the detailed differences between
Dental implant
models.
Finite element analysis
Method: A 3D finite element model was generated using Abaqus for titanium and titanium foam
Mechanical properties
AI (artificial intelligence) implants with different porosities (50,60,62.5,70, and 80%) fixed in cortical and cancellous bone.
The mechanical data for titanium foam is extracted from published literature. We evaluate an
artificial intelligent equation for the stress-strain response of titanium foam with various poros­
ities to describe their variations.
Results: To evaluate the stress-strain variations for different porosities, exponential artificial in­
telligence provides high accuracy (>0.99). The numerical results show that titanium foam im­
plants appear to transfer more loads to the bordering bones due to their lower stiffness and higher
energy absorption, which can help reduce stress shielding problems. In surrounding bones, the
maximum VM stress occurs at the neck region from 5.42 MPa for pure titanium to 21.53 MPa for
titanium foam with 80% porosity. Additionally, a porosity of 62.5% appears to be the most
suitable since Young’s modulus for this porosity (13.82 GPa) is close to the cortical bone’s
modulus (14.5 GPa). This suitability is shown in FEA by the similarity in stress level between pure
titanium and the corresponding porosity. Overall, titanium foam implants appear to be a prom­
ising option for improving the effectiveness and longevity of bone implants in surgical dentistry.
Conclusion: Systematic numerical studies on titanium foam dental implants with different po­
rosities. Analysis of the FE results shows that titanium foam with a porosity of 62.5% is more
beneficial for use in dental implants.

1. Introduction

1.1. Background

Over the past few decades, implantology in dentistry has seen significant advancements, transforming treatment approaches for
oral rehabilitation, and resolving issues with partial and complete edentulism. Titanium implants, which maintain natural teeth

* Corresponding author.
E-mail addresses: hussein.farroukh@net.usj.edu.lb (H. Farroukh), fouad.kaddah@usj.edu.lb (F. Kaddah), toufic.wehbe@usj.edu.lb (T. Wehbe).

https://doi.org/10.1016/j.heliyon.2024.e28063
Received 8 November 2023; Received in revised form 5 March 2024; Accepted 11 March 2024
Available online 13 March 2024
2405-8440/© 2024 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
H. Farroukh et al. Heliyon 10 (2024) e28063

proportions, offer significant benefits over adhesive or mucous prostheses. The continuous connection between prosthetics, implant
research, and industry has led to the incorporation of more diverse components for oral rehabilitation [1]. Surface topography affects
cell shape, orientation, function, and proliferation. The chemical composition of the implant surface affects biocompatibility, and
understanding a structure’s surface properties requires both topographical and chemical techniques [1,2]. Surface treatments have
been applied to titanium implants to adapt their characteristics and increase osteo-conductivity.
Studies show that implant surface qualities are crucial for optimal outcomes, with microscopical features affecting bone formation.
Metal surface treatments aim to create physiologically active surfaces, with retention increasing in developing tissues. Smooth and
rough implant surfaces are divided into additive and subtractive methods, hybrid procedures combining sandblasting and thermal
etching, and acid etching and sanding. Smooth implants can be machined or electropolished, while plasma surface treatment (TPS)
with titanium powders can present a risk of particle detachment [2–4].
Because of titanium’s outstanding mechanical properties, low density, and high chemical resistance, titanium foams have a wide
range of potential uses. These foams are primarily utilized in mechanical applications, such as sandwich materials in submarines or
aircraft, and are widely used for this purpose [5]. This application utilizes titanium’s excellent stiffness and strength relative to its
density, along with its outstanding corrosion resistance [6]. Porous dental implants represent another area of application where ti­
tanium’s exceptional biomechanical properties, especially its resistance to fatigue, are vital [7,8]. The reduced stiffness of titanium
foams represents an advantage for reducing stress-shielding in the bordering bone. Indeed, stress-shielding can lead to bone resorption
and loosening of the implant strength [9–11].
Dental implants can replace a natural tooth’s root to securely anchor dental prostheses. Osseointegration fosters a strong bone-
implant connection by directly linking living bone and implant surfaces, ensuring a durable interlock over time [12,13]. Osseointe­
gration is influenced by bone health, implant design, loadings, and other biological factors [14]. The biological integration between
the implant and tissue results from the porous features improved ability to transmit stress between the implant and the bone. Because
of their larger surface, they also prevent premature mobility and micromotion and offer early mechanical stability. Thus, by providing
a larger fixing surface, osseointegration is promoted. Additionally, a porous biomaterial facilitates vascularization and bone regen­
eration by drawing and allowing cells from the surrounding bone to penetrate its interior [2,4,14].
Fundamentally, implants should be made of biomaterials that are compatible with conditions in the human body. However,
compared to mineralized tissue, these materials have a larger rigidity. A difference between the titanium implant’s Young’s modulus
(110 GPa), natural cortical (13–18 GPa), and cancellous bone (around 2 GPa) might result from the typical non-porous titanium used
for dental implants [15–18].
Kayabasi et al. [19] demonstrates numerically that the high-stress values were located within the cortical bone at the neck of the
implant. Piotrowski et al. [20] discovered that when a low Young’s modulus implant was used, the stress at the boundary between the
cortical bone and the implant decreased notably. This reduces micromotion within the cortical bone-implant interface. Shirazi et al.
[21] showed that the FGBM (functionally graded biomaterial) dental implant can reduce the high-stress values and the stress shielding
effect in natural bones. Salehi et al. [22] indicated analytically that the 50 and 60% porosity of samples of titanium foam were suitable
for biomedical applications. To ensure the ingrowth of bone into the pores, Nouri [23] showed experimentally that the titanium foam
must have at least 60% porosity. Huang et al. [7] found that implants with porous structures can greatly enhance osseointegration
following 12 weeks of bone healing. Yaqoob et al. [24] mechanically approved that Ti foams can reduce the problem of stress shielding.
Ouldyerou et al. [25] show that the 3D printing manufacturing of Ti foam reduces the stress shielding problem and that 77% porosity
of Ti foam transfers more load to the nearby bone than 63%.
A variety of processes, such as foaming, additive manufacturing (AM), and powder metallurgy (PM), can be employed to fabricate
porous Ti structures. The cost-effectiveness and efficiency of PM have resulted in its extensive use across industries for blending and
compacting materials, enabling the production of implant materials with precise shapes, mechanical characteristics, and compositions
[26,27]. Furthermore, the combination of space holder (SH) techniques with powder sintering allows the production of a highly porous
foam structure with an accurately controlled morphology. This method, described by many authors, entails mixing and compacting
space holder particles and metallic powders and then removing the SH during sintering. Magnesium (Mg) is also a viable material for
space holding in the creation of porous titanium scaffolds [28,29].
Previous work demonstrated that the stiffness of porous materials improves by 13.7% and 21.1%, respectively, surpassing that of a
uniform porous medium [30]. Consequently, the production of Ti foam requires high accuracy in the pore distribution [31]. While
most titanium foam manufacturers offer data on porosity, mean average, and pore diameters, they often lack descriptions of me­
chanical properties.
Imwinkelried [32] conducted experimental research on the mechanical characteristics of open-pore Ti-foams made using the
PM/SH technique. Ti-foam with porosity of 50–80% and pores ranging from 100 to 500 μm were investigated thanks to different tests:
static compression, tension, bending, torsion, cyclic compression, and permeability tests. Anisotropic Ti-foam was discovered to have
been utilized in the experiments. Imwinkelried [32] has deduced that the stiffness of titanium foam can be improved by controlling the
porosity and making it vary from 50% (to be close to the cortical bone’s porosity) to 80% for the cancellous bone.

1.2. Aims

In our study, we used an optimized 3D finite element model (Implant, Cortical, and Cancellous Bone) with appropriate boundary
conditions in the Abaqus 6.14 software, where the mechanical properties of Ti-foam with various porosities were extracted from
Imwinkelried experimental tests [32]. The objective of our study can be summarized by:

2
H. Farroukh et al. Heliyon 10 (2024) e28063

• Evaluate the precision of the mechanical properties used by modeling the stress-strain data obtained from a compression test with
an artificially intelligent regression using MATLAB.
• Shown numerically the difference between solid titanium and Ti-foam used in dental implants and explain the importance of
balance between energy absorption and stiffness.
• Determine numerically the most accurate and optimal porosity for titanium foam used in dental implants by modifying the porosity
of titanium foam implants (50, 60, 62.5, 70, and 80%).

The novelty and originality of our work are first to implement results from the in vitro experimental work of Imwinkelried in the
physical Deshpande and Fleck model of Abaqus software. Secondly, we apply it to an optimized 3D finite element model of implants
and bones to investigate the optimal porosity of titanium foam implants used in dental applications and discuss the results from a
biological and mechanical point of view. After the literature review and wide research on Ti foam used in dental implants, we choose
this new idea to verify using FEA and the Deshpande and Fleck model the biomechanical advantages of Ti foam and the optimal
porosity used in dental implants without any expensive experimental work.

2. Materials and methods

2.1. Mechanical properties

2.1.1. Titanium and titanium foam


Their microstructure is the primary distinction between foam materials and solid materials. Imwinkelried experimental data on
titanium foam [32] was used in our study. In this experimental test, the titanium foam was produced using a PM process with a space
holder. The manufacturing process included the following operations: mixing the space holder material with the fine titanium powder
grade four, pressing the resulting body, removing the space holder, and sintering. According to the British Pharmacopoeia BP E503,
Imwinkelried specifies ammonium hydrogen carbonate ((NH4)HCO3) as the SH substance. The SH particles are sieved to a suitable
grain size (425–710 μm). The grain size is selected to provide a final pore size (100–500 μm), which is known to prompt bone for­
mation. In a convection oven operating at 958◦ C for 12 h, the space holder is removed. The components are then moved to the sintering
oven where they are sintered for 3 h at 1300◦ C in an argon atmosphere (400 mbar). By adjusting the quantity of the SH particles, the
overall porosity is changed between 50 and 80% [32]. The cell topology may describe the microstructure of metal foam, such as the
relative density, size, and form of the cells [33]. The percentage of porous area in foams is denoted by the word "porosity" on a
macroscopic scale. Cellular metals’ mechanical behaviors are influenced by their microstructural compositions. Three regimes can be
identified on the stress-strain curve for a porous material under compression: the linear elastic regime, which is related to cell bending;
the plateau regime, which appears with progressive cell collapse due to elastic buckling; and the densification zone, which takes place
when the cells begin to collapse throughout the material [34,35]. Imwinkelried In Vitro experimental data are widely used in Ti foam
research [28,36], because of their manufacturing precision and the high reproducibility and repeatability of testing.
In our numerical study, we considered the data of Ti foam with specific porosity (50, 60, 62.5, 70, and 80%) and a low strain rate of
0.005 s− 1 determined by Imwinkelried [32] with the static compression test. The author extracted the young modulus (E) and yielded
stress (Y) from stress-strain curves. The plastic Poisson’s ratio (υp ) characterizing the specimen’s radial shape, was established at 0.34.
The compressible stress ratio (k) was found to be 0.98.
The goal of optimization is to identify the best solution to a problem within the specified parameters [37,38]. There are two popular
approaches to solving optimization problems: the metaheuristic approach and the mathematical approach. The gradient-dependent

Fig. 1. Stress-strain curve for experimental data and fitting curve with various porosities. a) 70, and 80%, b) 50, 60, and 62.5%.

3
H. Farroukh et al. Heliyon 10 (2024) e28063

mathematical techniques are reliant on the original starting point [39]. High dimensionality, non-convexity, non-linearity, and
complexity are traits of real-world optimization problems [40].
Fig. 1 shows the experimental stress-strain curve with different porosities. To evaluate and describe the mechanical behavior of Ti-
foam for each porosity, we use the regression learner (AI tools on MATLAB R2021b). Like polynomial, Gaussian, Fourier, and other
mathematical optimizations, the exponential regression with two terms represents the best relation to describe the stress-strain
variation of Ti foam [41]. It can be defined by the following equation (1):

σ(ε) = aeb∗ε + ced∗ε (1)

where σ represents the compression stress (MPa) and ε the corresponding strain(mm/mm). Table 1 gives the computed constants a, b,
c, d and R-square for each porosity.
The exponential fitting generates a high precision relation where the R-square is higher than 0.99. Then all porosities of titanium
foam represent the same mathematical variations between compressive stress and strain. The difference between Imwinkelried data
and the correlated curve is represented in Fig. 1a) for porosity 70 and 80%, and Fig. 1b) for others 50, 60, and 62.5%.
Young’s modulus may be calculated using a linear fitting on the elastic zone of the stress-strain variation curve. The yield stress is
the stress at which residual deformation reaches 0.2% of the elastic limit. These mechanical properties of each porosity are filled out in
Table 2, and they will be used in the Deshpande and Fleck model in Abaqus software later.
The energy absorbed per unit volume of the material is related to the surface under the stress-strain curve. Additionally, the energy
absorption is influenced by the cell morphology, base material, foam density, and other factors affecting the length of the visible
plateau in the compression curve of these materials [42]. The energy absorption per unit volume (W) can be described by the equation
(2):
∫ εm
W= σ (ε) dε (2)
0

The energy absorption for each porosity was calculated from the stress-strain data exposed in Fig. 1 using MATLAB. To further
emphasize the accuracy of the mathematical exponential equation, we compared the energy under experimental and predicted stress-
strain curves using the equation (2) and a little difference exists for all porosities (Table 3).
According to Ref. [32], the stiffness of the solid titanium grade 4 is: Young’s modulus E = 110 GPa, Poisson’s ratio ν = 0.33, and the
yield strength σy = 650 MPa.

2.1.2. Cortical and cancellous bone


Cortical and cancellous bones are the two main categories of bones. Compact bones called cortical bones are in charge of mobility,
structural stiffness, and mechanical strength. They represent 80% of the mass of the bones. Cancellous bones are soft, spongy bones
that give the cortical bones structural support, flexibility, and weight loss. In terms of porous structure, cortical bone has a porosity of
less than 10% while cancellous bone has porosities ranging from 50% to 90% [43].
To simplify our numerical simulation, cortical and cancellous bones are assumed to be isotropic, homogenous, and linearly elastic.
This assumption is used due to our simulation goal to show the first chewing function during 0.5 s, where the stress does not exceed the
elastic zone. In real cases, this assumption is not useful to simulate the fatigue of the implant versus time. Those human mechanical
properties are influenced by their mandibular and gum health, genetic makeup, the pace of healing, and a few other variables. As a
result, the information in Table 4 refers to a person with normal tissues [44].

2.2. 3D optimized model

The Finite Element Method (FEM) numerically discretizes a continuous structure into simplified small elements, in order to model
complex mechanical and physical phenomena for engineering applications. It is now a vital tool in several technical domains,
particularly in dental implantology and biomechanics. Here are some reasons why the FEM is important in the dental implant field:
Biomechanical Analysis, Optimization of Implant Design, Predicting Failure Modes, Stress Analysis, and Personalized Treatment
Planning [45].
To initialize our simulations, a 3D finite element model was generated using Abaqus 6.14. Cortical and cancellous bone with
specific implant dimensions was modeled with an optimized shape Fig. 2. The highest level of stress is applied to the top section of the
bone. Suggests that the bone receives a direct signal from the imposed load to the crown [46]. For this reason, to improve our mesh
quality and find the stress distribution from implant to bones with high precision at the neck region, we optimized our implant by

Table 1
The constants a, b, c, and d for exponential model with the corresponding R-square.
Porosity (%) a b c d R-square

50 164 ± 9 2.07 ± 0.2 − 158.3 ± 11 − 82.54 ± 15 0.996


60 108.2 ± 4 2.04 ± 0.1 − 102.3 ± 9 − 111.8 ± 20 0.993
62.5 86.61 ± 1 2.14 ± 0.2 − 86.82 ± 8 − 163.4 ± 22 0.993
70 45.27 ± 2 2.051 ± 0.1 − 44.8 ± 7 − 132.7 ± 33 0.990
80 12.98 ± 1 2.08 ± 0.2 − 12.4 ± 3 − 114.3 ± 35 0.992

4
H. Farroukh et al. Heliyon 10 (2024) e28063

Table 2
The Young modulus and yield stress of Ti foam with various porosities.
Porosity (%) Young modulus E(GPa) Yield stress σy (MPa)

50 40.47 170.63
60 16.32 105.15
62.5 13.82 52.362
70 4.32 25.476
80 0.711 15.3

Table 3
Energy absorption (Mj/ m3 ) for experimental data and fitting curve with different porosities.
Porosity (%) Energy Absorption (Mj/ m3 )

Experimental data Fitting curve


50 141.129 140.991
60 90.462 91.451
62.5 77.038 76.778
70 38.360 38.648
80 11.213 11.296

Table 4
Mechanical properties of bones in the current model [44].
Young’s modulus (GPa) Poisson’s ratio

Cortical bone 14.5 0.323


Cancellous bone 1.37 0.3

suppressing the corresponding thread and replacing that with tie constraints between implants and bones. The specific tie constraint
can be applied to simulate how the implant is anchored or connected to surrounding structures. This constraint can help to accurately
model the implant’s behavior and optimize its thread design accordingly. The other contact between parts is considered frictional with
a coefficient of 0.5 [46]. The implant diameter is 4.1 mm, and the thickness of the cortical bone varies within 1.3–2.0 mm according to
Ref. [47] shown in Fig. 3.
The Deshpande and Fleck model of crushable foam included in Abaqus software is used in our 3D finite element simulation [48].
The simulation utilizes a shape factor parameter to approximate the effect of mean stress on the yield function of foam. This shape
factor is used to distinguish the plastic behavior of foam metals, and it creates an elliptical stress aspect ratio. The yield function (Y) is
represented by equation (3) and presented in Fig. 4 [48,49].
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
Y = q2 + p2 α2 − B (3)

Fig. 2. 3D optimized finite element model.

5
H. Farroukh et al. Heliyon 10 (2024) e28063

Fig. 3. Dimensions (mm) of the 3D model.

where p is the mean stress and q represents the Von Mises equivalent stress. B defines the size of the ellipse and α the shape factor of the
surface.
This specific model is used in our analysis to simulate the model’s response to the loads and boundary conditions defined in
subsequent parts.

2.3. Finite element model

2.3.1. Mesh
In this section, we describe the mesh process of implants, cortical bone, and cancellous bone for the 3D model shown in Fig. 5. The
model is divided into parts and meshed with a global size of 0.3 mm for the implant and 0.7 mm for the bone using quadratic
hexahedron elements (C3D8RH). The quadratic hexahedral elements offer distinct advantages over other mesh forms, such as tetra­
hedron or pyramid elements: higher accuracy in representing complex geometries, volume meshing efficiency, reduced element count,
reduced mesh sensitivity, and compatibility with CAD models.
The resulting mesh has 48026 elements which are of particularly high quality without any poor elements. The mesh quality is
assessed and optimized through specific parameters like geometric deviation factor, average aspect ratio, minimum edge length, and
maximum angle of quadrilateral faces. These parameters are summarized in Table 5.
Since the FEA only provides an approximate solution, a convergence analysis of mesh is conducted to ensure that the mesh quality
doesn’t distort the provided results. The mesh convergence study shows that doubling the mesh resolution in peak stress-prone areas
has less than 2% impact on results. We expose this slight difference by reporting the maximum von Mises stress (MPa) of the whole
model concerning the number of elements in the case of a titanium foam implant with 80% porosity (Fig. 6). This porosity was chosen
because it’s the most affected by the refinement of elements due to the high deformations obtained for this high porosity 80% seen in
the results section.

Fig. 4. Yield surfaces and flow potential of crushable foam model [49].

6
H. Farroukh et al. Heliyon 10 (2024) e28063

Fig. 5. Meshed model.

Table 5
Meshing parameters and quality of the model.
Number of elements Avg Aspect ratio Avg deviation factor Avg Max angle Avg Min edge length

Bone 46316 1.14 8.01 ∗10 − 4 94 6.55 ∗10− 4


Implant 1710 1.75 0.0367 107 3.3 ∗10− 4

2.3.2. Boundary conditions


The lower surfaces of the cortical bone are considered fixed supports (zero DOF, (Ui = URi = 0; i = 1,2,3)), and the dynamic load
simulates average chewing forces of 17.1 N in the lingual direction (-X), 114.6 N in the axial direction (-Y), and 23.4 N in the
mesiodistal direction (+Z) (Fig. 7) [19]. These forces are combined to produce a masticatory force of 118.2 N at a roughly 75◦ angle to
the occlusal plane. The force magnitudes and acting points were selected based on Mericske-Stern’s clinical work. The simulation is an
explicit dynamic simulation with a period of 0.5 s, representing the first chewing phase based on Kayabasi et al.(Fig. 8) [19]. To take
into account the worst case, the force applied during chewing is the one used to model the strongest loads in dental applications. It’s
crucial to underline that each person’s chewing force differs depending on variables like biting strength, dental alignment, and jaw
function.

3. Results

To find the distribution of stress and strain in dynamic dental applications, five criteria can be used: Von Mises (VM), Tresca,
maximum principal (S1), minimum principal (S3), and hydrostatic pressure. Since the latest three presented a variety of unusual
biomechanical stress, only Tresca and VM criteria displayed biomechanically appropriate stress [50]. The fact that foam materials can
undergo plastic deformation in the plateau regime justifies the use of VM stress, which is more accurate and appropriate for our study.

Fig. 6. Convergence study: influence of the number of elements on the maximum stress.

7
H. Farroukh et al. Heliyon 10 (2024) e28063

Fig. 7. Load and boundary conditions.

Fig. 8. Dynamic load in 0.5 s [19].

3.1. Implants

Fig. 9 (a-f), shows the Von Mises stress distribution along implants for pure titanium and Ti-foam with different porosities. The
stress distributions are inhomogeneous and depend on the direction of masticatory forces. The maximum concentration of stress occurs
near the neck region. In the case of pure titanium, the stress is distributed along the length of the implant more evenly than in the case
of Ti-foam with various porosities, where the stress is negligible after the neck region. Maximum stress increases with porosity, rising
from 24.2 MPa with 50% porosity to 46.5 MPa with 80% porosity (Table 6). We observe that the stress level between pure titanium and
Ti-foam with 62.5% porosity is approximately the same. In Fig. 9 (g), the comparison between the yield stress and the maximum stress
on the implant shows that where the porosity is higher than 70%, the maximum stress exceeds the yield stress value. So, the risk of
implant fracture occurs.
Fig. 10 shows the magnitude of deformation of titanium and Ti-foam implants with different porosities. Maximum deformations in
pure titanium implants are greater than Ti-foam with 50, 60, and 62.5% porosity, but less than those with porosities of 70 and 80%. We
notice that, in cases of 50, 60, and 80% porosities, the deformation remains negligible away from the neck region in the -Y direction.
The magnitude of deformations increases with porosity.

3.2. Cortical bone

Fig. 11 presents the Von Mises stress distribution within the surrounding cortical bones in the case of solid titanium and Ti-foam
implants. Stresses in titanium foam implants with different porosities are higher than those in solid titanium, especially at the neck

8
H. Farroukh et al. Heliyon 10 (2024) e28063

Fig. 9. Von Mises stress (Pa) distribution in titanium and Ti-foam implant with different porosities; a) Pure titanium b) P = 50%, c) P = 60%, d) P =
62.50%, e) P = 70%, f) P = 80%; (g) Comparison between Max von Mises stress and Yield stress for each porosity.

Table 6
The maximum value of Von Mises stresses (MPa) for each part of the model with different porosities.
Porosity (%)
Titanium 50 60 62.5 70 80
Von Mises stress (MPa) Implant 24.71 24.28 24.6 24.81 25.63 46.58
Cortical bone 5.427 8.236 8.357 8.463 9.619 21.53
Cancellous bone 0.53 0.45 0.476 0.498 0.592 1.16

region where the maximum value occurs. That means more loads can be transferred to the cortical bone in the case of foam implants.
The maximum stress at the neck region of cortical bone increases with porosity from 8.2 MPa for 50% porosity to 21.5 MPa for 80%
(Table 6). In another region of surrounding cortical bone away from the neck interaction surface, a little difference in stress was shown.
The distribution of stress was higher in the cervical line at the bone-implant interaction surface for all porosities; this area has
experienced higher levels of stress.
The porosity of Ti foam affects the stress distribution in the cortical bone due to the differentiation in the mechanical properties of
the implants. In the case of 80% porosity, the distribution of stress is higher than others due to their lower young modulus (0.71 GPa).
The magnitude of deformation of cortical bones presents a little difference between solid titanium and Ti-foam with different
porosities, as shown in Fig. 12.

3.3. Cancellous bone

Fig. 13 shows the Von Mises stress distribution within the surrounding cancellous bones in the case of solid and foam titanium
implants. The maximum stress for pure titanium implants occurs at the bottom face contact between implants and cancellous bone,

9
H. Farroukh et al. Heliyon 10 (2024) e28063

Fig. 10. Magnitude of deformation (m) for titanium and Ti-foam implant with different porosities; a) Pure titanium b) P = 50%, c) P = 60%, d) P =
62.50%, e) P = 70%, f) P = 80%.

with a small value of 0.53 MPa. In the case of Ti-foam implants, the maximum stress occurs at the top contact face between cortical and
cancellous bone, ranging from 0.45 MPa for 50% porosity to 1.16 MPa for 80% (Table 6). In other regions, we show a little difference,
and we notice that the stress transferred to the cancellous bone is very small.
Due to the low transferred stress, the magnitude of deformation of cancellous bones presents a small value with little difference
between solid and foam titanium implants with different porosities, as shown in Fig. 14.

4. Discussion

The solid titanium is stiffer than the titanium foam, for this reason, the Ti-foam implants transferred more loads to the surrounding
bones (Fig. 11). So, it’s more beneficial to use titanium foam to reduce the probability of a stress shielding problem. This can cause the
bone to weaken and deteriorate over time due to the lack of mechanical stimulation [21,51]. Furthermore, solid titanium has a higher
Young’s modulus than mineralized tissues. As a result, it causes insufficient loading of the underlying bone tissue and stress shielding,
which eventually leads to implant loosening, bone resorption, and failure.
Then, to choose the accurate porosity for Ti-foam dental implants, a balance should be found between stiffness and energy ab­
sorption. Titanium foam implants with porosities between 70 and 80 % present a high magnitude of deformation compared to lower
foam porosities, and this can affect the implant connections with surrounding bones because of their low stiffness (Fig. 10). So clin­
ically the use of Ti foam dental implants with porosity higher than 70% can cause a fracture of implants under natural masticatory
loadings, or disconnection with the surrounding bones due to their low stiffness and energy absorption (Table 3). A simulation
depicting the actual fracture of an implant is observable within a fatigue-versus-time Finite Element Model (FEM). While the use of
porosities 50, 60, and 62.5% is more interesting because of the smaller gap in stress distributions.
We show that the stress leveling of Ti-foam with a porosity of 62.5% and solid titanium implants are approximately the same, with
an equal maximum value and location of von Mises stress (Fig. 9). This similarity is due to the rapprochement of Young’s modulus
between the cortical bone (14.5 GPa) and Ti-foam with this specific porosity (14.8 GPa) according to Table 2. For this reason, the Ti-
foam implants with a porosity of 62.5% are a great solution among the tested porosities to be used for dental applications. In clinical
practice, the young’s modulus of the employed implant is close to 14.5 GPa, which corresponds to the stiffness of cortical bone, in order

10
H. Farroukh et al. Heliyon 10 (2024) e28063

Fig. 11. Von Mises stress (Pa) distribution in Cortical bone for titanium and Ti-foam implants with different porosities; a) Pure titanium b) P = 50%,
c) P = 60%, d) P = 62.50%, e) P = 70%, f) P = 80%.

to lessen stress shielding problems without sacrificing implant stiffness and lifetime.
Because of the high stiffness of cortical bone compared to cancellous bone, the stress transferred to the cancellous surrounding bone
is very low [52].
In the FEM with a dynamic explicit problem, we used a hexahedron structure mesh with an appropriate element quality applied to
an optimized shape of dental osseointegration. We obtained very interesting and promising results based on these numerical im­
provements, with an accuracy of less than 2% after the convergence study. However, we have some limitations, like that the bone
properties are considered to be isotropic and homogeneous, but the biological tissue is anisotropic and porous.
Surface chemistry influences protein, bacteria, and cell adhesion to implants. Dental implant companies use hydrophilic surfaces,
promoting better osseointegration as cells move differently on hydrophilic surfaces [53].
Even though solid titanium implants present some disadvantages in osseointegration the surface treatment can improve their
biomechanical ability. Surface treatment is crucial for improving titanium mechanical properties (Anodic Oxidation, Chemical Vapor
Deposition (CVD)), bioactivity, and osseointegration (Sandblasting/Grit-Blasting, Acid-Etching, Alkaline Treatment, Acid Treatment),
antibacterial effect (Antibiotic and Nonantibiotic Organic Coatings - Inorganic Antimicrobial Coatings) [1–3,54].
From a biological point of view, the porous titanium foam facilitates vascularization and bone regeneration by drawing and
allowing cells from the neighboring bone tissue to penetrate its interior [2,26]. To alleviate the issue of stress concentrations within the
pores, one can employ silanization of the pore walls, followed by filling them with poly-methylmethacrylate (PMMA) [54].
High manufacturing precision is necessary to achieve the exact porosity required for titanium foam dental implants. Powder
metallurgy provides several advantages when it’s combined with the SH. Powder metallurgy techniques are simpler to industrialize,
less expensive, and less consuming time in comparison to rapid prototyping methods such as SLM or 3D printing [55]. For titanium, its
strong chemical reactivity with surrounding mold materials and gases, and high melting point make solid foaming using powder
metallurgy more appealing than liquid foaming methods. Additionally, the pores are randomly distributed and exhibit a wide range of
diameters [56]. This could be seen as a drawback compared to the SLM method, which provides more precise control over the dis­
tribution of pores. Albert Barba et al. [57] demonstrated that only foams with spherical macro-pores significantly promoted ectopic
bone growth, contrary to those with prismatic macropores created by 3D printing. This unique property makes the PM/SH

11
H. Farroukh et al. Heliyon 10 (2024) e28063

Fig. 12. Magnitude of deformation (m) in Cortical bone for titanium and titanium foam implants with different porosities; a) Pure titanium b) P =
50%, c) P = 60%, d) P = 62.50%, e) P = 70%, f) P = 80%.

combination an extremely attractive approach for manufacturing porous metal for osteoregeneration.
Then, PM/SH is an appropriate approach to reduce these homogeneity problems of Ti-foam. The ammonium bicarbonate used in
the Imwinkelried experiment is the best SH because it can be completely and easily removed. Its moderate decomposition temperature
ensures minimal uptake of impurities like carbon, nitrogen, and oxygen [58]. The mechanical properties of the specimen are directly
influenced by its manufacturing imperfections. This is due to the uneven geometry of the pores and pre-existing crack tips. Some
studies have shown that samples with irregular pores experience yielding at lower stress compared to samples with regularly shaped
pores [36].
Previous experimental and numerical works only show an acceptable interval of porosity used in dental implants [7,20–23], but the
novelty of our study is to indicate the accurate porosity of Ti foam at 62.5% using an optimized FEM with the intermediary Deshpande
and Fleck model.

5. Conclusions and future works

In the current study, we extracted the mechanical properties of Ti-foam used in dental implants from published data with different
porosities (50, 60, 62.5, and 80%). We describe the stress-strain response with an artificially intelligent exponential fitting. A good
agreement is obtained with high porosity, especially between the elastic and plastic zones. Then, in Abaqus software, we generate a 3D
explicit finite element model with implants, cortical, and cancellous bone using crushable foam sections. Our goal was to numerically
determine the more adequate porosities used in dental implants after applying the corresponding masticatory forces.
The numerical results show that the titanium foam transferred more loads than solid titanium to the surrounding bones due to its
low stiffness and high energy absorption. Then the porous foam reduces the stress shielding problems. The 62.5% porosity of Ti-foam is
more interesting to be used than other tested porosities because of the similarity of stress leveling with solid titanium implants.
Based on these findings, our work demonstrates that Ti foam dental implants can significantly improve implant success rates and
patient outcomes. It can guide in developing calculation methodology to enable manufacturers and surgeons to select the efficient
implant with the best porosity depending on bone quality and stiffness range, leading to better clinical outcomes and patient
satisfaction.
In the future, new precise methods of manufacturing, such as 3D printing, can reduce the production time and cost of this type of

12
H. Farroukh et al. Heliyon 10 (2024) e28063

Fig. 13. Von Mises stress distribution (Pa) in Cancellous bone for titanium and titanium foam implants with different porosities; a) Pure titanium b)
P = 50%, c) P = 60%, d) P = 62.50%, e) P = 70%, f) P = 80%.

implant without sacrificing accuracy or mechanical characteristics.


In our future works, we will enhance these dynamic numerical simulations (full and partial osseointegration) by incorporating the
actual geometry of the titanium foam implant with various porosity levels, and the inhomogeneous structure of bone with anisotropic
properties. In addition, we will develop a new AI mathematical relation between mechanical properties and fatigue behavior of im­
plants. This will enable a comprehensive analysis of the fatigue behavior of thread connections under varying load configurations.

Ethics approval

The submitted manuscript is original and has not been published elsewhere in any form or language.

Data availability

All data are available on request to the corresponding author.

Funding

This research received no external funding.

CRediT authorship contribution statement

Hussein Farroukh: Writing – review & editing, Writing – original draft, Visualization, Validation, Software, Methodology,
Investigation, Formal analysis, Data curation, Conceptualization. Fouad Kaddah: Writing – review & editing, Writing – original draft,
Validation, Supervision, Project administration, Formal analysis, Conceptualization. Toufic Wehbe: Writing – review & editing,
Writing – original draft, Visualization, Validation, Supervision, Methodology, Formal analysis, Conceptualization.

13
H. Farroukh et al. Heliyon 10 (2024) e28063

Fig. 14. Magnitude of deformation (m) in Cancellous bone for titanium and Ti-foam implants with different porosities; a) Pure titanium b) P = 50%,
c) P = 60%, d) P = 62.50%, e) P = 70%, f) P = 80%.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

References

[1] G. Cervino, A. Meto, L. Fiorillo, A. Odorici, A. Meto, C. D’Amico, et al., Surface treatment of the dental implant with Hyaluronic acid: an overview of recent data,
Int. J. Environ. Res. Publ. Health 18 (2021) 4670, https://doi.org/10.3390/ijerph18094670.
[2] M. Cicciù, L. Fiorillo, A.S. Herford, S. Crimi, A. Bianchi, C. D’Amico, et al., Bioactive titanium surfaces: interactions of Eukaryotic and Prokaryotic cells of nano
devices applied to dental practice, Biomedicines 7 (2019) 12, https://doi.org/10.3390/biomedicines7010012.
[3] L. Laino, M. La Noce, L. Fiorillo, G. Cervino, L. Nucci, D. Russo, et al., Dental pulp stem cells on implant surface: an in vitro study, BioMed Res. Int. 2021 (2021)
e3582342, https://doi.org/10.1155/2021/3582342.
[4] L. Li, J. Shi, K. Zhang, L. Yang, F. Yu, L. Zhu, et al., Early osteointegration evaluation of porous Ti6Al4V scaffolds designed based on triply periodic minimal
surface models, Journal of Orthopaedic Translation 19 (2019) 94–105, https://doi.org/10.1016/j.jot.2019.03.003.
[5] Z. Li, B. Zhao, J. Shao, S. Liu, Deformation behavior and mechanical properties of periodic topological Ti structures fabricated by superplastic forming/diffusion
bonding, International Journal of Lightweight Materials and Manufacture 2 (2019) 1–30, https://doi.org/10.1016/j.ijlmm.2018.11.001.
[6] S. Lascano, C. Arévalo, I. Montealegre-Melendez, S. Muñoz, J.A. Rodriguez-Ortiz, P. Trueba, et al., Porous titanium for biomedical applications: evaluation of the
conventional powder metallurgy frontier and space-holder technique, Appl. Sci. 9 (2019) 982, https://doi.org/10.3390/app9050982.
[7] C.-C. Huang, M.-J. Li, P.-I. Tsai, P.-C. Kung, S.-Y. Chen, J.-S. Sun, et al., Novel design of additive manufactured hollow porous implants, Dent. Mater. 36 (2020)
1437–1451, https://doi.org/10.1016/j.dental.2020.08.011.
[8] H.J. Haugen, H. Chen, Is there a better biomaterial for dental implants than Titanium?-A review and meta-study analysis, J. Funct. Biomater. 13 (2022) 46,
https://doi.org/10.3390/jfb13020046.
[9] J. Li, J.A. Jansen, X.F. Walboomers, J.J. van den Beucken, Mechanical aspects of dental implants and osseointegration: a narrative review, J. Mech. Behav.
Biomed. Mater. 103 (2020) 103574, https://doi.org/10.1016/j.jmbbm.2019.103574.
[10] S. Ahmad, S. Mujumdar, V. Varghese, Role of porosity in machinability of additively manufactured Ti-6Al-4V, Precis. Eng. 76 (2022) 397–406, https://doi.org/
10.1016/j.precisioneng.2022.04.010.
[11] A. Ouldyerou, A. Merdji, L. Aminallah, S. Roy, H. Mehboob, M. Özcan, Biomechanical performance of Ti-PEEK dental implants in bone: an in-silico analysis,
J. Mech. Behav. Biomed. Mater. 134 (2022) 105422, https://doi.org/10.1016/j.jmbbm.2022.105422.
[12] B. Yang, A. Irastorza-Landa, P. Heuberger, H.-L. Ploeg, Analytical model for dental implant insertion torque, J. Mech. Behav. Biomed. Mater. 131 (2022)
105223, https://doi.org/10.1016/j.jmbbm.2022.105223.
[13] T. Fujii, R. Murakami, N. Kobayashi, K. Tohgo, Y. Shimamura, Uniform porous and functionally graded porous titanium fabricated via space holder technique
with spark plasma sintering for biomedical applications, Adv. Powder Technol. 33 (2022) 103598, https://doi.org/10.1016/j.apt.2022.103598.

14
H. Farroukh et al. Heliyon 10 (2024) e28063

[14] H. Mehboob, A. Ouldyerou, M.F. Ijaz, Biomechanical investigation of patient-specific porous dental implants: a finite element study, Appl. Sci. 13 (2023) 7097,
https://doi.org/10.3390/app13127097.
[15] M. Özcan, C. Hämmerle, Titanium as a reconstruction and implant material in dentistry: advantages and pitfalls, Materials 5 (2012) 1528–1545, https://doi.
org/10.3390/ma5091528.
[16] A. Nouri, P.D. Hodgson, C. Wen, A. Nouri, P.D. Hodgson, C. Wen, Biomimetic Porous Titanium Scaffolds for Orthopedic and Dental Applications, IntechOpen,
2010, https://doi.org/10.5772/8787.
[17] B.V. Krishna, S. Bose, A. Bandyopadhyay, Low stiffness porous Ti structures for load-bearing implants, Acta Biomater. 3 (2007) 997–1006, https://doi.org/
10.1016/j.actbio.2007.03.008.
[18] V.K. Dommeti, S. Roy, S. Pramanik, A. Merdji, A. Ouldyerou, M. Özcan, Design and development of tantalum and strontium ion doped hydroxyapatite composite
coating on titanium substrate: structural and human osteoblast-like cell viability studies, Materials 16 (2023) 1499, https://doi.org/10.3390/ma16041499.
[19] O. Kayabaşı, E. Yüzbasıoğlu, F. Erzincanlı, Static, dynamic and fatigue behaviors of dental implant using finite element method, Adv. Eng. Software 37 (2006)
649–658, https://doi.org/10.1016/j.advengsoft.2006.02.004.
[20] B. Piotrowski, A.A. Baptista, E. Patoor, P. Bravetti, A. Eberhardt, P. Laheurte, Interaction of bone–dental implant with new ultra low modulus alloy using a
numerical approach, Mater. Sci. Eng. C 38 (2014) 151–160, https://doi.org/10.1016/j.msec.2014.01.048.
[21] H. Asgharzadeh Shirazi, M.R. Ayatollahi, A. Asnafi, To reduce the maximum stress and the stress shielding effect around a dental implant–bone interface using
radial functionally graded biomaterials, Comput. Methods Biomech. Biomed. Eng. 20 (2017) 750–759, https://doi.org/10.1080/10255842.2017.1299142.
[22] A. Salehi, F. Barzegar, H. Amini Mashhadi, S. Nokhasteh, M.S. Abravi, Influence of pore characteristics on electrochemical and biological behavior of Ti foams,
J. Mater. Eng. Perform. 26 (2017) 3756–3766, https://doi.org/10.1007/s11665-017-2829-x.
[23] A. Nouri, 5 - titanium foam scaffolds for dental applications, in: C. Wen (Ed.), Metallic Foam Bone, Woodhead Publishing, 2017, pp. 131–160, https://doi.org/
10.1016/B978-0-08-101289-5.00005-6.
[24] K. Yaqoob, I. Amjad, M.A. Munir Awan, U. Liaqat, M. Zahoor, M. Kashif, Novel method for the production of titanium foams to reduce stress shielding in
implants, ACS Omega 8 (2023) 1876–1884, https://doi.org/10.1021/acsomega.2c02340.
[25] A. Ouldyerou, L. Aminallah, A. Merdji, A. Mehboob, H. Mehboob, Finite element analyses of porous dental implant designs based on 3D printing concept to
evaluate biomechanical behaviors of healthy and osteoporotic bones, Mech. Adv. Mater. Struct. 30 (2023) 2328–2340, https://doi.org/10.1080/
15376494.2022.2053908.
[26] X. Wang, F. Zhang, Bioactive submicron-pore design of microarc oxidation coating on Ti6Al4V alloy prepared by selective laser melting method, Surf. Coating.
Technol. 444 (2022) 128696, https://doi.org/10.1016/j.surfcoat.2022.128696.
[27] X. Yan, Q. Li, S. Yin, Z. Chen, R. Jenkins, C. Chen, et al., Mechanical and in vitro study of an isotropic Ti6Al4V lattice structure fabricated using selective laser
melting, J. Alloys Compd. 782 (2019) 209–223, https://doi.org/10.1016/j.jallcom.2018.12.220.
[28] O. Cetinel, Z. Esen, B. Yildirim, Fabrication, morphology analysis, and mechanical properties of Ti foams manufactured using the space holder method for bone
substitute materials, Metals 9 (2019) 340, https://doi.org/10.3390/met9030340.
[29] Y. Hou, Y. Li, X. Cai, C. Pan, J. Wang, W. Zhang, et al., Mechanical response and response mechanism of AlSi10Mg porous structures manufactured by laser
powder bed fusion: experimental, theoretical and numerical studies, Mater. Sci. Eng., A 849 (2022) 143381, https://doi.org/10.1016/j.msea.2022.143381.
[30] S. Ruiz de Galarreta, R.J. Doyle, J. Jeffers, S. Ghouse, Laser powder bed fusion of porous graded structures: a comparison between computational and
experimental analysis, J. Mech. Behav. Biomed. Mater. 123 (2021) 104784, https://doi.org/10.1016/j.jmbbm.2021.104784.
[31] H. Singh, P. Saxena, Y.M. Puri, The manufacturing and applications of the porous metal membranes: a critical review, CIRP Journal of Manufacturing Science
and Technology 33 (2021) 339–368, https://doi.org/10.1016/j.cirpj.2021.03.014.
[32] Mechanical Properties of Open-pore Titanium Foam - Imwinkelried - 2007 - Journal of Biomedical Materials Research Part, A - Wiley Online Library, 2023.
https://onlinelibrary.wiley.com/doi/abs/10.1002/jbm.a.31118.
[33] A. Nouri, X.B. Chen, P. Hodgson, C.E. Wen, Preparation and characterisation of new titanium based alloys for orthopaedic and dental applications, Adv. Mater.
Res. 15–17 (2007) 71–76, https://doi.org/10.4028/www.scientific.net/AMR.15-17.71.
[34] L.J. Gibson, M.F. Ashby, Cellular Solids: Structure and Properties, second ed., Cambridge University Press, Cambridge, 1997 https://doi.org/10.1017/
CBO9781139878326.
[35] M.F. Ashby, T. Evans, N.A. Fleck, J.W. Hutchinson, H.N.G. Wadley, L.J. Gibson, Metal Foams: A Design Guide, Elsevier, 2000.
[36] M. Shbeh, E. Oner, A. Al-Rubaye, R. Goodall, Production and digital image correlation analysis of titanium foams with different pore morphologies as a bone-
substitute material, Adv. Mater. Sci. Eng. 2019 (2019) e1670837, https://doi.org/10.1155/2019/1670837.
[37] G. Hu, Y. Guo, G. Wei, L. Abualigah, Genghis Khan shark optimizer: a novel nature-inspired algorithm for engineering optimization, Adv. Eng. Inf. 58 (2023)
102210, https://doi.org/10.1016/j.aei.2023.102210.
[38] M. Ghasemi, M. Zare, A. Zahedi, M.-A. Akbari, S. Mirjalili, L. Abualigah, Geyser inspired algorithm: a new geological-inspired Meta-heuristic for real-parameter
and constrained engineering optimization, J Bionic Eng (2023), https://doi.org/10.1007/s42235-023-00437-8.
[39] J.O. Agushaka, A.E. Ezugwu, L. Abualigah, Dwarf Mongoose optimization algorithm, Comput. Methods Appl. Mech. Eng. 391 (2022) 114570, https://doi.org/
10.1016/j.cma.2022.114570.
[40] G. Hu, Y. Zheng, L. Abualigah, A.G. Hussien, DETDO: an adaptive hybrid dandelion optimizer for engineering optimization, Adv. Eng. Inf. 57 (2023) 102004,
https://doi.org/10.1016/j.aei.2023.102004.
[41] P. Glaister, Exponential curve fitting with least squares, Int. J. Math. Educ. Sci. Technol. 38 (2007) 422–427, https://doi.org/10.1080/00207390601128099.
[42] B. Xie, Y.Z. Fan, T.Z. Mu, B. Deng, Fabrication and energy absorption properties of titanium foam with CaCl2 as a space holder, Mater. Sci. Eng., A 708 (2017)
419–423, https://doi.org/10.1016/j.msea.2017.09.123.
[43] X.-Y. Zhang, G. Fang, J. Zhou, Additively manufactured scaffolds for bone tissue engineering and the prediction of their mechanical behavior: a review,
Materials 10 (2017) 50, https://doi.org/10.3390/ma10010050.
[44] A. Merdji, B. Bachir Bouiadjra, T. Achour, B. Serier, B. Ould Chikh, Z.O. Feng, Stress analysis in dental prosthesis, Comput. Mater. Sci. 49 (2010) 126–133,
https://doi.org/10.1016/j.commatsci.2010.04.035.
[45] G. Cervino, L. Fiorillo, A.V. Arzukanyan, G. Spagnuolo, P. Campagna, M. Cicciù, Application of bioengineering devices for stress evaluation in dentistry: the last
10 years FEM parametric analysis of outcomes and current trends, Minerva Stomatol. 69 (2020) 55–62, https://doi.org/10.23736/S0026-4970.19.04263-8.
[46] Y.C. Chikr, B. Boutabout, A. Merdji, K. Bouzouina, Numerical analysis of the equivalent stress at the interface bone/Threaded dental implant, Journal of
Biomimetics, Biomaterials and Biomedical Engineering 34 (2017) 82–93, https://doi.org/10.4028/www.scientific.net/JBBBE.34.82.
[47] L. Kong, Y. Zhao, K. Hu, D. Li, H. Zhou, Z. Wu, et al., Selection of the implant thread pitch for optimal biomechanical properties: a three-dimensional finite
element analysis, Adv. Eng. Software 40 (2009) 474–478, https://doi.org/10.1016/j.advengsoft.2008.08.003.
[48] V.S. Deshpande, N.A. Fleck, Isotropic constitutive models for metallic foams, J. Mech. Phys. Solid. 48 (2000) 1253–1283, https://doi.org/10.1016/S0022-5096
(99)00082-4.
[49] T.H. Reddy, S. Pal, K.C. Kumar, M.K. Mohan, V. Kokol, Finite element analysis for mechanical response of magnesium foams with regular structure obtained by
powder metallurgy method, Procedia Eng. 149 (2016) 425–430, https://doi.org/10.1016/j.proeng.2016.06.688.
[50] R.A. Moga, C.D. Olteanu, B.M. Daniel, S.M. Buru, Finite elements analysis of tooth-A comparative analysis of multiple failure criteria, Int. J. Environ. Res. Publ.
Health 20 (2023) 4133, https://doi.org/10.3390/ijerph20054133.
[51] X. Lu, S.D. Rawson, P.J. Withers, Effect of hydration and crack orientation on crack-tip strain, crack opening displacement and crack-tip shielding in elephant
dentin, Dent. Mater. 34 (2018) 1041–1053, https://doi.org/10.1016/j.dental.2018.04.002.
[52] G. Osterhoff, E.F. Morgan, S.J. Shefelbine, L. Karim, L.M. McNamara, P. Augat, Bone mechanical properties and changes with osteoporosis, Injury 47 (2016)
S11–S20, https://doi.org/10.1016/S0020-1383(16)47003-8.
[53] M.M. El-Saies, I. Hassan, A.H. El-Shazly, M.T. El-Wakad, Titanium foam for dental implant applications: a review, SVU-International Journal of Engineering
Sciences and Applications 4 (2023) 107–112, https://doi.org/10.21608/svusrc.2023.197066.1110.

15
H. Farroukh et al. Heliyon 10 (2024) e28063

[54] Y. Kirmanidou, M. Sidira, M.-E. Drosou, V. Bennani, A. Bakopoulou, A. Tsouknidas, et al., New Ti-alloys and surface modifications to improve the mechanical
properties and the biological response to orthopedic and dental implants: a review, BioMed Res. Int. 2016 (2016) e2908570, https://doi.org/10.1155/2016/
2908570.
[55] N. Tuncer, G. Arslan, E. Maire, L. Salvo, Investigation of spacer size effect on architecture and mechanical properties of porous titanium, Mater. Sci. Eng., A 530
(2011) 633–642, https://doi.org/10.1016/j.msea.2011.10.036.
[56] B. Dabrowski, W. Swieszkowski, D. Godlinski, K.J. Kurzydlowski, Highly porous titanium scaffolds for orthopaedic applications, J. Biomed. Mater. Res. B Appl.
Biomater. 95B (2010) 53–61, https://doi.org/10.1002/jbm.b.31682.
[57] A. Barba, A. Diez-Escudero, Y. Maazouz, K. Rappe, M. Espanol, E.B. Montufar, et al., Osteoinduction by foamed and 3D-printed calcium phosphate scaffolds:
effect of nanostructure and pore architecture, ACS Appl. Mater. Interfaces 9 (2017) 41722–41736, https://doi.org/10.1021/acsami.7b14175.
[58] 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.

16

You might also like