You are on page 1of 21

Journal Pre-proof

Biomechanical effects of dental implant diameter, connection type,


and bone density on microgap formation and fatigue failure: A finite
element analysis

Hyeonjong Lee , Minhye Jo , Gunwoo Noh

PII: S0169-2607(20)31696-5
DOI: https://doi.org/10.1016/j.cmpb.2020.105863
Reference: COMM 105863

To appear in: Computer Methods and Programs in Biomedicine

Received date: 29 July 2020


Accepted date: 17 November 2020

Please cite this article as: Hyeonjong Lee , Minhye Jo , Gunwoo Noh , Biomechanical effects of
dental implant diameter, connection type, and bone density on microgap formation and fatigue
failure: A finite element analysis, Computer Methods and Programs in Biomedicine (2020), doi:
https://doi.org/10.1016/j.cmpb.2020.105863

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier B.V.


1

Highlights

 Three-dimensional dental implant finite element models with vertical and

oblique loadings

 Influence of connection types, implant diameters, and bone densities on

microgap formation and fatigue life

 Tissue-level connection type-induced low microgap formation and high

resistance to fatigue failure


2

Biomechanical effects of dental implant diameter, connection type, and bone density on microgap

formation and fatigue failure: A finite element analysis

Hyeonjong Lee, DMD, PhD,a,† Minhye Jo, BS,b, † and Gunwoo Noh, PhDc

a
Assistant Professor, Department of Prosthodontics, Dental Research Institute, Dental and Life Science Institute,

School of Dentistry, Pusan National University, Yangsan, Korea.

b
Graduate Student, School of Mechanical Engineering, Korea University, Seoul, Korea

c
Assistant Professor, School of Mechanical Engineering, Kyungpook National University, Daegu, Korea


Hyeonjong Lee and Minhye Jo contributed equally as first authors

Corresponding author:

Dr Gunwoo Noh

School of Mechanical Engineering, Kyungpook National University

1370 Sankyuk Dong, Buk-gu, Daegu 702-701, Republic of Korea

Email: gunwoonoh@gmail.com

Telephone: 82-53-950-5575; Fax: 82-53-950-6550


3

ABSTRACT

Background and Objective. Understanding fatigue failure and microgap formation in dental implants,

abutments, and screws under various clinical circumstances is clinically meaningful. In this study, these aspects

were evaluated based on implant diameter, connection type, and bone density.

Methods. Twelve three-dimensional finite element models were constructed by combining two bone densities

(low and high), two connection types (bone and tissue levels), and three implant diameters (3.5, 4.0, and 4.5

mm). Each model was composed of cortical and cancellous bone tissues, the nerve canal, and the implant

complex. After the screw was preloaded, vertical (100 N) and oblique (200 N) loadings were applied. The

relative displacements at the interfaces between implant, abutment, and screw were analyzed. The fatigue lives

of the titanium alloy (Ti–6Al–4V) components were calculated through repetitive mastication simulations.

Mann–Whitney U and Kruskal–Wallis one-way tests were performed on the 50 highest displacement values of

each model.

Results. At the implant/abutment interface, large microgaps were observed under oblique loading in the buccal

direction. At the abutment/screw interface, microgap formation increased along the implant diameter under

vertical loading but decreased under oblique loading (p < 0.001); the largest microgap formation occurred in the

lingual direction. In all cases, the bone-level connection induced larger microgap formation than the tissue-level

connections. Moreover, only the bone-level connection models showed fatigue failure, and the minimum fatigue

life was observed for the implant diameter of 3.5 mm.

Conclusions. Tissue-level implants possess biomechanical advantages compared to bone-level ones. Two-piece

implants with diameters below 3.5 mm should be avoided in the posterior mandibular area.

Keywords: dental implant, microgap, fatigue failure, finite element analysis

Abbreviations: IAI, implant/abutment interface; FEM, finite element analysis; IT, internal tissue-level; IB,

internal bone-level.

1. INTRODUCTION

The success of a dental implant depends on multiple factors such as the prosthetic contour, oral hygiene,

masticatory force, and biotype of tissue. A critical factor is the microgap formation at the implant/abutment

interface (IAI) under loading conditions since it not only leads to screw loosening but also allows infiltration of
4

micro-organisms into the IAI, resulting in contamination by bacteria and acidic compounds as well as cell

inflammation [1-3]. Durability of the dental implant under repeated loading is also an important factor since the

implant is exposed to nearly 200,000 cycles of masticatory load in a year [4]. The details of fatigue failure of

dental implants can be revealed through computational simulation.

Regarding the microgap analysis of dental implants at the IAI, many in vitro studies had been

performed [1-3,5]; however, several studies only measured the gap distance under preloaded conditions due to

screw tightening [6-8]. Via high-resolution radiographic imaging, various researchers observed the formation of

a 22 mm microgap at the IAI under a lateral force of 100 N [1,8-10]. The radiographic images measured by

Zipprich et al. revealed 0–18.6 and 4.8–42 mm microgaps for, conical and flat connections under 30° oblique

loading, respectively [2].

Previously, several studies regarding microgap formation had been simulated through finite element

analysis (FEA) [11-13]. Saidin et al. evaluated the microgap formation and von Mises stress of four different

implant/abutment connections under a 360 N 15° oblique loading for an implant diameter of 4.5 mm, observing

1.22 mm microgaps in the internal conical connection; however, they did not consider the preloading [11]. Li et

al. investigated the correlation between FEA and in vitro performance for the micromotion of the IAI under

loading conditions, concluding that FEA is an appropriate method to evaluate microgap formation in implants

and abutments [12].

Fatigue failure experiments of dental implants have also been performed [14-20]. The mechanical

aspects of fatigue failure should be examined especially for short implants since their long crown height space

leads to high stress concentrations [21,22]. Some studies have demonstrated that prosthetic complications do not

differ between standard and short implants [23].

Although there have been various studies regarding microgap formation and fatigue failure of dental

implants as mentioned above, there is still a lack of detailed and comprehensive information. To the authors’

best knowledge, a comprehensive FEA study to reveal the correlation between stress, lever effect, and

mechanical failure based on the connection type, implant diameter, and bone density has not been perfo rmed.

The purpose of this study is to evaluate the microgap formation and fatigue failure in dental implants via FEA

for different diameters, connections, and bone qualities.

2. MATERIALS AND METHODS

Twelve three-dimensional finite element models were constructed by considering three parameters with
5

different values: two levels of bone density (low, high), two connection types (internal tissue-level IT, internal

bone-level IB), and three implant diameters (3.5 mm, 4.0 mm, and 4.5 mm) (Table 1). Each model represents a

mandibular bone section of the molar region and includes both the nerve canal and the implant complex. The

bone tissue is composed of cancellous bone in the center, surrounded by 2 mm of cortical bone, and the

cylindrical bone part near the implant has been modeled. The implant complex, simulated based on an Osstem

implant, consists of the crown, cement layer, abutment, screw, and implant (Fig. 1). The implant length was 10

mm and the cement layer had a thickness of 0.03 mm [24]. All the models were built by commercial modeling

software (Mimics 19.0, Materialise Group).

Table 1. Specifications of the finite element models (IT: internal tissue; IB: internal bone).

Cancellous Connection
Implant diameter (mm)
bone density type

Low IT 3.5

4.0

4.5

IB 3.5

4.0

4.5

High IT 3.5

4.0

4.5

IB 3.5

4.0

4.5
6

Fig. 1. Finite element models: (a) dimensions of the implant complexes with the internal tissue (IT ) and internal bone (IB)

level connections and (b) three-dimensional model of the implant complex.

Table 2 summarizes the material properties of the model components [22,25-30]. The implant,

abutment, and screw were made of a titanium alloy (Ti–6Al–4V) with a yield strength of 847 MPa [30]. All

materials were considered linearly elastic, homogenous, and isotropic. Table 3 lists the number of elements used

for each model. The implant/bone interface was defined as a “tie” to achieve complete osseointegration; the

interfaces between abutment, implant, screw, and cement layer were defined as “contact” to analyze the

displacements within the implant complex (Fig. 2). The friction coefficients were 0.441, 0.16, and 0.25 for the

abutment/screw, abutment/implant, and implant/cement layer interfaces, respectively [31,32]. The boundary

conditions were fixed along all axes on the mesial and distal surfaces of the cortical and cancellous bones.

The analysis consisted of two steps. First, an appropriate preloading was applied to the screw to

simulate its tightening process for a torque of 32 N∙cm (Fig. 3a) [6,32]; the preloading was calculated with the

formula reported by Bickford [33]. Second, external loadings were applied along the vertical and oblique

directions to simulate masticatory loading (Fig. 3b). In particular, a loading of 200 N was applied to each of the

60 nodes on the three cusps and three fossae in the vertical direction, and another of 100 N was exerted on each

of the 30 nodes on the three cusps in the oblique direction [24,34,35].

Table 2. Mechanical properties of the finite element model components.


7

Young’s
Material Poisson’s ratio Reference
modulus (MPa)

Crown 140,000 0.28 [26]

T itanium alloy 110,000 0.34 [22,30]

Cement 10,760 0.35 [27]

Cortical Bone 13,700 0.3 [25]

Cancellous Bone

Low Density 259 0.3 [29]

High Density 3,507 0.3 [29]

Nerve Canal 70 0.45 [28]

Table 3. Numbers of tetrahedral elements of each finite element model component .

Cortical
Implant Cancellous
cylinder Cortical Cancellous Crown Cement Abutment Implant Screw Nerve
system cylinder part
part

IT 3510 17,372 159,284 132,128 330,683 57,212 173,234 183,492 139,662 102,420 5061

IT 4010 19,905 151,315 144,173 352,834 58,510 180,932 181,634 193,826 102,421 4993

IT 4510 25,597 146,389 180,052 350,615 57,909 183,694 181,926 210,784 101,168 5061

IB3510 18,654 177,047 127,515 357,532 53,883 142,646 286,475 97,065 155,269 5022

IB4010 20,659 151,326 157,806 340,797 61,137 217,898 409,568 127,625 148,590 5022

IB4510 20,706 145,883 167,805 350,924 60,033 214,524 380,838 167,893 147,712 5022

Fig. 2. Contact surfaces for the analysis of the microgap formation between implant model components (IB: internal tissue;

IB: internal bone).


8

Fig. 3. Boundary and loading conditions. (a) Screw preloading, with the bone block fixed along all the axes; the red arrows

indicate the preloading to achieve a tightening torque of 32 N∙cm. (b) Applied forces: 200 N vertically (to each of the 60

nodes on three cusps and three pits) and 100 N obliquely (to each of the 30 nodes on three cusps).

The FEA was conducted with ABAQUS 6.14 software (Dassault Systèmes SIMULIA Corp.). The

convergence test was performed using four relative characteristic element sizes [36]. Table 4 shows the

maximum von Mises stress value to determine the mesh size. Based on the static FEA results (in particular, the

elastic stress), the fatigue analysis of the implant complex was performed to estimate the fatigue lives of its

titanium alloy components [35-37]. The fatigue life was calculated via repetitive mastication simulations with

alternating vertical and oblique loadings by using a multiaxial fatigue algorithm in the plasticity mode. For these

calculations, also cycling hysteresis material data of titanium were used; the elastic–plastic correction (biaxial

Neuber rule) and the cycling material data were utilized to translate the elastic stress from the FEA into elastic–

plastic stress [20,38]. The fatigue calculations were conducted with commercial software (FE-Safe 6.5, Safe

Technology Ltd.). The lowest failure lives and failure regions were predicted for 107 cycles.

Table 4. Maximum von Mises stress and stress variation of IT with 4.0 mm diameter for low density bone under oblique

loading (1.0: relative characteristic element size used in the following finite element analysis).

Relative element size

1.0 1.25 1.5 2.0

Von Mises stress Implant 320.94 322.37 313.30 325.09

(MPa) Abutment 293.27 293.66 268.87 244.76


9

Variation Implant -0.44 2.89 -3.63

(%) Abutment -0.13 9.22 9.85

Statistical analysis was performed with SPSS 20.0 software (IBM). The analysis for normal distribution

was conducted with the Shapiro–Wilk test [24,39-41]. The Kruskal–Wallis one-way and Mann–Whitney U tests

were carried out to verify the effects of implant diameter (3.5 mm, 4.0 mm, and 4.5 mm) and bone density (low,

high) [13,24,40-43]; the Bonferroni correction method was also adopted for the implant diameter. In all analyses,

the 50 highest displacement values for each model were used to evaluate the separations at the interfaces [40]. p-

Values below 0.05 were considered statistically significant [43,44].

3. RESULTS

The relative displacements, or microgaps, of the implant components were measured at the contact

surfaces between implant, abutment, and screw to predict the risks of screw loosening and infection by possible

mechanical problems [45]. Tables 5 and 6 summarize the statistical analysis results.

Table 5. Results of the statistical analysis regarding the effects of bone density and implant diameter on the maximum

microgap formation, expressed by the p-value (IT : internal tissue; IB: internal bone).

Bone density Implant diameter

Implant/Abutment Abutment/Screw Implant/Abutment Abutment/Screw

Vertical 0.642 0.040 <0.001 <0.001


IT case
Oblique 0.424 0.707 <0.001 <0.001

Vertical 0.206 <0.001 <0.001 <0.001


IB case
Oblique 0.129 0.365 <0.001 <0.001

Table 6. Effect of different implant diameters, i.e., 3.5, 4.0, and 4.5 mm, expressed by the p-value (IT : internal tissue; IB:

internal bone).

Implant/Abutment Abutment/Screw

3.5 & 4.0 4.0 & 4.5 3.5 & 4.5 3.5 & 4.0 4.0 & 4.5 3.5 & 4.5

Vertical <0.001 <0.001 <0.001 <0.001 <0.001 <0.001


IT case
Oblique <0.001 <0.001 0.320 <0.001 <0.001 <0.001
10

Vertical <0.001 0.466 <0.001 <0.001 <0.001 <0.001


IB case
Oblique <0.001 0.007 <0.001 <0.001 <0.001 <0.001

Fig. 4. Maximum microgap formation at the implant/abutment and abutment/screw interfaces (IT: internal tissue; IB: internal

bone).

Fig. 4 illustrates the maximum microgaps within the implant complex. For the IAI, the IB level

connection resulted in larger microgaps than the IT one and the oblique loading increased the microgap size

compared to the vertical loading. All models exhibited no significant differences in microgap formation

according to the bone density (p > 0.05). For the IT and IB cases, the largest microgaps were observed with the

4.0 and 3.5, mm diameter implants (p < 0.001), respectively. At the abutment/screw interface, microgap
11

formation was mainly influenced by implant diameter only in the IB case; as implant diameter increased, the

maximum microgap size increased under vertical loading and decreased under oblique loading (p < 0.001).

Fig. 5 shows the displacement patterns under oblique loading. At the IAI, they were similar for the two

connection types; larger microgaps were formed in the buccal direction in both cases. At the abutment/screw

interface, the largest microgaps were observed in the lingual direction for all the models and the microgap

formation area was the smallest for the 4.5 mm diameter implant.

Fig. 6 and Table 7 present the minimum fatigue lives for 107 repetitions of the test. Fatigue failure was

only predicted for the IB case. The minimum fatigue life of the abutment was obtained with the implant

diameter of 3.5 mm. The fatigue life of the abutment in the high density bone was 1.44 times that in the low

density one. The fatigue lives of the implants were much longer than those o f the abutments.

Fig. 7 displays the von Mises stress distribution and the areas where the fatigue failures for the

minimum fatigue lives were predicted for the IB case. The stress distribution patterns were similar in all models.

In the abutment and implant, high stress concentrations were observed in the lingual direction, which induced

fatigue fractures. For the abutment with a 3.5 mm diameter, fatigue failure occurred at the abutment/screw

interface. The maximum von Mises stress under oblique loading exceeded the yield strength for both bone

densities tested.

4. DISCUSSION

Clinicians must consider many factors when selecting dental implants. Several studies have analyzed

implant complexes through finite element modeling and loading tests, providing new insights. In the present

work, the effects of loading direction, connection type, implant diameter, and bone density on the microgap

formation and fatigue failure were investigated.


12

Fig. 5. Top views of the microgaps between the implant model components under oblique loading; yellow regions indicate

contact area for measurement of the microgaps. (IT: internal tissue; IB: internal bone; L: lingual direction; B: buccal

direction): interfaces (a) between implant and abutment for IT, (b) between implant and abutment for IB, and (c) between

abutment and screw for IB.


13

Figure 6. Fatigue lives of the implants for the internal bone connection.

Table 7. Minimum fatigue lives of the implants with different diameters, for the internal bone connection.

Abutment Implant

3.5 mm 4.0 mm 4.5 mm 3.5 mm 4.0 mm 4.5 mm

Low density 839 1,542,055 1,669,168 945,802 101,531 535,427

High density 1,221 1,842,977 1,531,793 808,444 81,433 472,716

The stress distribution patterns were similar for all finite element models. Under vertical loading, a 200

N loading was applied on the three buccal cusps and fossae (Fig. 2b). Its mean vector was directed nearly inside

the connection area in the IT case but outside it in the IB one; in the second case, a lever effect inducing

microgap formation was observed. The microgaps were 1.5–2 times higher in the IB case than in the IT one at

the IAI, and the microgap formation at the abutment/screw interface was zero in the IT model even under

oblique loading while 6–10 mm microgaps were formed in the IB one (Fig. 5). This indicates that the geometric

differences in the connection highly influence the microgap formation. A 22 mm microgap was observed on a

conical platform under a lateral force of 100 N [1]. High-resolution radiographic imaging revealed 0–18.6 and

4.8–42 mm microgaps on the conical and flat connections under 30° oblique loading, respectively [2]. FEA

results regarding microgap formation are comparable to those of previous in vitro mechanical loading studies.
14

Fig. 7. Von Mises stress distribution and fatigue analysis results for the internal bone connection case, implant diameter of

3.5 mm, and low density bone. T he yellow regions for abutment and arrows indicate the areas of predicted fractures.
15

Rarely performed in past investigations, analyses of the microgaps between screw head and abutment

were also conducted. Under a 45° oblique loading, the vector of the net lateral force was oriented from the

buccal to the lingual area in all models. Thus, the microgaps were opened on the buccal side of the IAI but

located on the lingual side of the screw/abutment interface. Despite the preloading application on the implant

complex, there was rotational movement of the abutment due to the lateral force vector; therefore, if the

preloading decreased, microgap formation within the implant complex was promoted.

As for the fatigue failure simulation, one cyclic loading was composed of one vertical and one oblique

loading to better approximate real oral conditions. In this study, one million loading cycles performed

corresponds to roughly 10 years of survival. All the IT models survived up to 107 vertical and oblique loading

cycles, which means the tissue-level implants have high resistance to fatigue failure during their whole life. The

abutment of the IB models having 4.0- and 4.5-mm implant diameters survived 500,000–1,000,000 loading

cycles before fatigue failure, corresponding to 5–10 years.

The abutment of the IB group with a 3.5-mm implant diameter failed near 1,000 loading cycles; this

may suggest that the wall thickness of the abutment connection area was not sufficiently thick enough to

withstand the cyclic loading on the molar area. The maximum stress observed was 878 MPa, which is close to

the yield strength of the titanium alloy considered for the abutment and implant in this study. Several companies

use other materials with higher mechanical properties than titanium grade 5 [46-48], and it is suggested to use

materials with better properties than titanium grade 5 in reduced diameter bone-level implants.

5. CONCLUSIONS

Microgap formation at the IAI was shown to be larger in IB than IT and also larger in oblique loading

than vertical loading. Fatigue failure on the 3.5-mm diameter implant was vulnerable. The outcomes

demonstrate that the connection type and implant diameter are critical factors in the biomechanical behavior of

implant systems. Tissue-level implants have advantages of less microgap formation and longer fatigue life

compared to bone-level implants. The use of 3.5-mm diameter implants should be carefully considered in the

posterior area.

ACKNOWLEDGMENTS

This work was partly supported by National Research Foundation of Korea (NRF) funded by the

Ministry of Education of Korea (No.2018R1D1A1B07049789, the Basic Science Research Program).


16

Declaration of interests

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] A. Rack, T. Rack, M. Stiller, H. Riesemeier, S. Zabler, K. Nelson, In vitro synchrotron -based radiography of

micro-gap formation at the implant–abutment interface of two-piece dental implants, J. Synchrotron Radiat. 17

(2010) 289-294.

[2] H. Zipprich, P. Weigl, C. Ratka, B. Lange, H.C. Lauer, The micromechanical behavior of implant‐abutment

connections under a dynamic load protocol, Clin. Implant. Dent. Relat. Res. 20 (2018) 814-823.

[3] E.C. Lopes de Chaves e Mello Dias, M. Sperandio, M. Henrique Napimoga, Association Between Implant-

Abutment Microgap and Implant Circularity to Bacterial Leakage: An In Vitro Study Using Tapered Connection

Implants, Int. J. Oral. Maxillofac. Implants. 33 (2018) 505-511.

[4] D. Koletsi, A. Iliadi, T. Eliades, G. Eliades, In vitro simulation and in vivo assessment of tooth wear: a meta-

analysis of in vitro and clinical research, Materials . 12 (2019) 3575.

[5] A. Berberi, D. Maroun, W. Kanj, E.Z. Amine, A. Philippe, Micromovement Evaluation of Original and

Compatible Abutments at the Implant-abutment Interface, J. Contemp. Dent. Pract. 17 (2016) 907-913.

[6] L.A. Lang, B. Kang, R.-F. Wang, B.R. Lang, Finite element analysis to determine implant preload, J.

Prosthet. Dent. 90 (2003) 539-546.

[7] A. Schwitalla, M. Abou-Emara, T. Spintig, J. Lackmann, W. Müller, Finite element analysis of the

biomechanical effects of PEEK dental implants on the peri-implant bone, J. Biomech. 48 (2015) 1-7.

[8] D. Jörn, P. Kohorst, S. Besdo, M. Rücker, M. Stiesch, L. Borchers, Influence of lubricant on screw preload

and stresses in a finite element model for a dental implant, J. Prosthet. Dent. 112 (2014) 340-348.

[9] H. Zipprich, F. Rathe, S. Pinz, L. Schlotmann, H.-C. Lauer, C. Ratka, Effects of Screw Configuration on the
17

Preload Force of Implant-Abutment Screws, Int. J. Oral. Maxillofac. Implants. 33 (2018) e25-e32.

[10] M.M. Al-Sahan, N.S. Al Maflehi, R.F. Akeel, The influence of tightening sequence and method on screw

preload in implant superstructures, Int. J. Prosthodont. 27 (2014) 76-79.

[11] S. Saidin, M.R.A. Kadir, E. Sulaiman, N.H.A. Kasim, Effects of different implant–abutment connections on

micromotion and stress distribution: Prediction of microgap formation, J. Dent. 40 (2012) 467-474.

[12] Z. Li, S. Gao, H. Chen, R. Ma, T. Wu, H. Yu, Micromotion of implant-abutment interfaces (IAI) after

loading: correlation of finite element analysis with in vitro performances, Med. Biol. Eng. Comput. 57 (2019)

1133-1144.

[13] H. Lee, S. Park, K.-R. Kwon, G. Noh, Effects of cementless fixation of implant prosthesis: A finite element

study, J. Adv. Prosthodont. 11 (2019) 341-349.

[14] K. Shemtov-Yona, D. Rittel, Fatigue failure of dental implants in simulated intraoral media, J. Mech. Behav.

Biomed. Mater. 62 (2016) 636-644.

[15] K. Shemtov‐Yona, D. Rittel, E.E. Machtei, L. Levin, E ffect of D ental I mplant D iameter on F atigue P

erformance. P art II: F ailure A nalysis, Clin. Implant. Dent. Relat. Res. 16 (2014) 178-184.

[16] S.-Y. Song, J.-Y. Lee, S.-W. Shin, Effect of implant diameter on fatigue strength, Implant. Dent. 26 (2017)

59-65.

[17] S. Yamaguchi, Y. Yamanishi, L.S. Machado, S. Matsumoto, N. Tovar, P.G. Coelho, V.P. Thompson, S.

Imazato, In vitro fatigue tests and in silico finite element analysis of dental implants with different

fixture/abutment joint types using computer-aided design models, J. Prosthodont. Res. 62 (2018) 24-30.

[18] R. Coray, M. Zeltner, M. Özcan, Fracture strength of implant abutments after fatigue testing: A systematic

review and a meta-analysis, J. Mech. Behav. Biomed. Mater. 62 (2016) 333-346.

[19] J.K. Foong, R.B. Judge, J.E. Palamara, M.V. Swain, Fracture resistance of titanium and zirconia abutments:

an in vitro study, J. Prosthet. Dent. 109 (2013) 304-312.

[20] M. Pérez, Life prediction of different commercial dental implants as influence by uncertainties in their

fatigue material properties and loading conditions, Comput. Methods. Programs. Biomed. 108 (2012) 1277-
18

1286.

[21] H. Lee, S. Park, G. Noh, Biomechanical analysis of 4 types of short dental implants in a resorbed mandible,

J. Prosthet. Dent. 121 (2019) 659-670.

[22] H.A. Bulaqi, M.M. Mashhadi, H. Safari, M.M. Samandari, F. Geramipanah, Effect of increased crown

height on stress distribution in short dental implant components and their surrounding bone: A finite element

analysis, J. Prosthet. Dent. 113 (2015) 548-557.

[23] C.A.A. Lemos, M.L. Ferro-Alves, R. Okamoto, M.R. Mendonça, E.P. Pellizzer, Short dental implants

versus standard dental implants placed in the posterior jaws: A systematic review and meta-analysis, J. Dent. 47

(2016) 8-17.

[24] L.B. Torcato, E.P. Pellizzer, F.R. Verri, R.M. Falcón-Antenucci, J.F.S. Júnior, D.A. de Faria Almeida,

Influence of parafunctional loading and prosthetic connection on stress distribution: a 3D finite element analysis,

J. Prosthet. Dent. 114 (2015) 644-651.

[25] L. Barbier, J.V. Sloten, G. Krzesinski, E.S. Van Der Perre, Finite element analysis of non‐axial versus axial

loading of oral implants in the mandible of the dog, J. Oral. Rehabil. 25 (1998) 847-858.

[26] C. Rungsiyakull, J. Chen, P. Rungsiyakull, W. Li, M. Swain, Q. Li, Bone’s responses to different designs of

implant-supported fixed partial dentures, Biomech. Model. Mechanobiol. 14 (2015) 403-411.

[27] K. Tolidis, D. Papadogiannis, Y. Papadogiannis, P. Gerasimou, Dynamic and static mechanical analysis of

resin luting cements, J. Mech. Behav. Biomed. Mater. 6 (2012) 1-8.

[28] H. Vaillancourt, R. Pilliar, D. McCammond, Finite element analysis of crestal bone loss around porous‐

coated dental implants, J. Appl. Biomater. 6 (1995) 267-282.

[29] T. Sugiura, K. Yamamoto, M. Kawakami, S. Horita, K. Murakami, T. Kirita, Influence of bone parameters

on peri-implant bone strain distribution in the posterior mandible, Med. Oral. Patol. Oral. Cir. Bucal. 20 (2015)

e66-e73.

[30] M. Niinomi, Mechanical properties of biomedical titanium alloys, Mater. Sci. Eng. A. 243 (1998) 231-236.

[31] T. Guda, T.A. Ross, L.A. Lang, H.R. Millwater, Probabilistic analysis of preload in the abutment screw of a
19

dental implant complex, J. Prosthet. Dent. 100 (2008) 183-193.

[32] R.-F. Wang, B. Kang, L.A. Lang, M.E. Razzoog, The dynamic natures of implant loading, J. Prosthet. Dent.

101 (2009) 359-371.

[33] J. Bickford, An introduction to the design and behavior of bolted joints, Revised and expanded, CRC press,

New York, 1995.

[34] J. Chen, Z. Zhang, X. Chen, X. Zhang, Influence of custom-made implant designs on the biomechanical

performance for the case of immediate post-extraction placement in the maxillary esthetic zone: a finite element

analysis, Comput. Methods. Biomech. Biomed. Engin. 20 (2017) 636-644.

[35] M. Pirmoradian, H.A. Naeeni, M. Firouzbakht, D. Toghraie, R. Darabi, Finite element analysis and

experimental evaluation on stress distribution and sensitivity of dental implants to assess optimum length and

thread pitch, Comput. Methods. Programs. Biomed. 187 (2020) 105258.

[36] H. Lee, S.-M. Park, K. Noh, S.-J. Ahn, S. Shin, G. Noh, Biomechanical stability of internal bone-level

implant: Dependency on hex or non-hex structure, Struct. Eng. Mech. 74 (2020) 567-576.

[37] A. Eram, M. Zuber, L.G. Keni, S. Kalburgi, R. Naik, S. Bhandary, S. Amin, I.A. Badruddin, Finite element

analysis of immature teeth filled with MTA, Biodentine and Bioaggregate, Comput. Methods. Programs.

Biomed. 190 (2020) 105356.

[38] E.P. Holmgren, R.J. Seckinger, L.M. Kilgren, F. Mante, Evaluating Parameters of osseointegrated dental

implants using finite element analysis a two-dimensional comparative study examining the effects of implant

diameter, implant shape, and load direction, J. Oral. Implantol. 24 (1998) 80-88.

[39] J.F.S. Junior, F.R. Verri, D.A. de Faria Almeida, V.E. de Souza Batista, C.A.A. Lemos, E.P. Pellizzer, Finite

element analysis on influence of implant surface treatments, connection and bone types, Mater. Sci. Eng. C. 63

(2016) 292-300.

[40] L. Minatel, F.R. Verri, G.A.H. Kudo, D.A. de Faria Almeida, V.E. de Souza Batista, C.A.A. Lemos, E.P.

Pellizzer, J.F.S. Junior, Effect of different types of prosthetic platforms on stress -distribution in dental implant-

supported prostheses, Mater. Sci. Eng. C. 71 (2017) 35-42.


20

[41] F.R. Verri, J.F.S. Junior, D.A. de Faria Almeida, G.B.B. de Oliveira, V.E. de Souza Batista, H.M. Honório,

P.Y. Noritomi, E.P. Pellizzer, Biomechanical influence of crown-to-implant ratio on stress distribution over

internal hexagon short implant: 3-D finite element analysis with statistical test, J. Biomech. 48 (2015) 138-145.

[42] N. Lümkemann, M. Eichberger, B. Stawarczyk, Bond strength between a high -performance thermoplastic

and a veneering resin, J. Prosthet. Dent. (2020). https://doi.org/10.1016/j.prosdent.2019.10.017.

[43] F.R. Verri, R.S. Cruz, V.E. de Souza Batista, D.A.d.F. Almeida, A.C.G. Verri, C.A.d.A. Lemos, J.F. Santiago

Júnior, E.P. Pellizzer, Can the modeling for simplification of a dental implant surface affect the accuracy of 3D

finite element analysis?, Comput. Methods. Biomech. Biomed. Engin. 19 (2016) 1665-1672.

[44] D.A. de Faria Almeida, E.P. Pellizzer, F.R. Verri, J.F. Santiago Jr, P.S.P. de Carvalho, Influence of tapered

and external hexagon connections on bone stresses around tilted dental implants: three‐dimensional finite

element method with statistical analysis, J. Periodontol. 85 (2014) 261-269.

[45] G.C. Silva, T.M. Cornacchia, C.S. de Magalhães, A.C. Bueno, A.N. Moreira, Biomechanical evaluation of

screw-and cement-retained implant-supported prostheses: a nonlinear finite element analysis, J. Prosthet. Dent.

112 (2014) 1479-1488.

[46] K.-S. Shih, C.-C. Hsu, T.-P. Hsu, S.-M. Hou, C.-K. Liaw, Biomechanical analyses of static and dynamic

fixation techniques of retrograde interlocking femoral nailing using nonlinear finite element methods, Comput.

Methods. Programs. Biomed. 113 (2014) 456-464.

[47] V.A.R. Barão, W.G. Assunção, L.F. Tabata, E.A.C. de Sousa, E.P. Rocha, Effect of different mucosa

thickness and resiliency on stress distribution of implant-retained overdentures-2D FEA, Comput. Methods.

Programs. Biomed. 92 (2008) 213-223.

[48] J. Park, S.-J. Ahn, H. Lee, G. Noh, Implant placement in the implant-supported mandibular advancement

device for completely edentulous patients: A finite element study, Journal of Computational Design and

Engineering, (2020) qwaa067.

You might also like