You are on page 1of 41

Journal Pre-proof

Fracture resistance of CAD/CAM occlusal veneers: A systematic review of laboratory


study

Eman Albelasy, Hamdi H. Hamama, James K.H. Tsoi, Salah H. Mahmoud

PII: S1751-6161(20)30501-4
DOI: https://doi.org/10.1016/j.jmbbm.2020.103948
Reference: JMBBM 103948

To appear in: Journal of the Mechanical Behavior of Biomedical Materials

Received Date: 4 May 2020


Revised Date: 18 June 2020
Accepted Date: 21 June 2020

Please cite this article as: Albelasy, E., Hamama, H.H., Tsoi, J.K.H., Mahmoud, S.H., Fracture
resistance of CAD/CAM occlusal veneers: A systematic review of laboratory study, Journal of the
Mechanical Behavior of Biomedical Materials (2020), doi: https://doi.org/10.1016/j.jmbbm.2020.103948.

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 Ltd.


Fracture Resistance of CAD/CAM Occlusal Veneers: A Systematic Review of

Laboratory Study

Authors: Eman Albelasya, Hamdi H Hamamaa,b*, James K.H. Tsoic*, and Salah H.
Mahmouda.
a
Operative Dentistry Department, Faculty of Dentistry, Mansoura University, Egypt.
b
Division of Restorative Dental Sciences, Faculty of Dentistry, The University of Hong
Kong, Hong Kong
c
Dental Material Science, Division of Applied Oral Sciences and Community Dental Care,
Faculty of Dentistry, The University of Hong Kong, Hong Kong

Corresponding authors:

James K.H. Tsoi


Dental Material Science, Division of Applied Oral Sciences and Community Dental Care,
Faculty of Dentistry, The University of Hong Kong
Mailing address: Prince Philip Dental Hospital, 34 Hospital Road, Sai Ying, Pun,
Hong Kong
Tel: +852-28590515
Fax: +852-25489464
Email: jkhtsoi@hku.hk

Hamdi H Hamama
Operative Dentistry Department, Faculty of Dentistry, Mansoura University, Egypt.
Mailing address: Operative Dentistry Dept, Faculty of Dentistry, Algomhoria Street,
Mansoura, Dakahlia, Egypt Po (box) 35516
Tel: +201153418154
Email: hamdy@connect.hku.hk

1
Abstract

Objective: The purpose of this systematic review was to summarize scientific evidence that

evaluates in vitro fracture and fatigue strength of occlusal veneers in different thicknesses,

CAD/CAM materials, and under different aging methodologies.

Materials and methods: An electronic search of 3 English databases (The National Library

of Medicine (MEDLINE/PubMed), ScienceDirect, and EBSCOhost) was conducted.

Laboratory studies published between September 2009 and October 2019 that evaluated

fracture or fatigue strength of CAD/CAM occlusal veneers and used human teeth were

selected. The included studies were individually evaluated for the risk of bias following a

predetermined criterion. The outcomes assessed including the types of the restorative

material, the thickness of the veneers, and aging methods.

Results: A total of 12 studies fulfilled the inclusion criteria. Most of the included studies

(86%) evaluated the fracture strength of occlusal veneers. Two studies evaluated fatigue

resistance. There was a significant relationship between the choice of materials and fracture

strength. Polymeric materials performed better in fatigue testing in comparison to ceramics.

Lithium silicate-based glass ceramics showed more favorable outcomes in a thickness of 0.7-

1.0 mm. Fracture resistance values in all the included studies exceeded maximum bite forces

in the posterior region.

Conclusions: The outcomes of this systematic review suggest that occlusal veneers can

withstand bite forces in the posterior region. whereas the measurement of thickness should be

standardized in order to have a fair comparison. Further research needs to be conducted to

evaluate the longevity of this type of restorations clinically.

Keywords: occlusal veneers, CAD/CAM, ceramics, dental materials, overlay, onlay

2
1. Introduction

Obtaining excellent aesthetic results with preservation of biological structures is a

prime objective of contemporary restorative dentistry (Angerame and De Biasi, 2019). With

the emergence of higher strength and tough restorative materials along with CAD/CAM

technology and improved adhesive techniques, conservative approaches can be considered

(Magne et al., 2010). Accordingly, restorative protocols became less complicated and capable

of providing satisfactory results (Angerame and De Biasi, 2019). Clinicians were able to shift

towards conservative approaches due to remarkable improvements in dental adhesive systems

regarding mechanical and optical properties (Bindl et al., 2006).

The term occlusal veneer was proposed in an attempt to adopt a minimally invasive

approach in cases of generalized wear in the posterior region (Magne et al., 2010). Occlusal

veneers are a non-invasive alternative for restoring worn posterior teeth requiring little

additional tooth structure removal (Johnson et al., 2014). Traditional approaches for

managing worn posterior teeth may involve invasive full mouth rehabilitation with biological

consequences including sacrificing sound tooth structure (Vailati and Belser, 2008a). In

patients where a substantial amount of tooth structure has been compromised by tooth wear,

extensive preparation may consider to be unacceptable (Schlichting et al., 2011). This said,

the choice of material is an important factor for a successful occlusal veneer.

New glass ceramics, such as lithium disilicate, are stronger than traditional feldspathic

porcelain(Leung et al., 2015), HF-etchable (Tian et al., 2014; Wong et al., 2017) and

machinable (Beuer et al., 2008; Magne et al., 2013). They have extended the indications of

bonded ceramics to include minimally invasive restorations. On the other hand, polymer-

infiltrated ceramics were shown to have good mechanical properties after luting at a reduced

thickness (Dirxen et al., 2013) due to its dual network structure, such that the interlinked

3
polymer network can mitigate crack propagation (Coldea et al., 2013b; Homaei et al., 2016b).

Polymer-infiltrated ceramics have demonstrated promising results as minimally invasive

restorations with a 3-year survival rate of 97.4% for inlays (Spitznagel et al., 2018).

In addition, the performance of machinable resin composite has improved

significantly in the last decade (Mainjot et al., 2016). CAD/CAM fabricated resin composite

overlays have demonstrated fatigue resistance values superior to that of porcelain (Magne and

Knezevic, 2009a, b). However, these polymer-based materials possess their own limitations,

e.g. wear, discoloration, and low fracture strength (Egilmez et al., 2018; Tekce et al., 2016).

Variations in mechanical properties of ceramic and resin-based materials raise the question

regarding which material is capable of survival in the load-bearing posterior region

(Morimoto et al., 2016). Long-term clinical data regarding survival rates, discoloration, and

degradation of CAD/CAM resin composite are still lacking and in need of further research.

Laboratory investigation of a restorative material fracture resistance prior to clinical

use is essential to decide if it can be regarded as a reliable treatment option. However, the

nature of force in the oral cavity is not strictly compressive as in fracture testing (Pjetursson

et al., 2007). In bonded ceramics, several factors affect the mechanical behavior of both

restoration/tooth complex including restoration thickness, quality of the adhesive interface,

the ratio of modulus of elasticity between restoration, cement, and dentin (Homaei et al.,

2016a). Thus, the strength of the tooth, restoration, adhesive system, and cement contributes

to the performance of indirect restorations (El-Damanhoury et al., 2015; Homaei et al., 2018;

Maghami et al., 2018). On the other hand, scientific evidence on the minimum acceptable

thickness for glass ceramic occlusal veneers is still scarce. An earlier study (Guess et al.,

2013) has concluded that decreasing the preparation depth to 0.5 and 1.0 mm had no

significant impact on fracture resistance of pressable lithium disilicate glass ceramic onlays

4
on premolars. Conversely, another study (Sasse et al., 2015) has shown the fracture strength

of ceramic occlusal veneers was significantly influenced by changes in thickness. When

polymer-infiltrated ceramics and CAD/CAM resin composite were compared, it was found

that when the thickness is increased to more than 0.5 mm, fracture resistance values that

exceed average posterior bite forces can be predicted (Chen et al., 2014). Moreover, all-

ceramic occlusal veneers were able to survive in simulated masticatory function when

subjected to artificial aging (Skouridou et al., 2013), while machinable resin composite

materials under similar comparisons regarding the strength of the restoration have shown

significantly better results. In the same context, all-ceramic crowns compared to their resin

composite counterparts showed inferior results regarding survival rates when subjected to

cyclic loading (Kassem et al., 2012).

As a result, previous studies that evaluated the mechanical behavior of occlusal

veneers seem to be conflicting. The selection of restorative material and the minimum

thickness that can bear occlusal forces in the posterior region is still under investigation.

Therefore, this systematic review was conducted to analyze the published data that evaluate

the fracture and fatigue resistance of occlusal veneers manufactured from different materials

and at variable thicknesses.

2. Material and methods

2.1 PIO

The PIO of this systematic review was defined as follows:

Population: occlusal veneers; Intervention: materials; thickness; thermomechanical aging;

Outcomes: fracture strength, fatigue strength;

2.2 Search strategy

5
The protocol of this systematic review was designed following the Preferred Reporting

Items Systematic Review and Meta-Analysis (PRISMA) guidelines.

2.3 Information source and search

A detailed electronic search of 3 English databases (The National Library of Medicine

(MEDLINE/PubMed), Science Direct, and EBSCOhost) was conducted by one of the

reviewers (EH). Furthermore, a subsequent manual search was conducted to check for non-

online resources. Only studies that were published on or after 2009 were selected. The

following keywords and MESH-terms were used ("occlusal veneers" OR “veneers” OR

"hybrid ceramics" OR "CAD/CAM composite" OR "CAD/CAM resin" OR "non-retentive

overlay" OR "ceramic onlay" OR "CAD/CAM" ).

2.4 Study selection

The selection of articles went through 3 stages, (1) selection in accordance with the

relevance of the title, (2) selection in accordance with the relevance of the abstract, and (3)

analysis of the full text. All articles found by the electronic and manual searches were

collected, and a copy was given to each author. The eligibility criteria for all included studies

were checked individually by each author. The agreement of 2 authors at least was required

for the study to be selected.

2.5 Eligibility criteria

2.5.1 Inclusion criteria

The studies included in this systematic review were all laboratory studies, written in

English and published between September 2009 and October 2019. For all the included

studies, fracture or/and fatigue resistance of occlusal veneers were evaluated. The research

6
question was as follows: is fracture resistance of minimally invasive occlusal veneers

influenced by variation in thickness, material selection, and thermomechanical aging?

2.5.2 Exclusion criteria

The following studies were excluded during the assessment process: studies that were

published before September 2009 and after October 2019, non-English manuscripts, review

articles, case reports, and clinical studies. Moreover, laboratory studies that used

endodontically treated teeth were excluded. In addition, all laboratory studies conducted to

evaluate implant-supported restorations were also excluded. Furthermore, studies that did not

utilize natural human teeth were excluded. All laboratory studies which evaluated full-

coverage crowns, onlays, and inlays were also excluded. Pilot studies and Studies that used

testing methodologies other than (fracture/fatigue) strength were also excluded. The

objectives and conclusions of each included study are summarized in table 1.

7
Table 1 Summary of studies included in the systematic review

Study Sample Year Objectives Conclusions


size
Al-Akhali et al., 64 2019 To evaluate the influence of Thermomechanical aging generally
2019 premolars thermomechanical aging on the durability and decreased the fracture resistance of all tested
fracture resistance behavior of occlusal CAD/CAM materials when bonded to
veneers fabricated from different biomedical enamel using a self-etching technique.
dental CAD/CAM materials.
Angerame and 16 molars 2019 To evaluate the fracture resistance and The study presented satisfactory results for
De Biasi, 2019 marginal quality of maxillary molars restored the two preparation designs for occlusal
with lithium disilicate glass ceramic occlusal veneers.
veneers with two preparation designs.
Ioannidis et al., 100 molars 2019 To test whether the load-bearing capacity of Regarding their maximum load-bearing
2019 occlusal veneers bonded to enamel and made capacities, minimally invasive occlusal
of ceramic or hybrid materials differs from veneers made of ceramic or hybrid materials
those of porcelain-fused-to-metal or lithium can be applied to correct occlusal tooth wear
disilicate glass ceramic crowns. and thus replace conventional crown
restorations.
Maeder et al., 100 molars 2019 To test whether the load-bearing capacity of Regarding their maximum load-bearing
2019 occlusal veneers made of ceramic or hybrid capacity, minimally invasive occlusal
materials bonded to dentin differs from those veneers made of ceramic, hybrid materials,
of porcelain-fused-to-metal or lithium or polymeric materials can be applied to
disilicate glass ceramic crowns. correct occlusal tooth wear with exposed
dentin and thus replace conventional crown
restorations in cases of normally expected
intraoral bite forces.
Andrade et al., 70 molars 2018 To evaluate in vitro, the influence of Ultrathin occlusal veneers appear to be a
2018 CAD/CAM restorative materials and their promising procedure for the restoration of
thickness on the fracture resistance of teeth eroded posterior teeth.
restored with occlusal veneers.
Yazigi et al., 48 2018 To evaluate the effects of artificial aging on 1. Optical coherence tomography allows an
2018 premolars thin glass ceramic occlusal premolar veneers easy and non-invasive method to
adhesively bonded to dentin, by examining the internally scan teeth and restorations.
changes caused by artificial aging using
spectral-domain optical coherence 2. The development of cracks in the ceramic
tomography. In addition, the development of did not affect the fracture strength of the
cracks in the veneer and their possible restorations but might lead to a more
influence on the behavior of ceramic catastrophic type of failure.
restorations were examined.
Al-Akhali et al., 64 2017 The purpose of this in vitro study was to All tested CAD/CAM materials exhibited a
2017 premolars evaluate the influence of thermomechanical fracture resistance considerably exceeding
aging on the durability and fracture resistance the average occlusal force in the posterior
behavior of occlusal veneers fabricated from dentition.
different CAD/CAM materials.
Yazigi et al., 96 2017 To evaluate the efficiency of immediate dentin Immediate dentin sealing protocol is
2017 premolars sealing and the effects of different bonding recommended whenever dentin is exposed
protocols on the fracture strength of during the preparation for thin glass-ceramic
CAD/CAM occlusal veneers bonded to occlusal veneers.
exposed dentin.
Sasse et al., 2015 72 molars 2015 To evaluate the influence of ceramic thickness The results suggest using a thickness of 0.7–
and type of dental bonding surface on the 1.0 mm for non-retentive full-coverage
fracture resistance of non-retentive full- adhesively retained occlusal lithium
coverage adhesively retained occlusal veneers disilicate ceramic restorations.
made from lithium disilicate ceramic.
Johnson et al., 60 molars 2014 To determine the effect of material selection Occlusal veneers fabricated from composite-
2014 and restoration dimension on the fracture ceramic hybrid materials are more likely to
resistance of occlusal veneers. survive heavier loads.

Schlichting et al., 40 molars 2011 To assess the influence of CAD/CAM Both composite resins (MZ100 and XR)
2011 restorative material (ceramic vs. resin increased the fatigue resistance of ultra-thin
composite) on the fatigue resistance of ultra- occlusal veneers when compared to the
thin occlusal veneers. ceramics (Empress CAD and e.max CAD).
Magne et al., 30 molars 2010 The purpose of this study was to assess and Posterior occlusal veneers made of
2010 compare the fatigue resistance of resin composite resin (Paradigm MZ100) had
composite and ceramic posterior occlusal significantly higher fatigue resistance
veneers. compared to IPS Empress CAD and IPS
e.max CAD.

8
2.6 Risk of bias evaluation

Two authors (EH and HH) independently assessed the risk of bias for each of the selected

studies according to the criteria applied by Rosa et al. and Sarkis-Onofre et al. (Rosa et al.,

2015; Sarkis-Onofre et al., 2014). The following parameters were evaluated: teeth

randomization, use of sound teeth (free from caries or restorations), use of materials

following the instructions of manufacturers, use of teeth with similar dimensions, teeth

preparation performed by the same operator, evaluation of failure mode, and description of

sample-size calculation. If the author reported the parameter, the article received a yes (Y) on

that specific parameter; if it was not possible to find the information, the article received a no

(N). The articles which reported 1 to 3 items were classified as having a high risk of bias, 4 or

5 items as medium risk of bias, and 6 or 7 items as low risk of bias.

3. Results

The online search in MEDLINE/PubMed database resulted in 125 articles being

identified. This was followed by a subsequent search of 2 more databases along with a hand

search. Human molars were utilized in 8 studies (Andrade et al., 2018; Angerame and De

Biasi, 2019; Ioannidis et al., 2019; Johnson et al., 2014; Maeder et al., 2019; Magne et al.,

2010; Sasse et al., 2015; Schlichting et al., 2011) while 4 studies used human premolars (Al-

Akhali et al., 2017; Al-Akhali et al., 2019; Yazigi et al., 2018; Yazigi et al., 2017). Nineteen

studies were excluded because they were published before 2009 and 2 manuscripts were

excluded because they were not written in English. Seven review articles were excluded.

Furthermore, 13 case reports were excluded. Of the remaining 79 studies, 5 clinical studies, 7

studies that utilized endodontically treated teeth in addition to 40 laboratory studies that did

not utilize natural human teeth were excluded. Ten studies that involved full-coverage

crowns, onlays, and inlays were also excluded. One pilot study was excluded. Moreover, 5

9
laboratory studies that used other testing methodologies (e.g. microtensile bond strength)

were also excluded. Twelve studies fulfilled the inclusion criteria for this systematic review.

The detailed selection process is illustrated in the flow chart (Figure 1). This systematic

review included 12 laboratory studies that were conducted to evaluate the impact of

(restoration thickness, material selection, and thermomechanical aging) on fracture (and/or)

fatigue resistance of occlusal veneers.

10
Figure 1. Flow chart of the study selection process, n= the number of articles

11
3.1 Risk of bias

Table 2 listed the summary of the risk of bias assessment. Three studies demonstrated

a low risk of bias (Al-Akhali et al., 2017; Angerame and De Biasi, 2019; Johnson et al.,

2014). While 6 studies showed a high risk of bias (Al-Akhali et al., 2019; Ioannidis et al.,

2019; Maeder et al., 2019; Magne et al., 2010; Sasse et al., 2015; Schlichting et al., 2011).

The remaining 3 showed a medium risk of bias (Andrade et al., 2018; Yazigi et al., 2018;

Yazigi et al., 2017). All studies except (Al-Akhali et al., 2017; Angerame and De Biasi,

2019) scored poorly on the item regarding the description of sample size calculation. All

studies scored high on the two items regarding the use of sound teeth and the use of materials

following manufacturer instructions. Two studies (Al-Akhali et al., 2017; Johnson et al.,

2014) scored high on the item regarding teeth preparation performed by the same operator.

12
Table 2 Risk of bias assessment summary

Parameter

Evaluation of
to
Use of teeth
from

Use of teeth
or

of

Description of
performed by
similar

same

size
Randomization

manufacture

failure mode
restorations

preparation
instructions

Risk of bias
dimensions

calculation
according
materials

operator
of teeth

sample
caries

Teeth
with
Study

free

Use

the
Al-Akhali et al. High
N Y Y N N Y N
(2019)
Angerame et al. Low
Y Y Y Y N Y Y
(2019)
Ioannidis et al. High
N Y Y N N Y N
(2019)
Maeder et al. High
N Y Y N N Y N
(2019)
Andrade et al. Medium
Y Y Y Y N Y N
(2018)
Yazigi et al. Medium
Y Y Y N N Y N
(2018)
Al-Akhali et al. Low
N Y Y Y Y Y Y
(2017)
Yazigi et al. Medium
Y Y Y N Y N N
(2017)
Sasse et al. High
N Y Y N N Y N
(2015)
Johnson et al. Low
Y Y Y Y Y Y N
(2014)
Schlichting et High
N Y Y N N Y N
al.(2011)
Magne et al. High
N Y Y N N Y N
(2010)
Abbreviations: N, no; Y, yes

13
3.2 Evaluation of thickness

Five studies (42%) evaluated the impact of restoration thickness on fracture resistance

of occlusal veneers (Andrade et al., 2018; Ioannidis et al., 2019; Johnson et al., 2014; Maeder

et al., 2019; Sasse et al., 2015). Two of them used occlusal veneers at a thickness of 0.5 and

1.0 mm (Ioannidis et al., 2019; Maeder et al., 2019). One study used a restoration thickness of

0.6 and 1.5 mm (Andrade et al., 2018). Of the remaining 2 studies, 1 study (Sasse et al.,

2015) used 3 different thicknesses (0.3 mm, 0.7 mm and 1.0 mm). The remaining study

(Johnson et al., 2014) evaluated occlusal veneers at a thickness of (0.3, 0.6 and 1.0 mm)

(Table 3). Mean fracture resistance values for different materials with variable thicknesses

are presented in Table 4. However, given that thickness measurement is not a standardized

procedure, quantitative analyses on these numbers are not feasible and thus meta-analysis

could not be performed (Habib et al., 2018).

Table 3 Different restoration thicknesses used in the included studies

Study Restoration Results


thickness / mm
Ioannidis et al. 0.5 and 1.0 Load bearing capacities of 0.5 and 1.0mm occlusal veneers
(2019) were similar to the control group (conventional PFM crowns)
Maeder et al. 0.5 and 1.0 Load bearing capacities of 1.0mm and 0.5 occlusal veneers and
(2019) the control group (conventional PFM crowns) were similar
This applies for all materials except for LDC
Andrade et al. 0.6 and 1.5 For LDC, fracture resistance was significantly higher at 1.5mm
(2018) thickness. For hybrid ceramics and composite, no significance
difference between different thicknesses was found.
Sasse et al. (2015) 0.3, .0.7 and 1.0 Thickness of occlusal ceramic veneers had a statistically
significant influence on fracture resistance.
Johnson et al. 0.3, 0.6, 1.0 No significant difference was found in fracture strength
(2014) between occlusal veneers thicknesses.
Abbreviations: LDC: lithium disilicate glass ceramics; PFM: porcelain fused to metal

14
Table 4 Mean fracture resistance values in Newton (N) for different materials at

different thicknesses

Study Material Thickness / mm Mean fracture resistance values in Newton / N


Ioannidis et RNC 0.5 1941 ± 631
al. (2019) 1.0 2274 ± 455
PIC 0.5 1952 ± 730
1.0 1839 ± 779
ZIR 0.5 Could not be fabricated
1.0 2256 ± 265
LDC (press) 0.5 1178 ± 588
1.0 1530 ± 440
Maeder et al. RNC 0.5 2092 ± 439

(2019) 1.0 2328 ± 288


PIC 0.5 1981 ± 617
1.0 2239 ± 493
ZIR 0.5 2382 ± 228
1.0 2483 ± 23
LDC (press) 0.5 1191 ± 382
1.0 1851 ± 631
Andrade et al. RNC 0.6 3384 ± 922

(2018) 1.5 3584 ± 954


PIC 0.6 2973 ± 635
1.5 3540 ± 986
LDC (milling) 0.6 3067 ± 933
1.5 4995 ± 855
Johnson et al. RNC 0.3 2078 ± 605
(2014) 0.6 2141 ± 473
1.0 2115 ± 462
RC 0.3 1620 ± 433
0.6 1830 ± 501
1.0 2027 ± 704

Abbreviations: RNC: resin nano-ceramic; PIC: polymer- infiltrated ceramic; LDC: lithium disilicate

glass ceramic; ZIR: zirconia; RC: resin composite

15
3.3 Evaluation of material selection

Eight studies (66.6%) evaluated the impact of materials selection on fracture and

fatigue resistance of occlusal veneers (Al-Akhali et al., 2017; Al-Akhali et al., 2019; Andrade

et al., 2018; Ioannidis et al., 2019; Johnson et al., 2014; Maeder et al., 2019; Magne et al.,

2010; Schlichting et al., 2011). In all of them, the impact of material type on fracture

resistance was significant. Eleven studies (91.6%) used lithium disilicate glass ceramic.

CAD/CAM fabricated blocks (e.max CAD, Ivoclar Vivadent, Schaan, Liechtenstein) were

used in 8 studies (66.6%) (Al-Akhali et al., 2017; Al-Akhali et al., 2019; Andrade et al.,

2018; Magne et al., 2010; Sasse et al., 2015; Schlichting et al., 2011; Yazigi et al., 2018;

Yazigi et al., 2017) whereas 2 studies used pressed lithium disilicate (Ioannidis et al., 2019;

Maeder et al., 2019). The use of zirconia-added lithium disilicate (Vita Suprinity, Vita

Zahnfabrik, Bad Säckingen, Germany) was reported in 2 studies (Al-Akhali et al., 2017; Al-

Akhali et al., 2019) (16.6%). Two studies (Ioannidis et al., 2019; Maeder et al., 2019)

(16.6%) used yttrium stabilized zirconia.

Polymer-infiltrated ceramics (VITA ENAMIC, Vita Zahnfabrik, Bad Säckingen,

Germany) were used in 5 studies (41.6%) (Al-Akhali et al., 2017; Al-Akhali et al., 2019;

Andrade et al., 2018; Ioannidis et al., 2019; Maeder et al., 2019). CAD/CAM resin composite

restorations were used in 50% of the studies (Andrade et al., 2018; Ioannidis et al., 2019;

Johnson et al., 2014; Maeder et al., 2019; Magne et al., 2010; Schlichting et al., 2011). Four

studies (Andrade et al., 2018; Ioannidis et al., 2019; Johnson et al., 2014; Maeder et al., 2019)

have used resin nanoceramic (Lava Ultimate 3M ESPE, St. Paul. Minn, USA) and three

studies (Johnson et al., 2014; Magne et al., 2010; Schlichting et al., 2011) have used

CAD/CAM resin composite (Paradigm MZ100 blocks; 3M ESPE, St. Paul, Minn, USA).

Only 2 studies (Ioannidis et al., 2019; Maeder et al., 2019) used porcelain fused to metal

16
crowns. Furthermore, 2 studies (Al-Akhali et al., 2017; Al-Akhali et al., 2019) (16.6%) used

PMMA blocks (Telio CAD, Ivoclar Vivadent). The materials used in each included study are

presented in Table 5.

Table 5 Materials used in the included studies.

Material
CAD/CAM resin Polymer- Leucite- Lithium disilicate glass PMMA
composite infiltrated reinforced ceramics blocks
ceramics glass
ceramics
Study Lava Paradigm Vita IPS IPS Zirconia- Heat Telio CAD
Ultimate MZ 100 Enamic Empress e.max containing pressed
CAD lithium (IPS
disilicate e.max
(Vita Press.)
Suprinity)
Al-Akhali et
al. (2017)
and Al-
Akhali et al.
(2019)
Angerame et
al. (2019)
Ioannidis et
al. (2019)
Maeder et al.
(2019)
Andrade et
al. (2018)
Yazigi et al.
(2017) and
Yazigi et al.
(2018)
Sasse et al.
(2015)
Johnson et
al. (2014)
Schlichting
et al. (2011)
Magne et al.
(2010)
indicates that the material has been used in the study. Abbreviations: PMMA: poly(methyl
methacrylate),

17
3.4 Evaluation of aging procedures

Thermomechanical aging was performed in 8 studies (Al-Akhali et al., 2017; Al-

Akhali et al., 2019; Angerame and De Biasi, 2019; Ioannidis et al., 2019; Maeder et al., 2019;

Sasse et al., 2015; Yazigi et al., 2018; Yazigi et al., 2017) (66.6%). In 3 of them (25%), only

half the specimens were subjected to thermomechanical aging to test its impact on fracture

resistance of the restored teeth. In one study (Andrade et al., 2018) (8.3%), a dynamic loading

test was solely used without thermal cycling with water. The number of mechanical cycles

performed varied among different studies. Six studies (Al-Akhali et al., 2017; Al-Akhali et

al., 2019; Ioannidis et al., 2019; Maeder et al., 2019; Yazigi et al., 2018; Yazigi et al., 2017)

used 1,200,000 cycles, which is claimed to simulate a 5-year of clinical service. One study

used 1,000,000 cycles (Andrade et al., 2018), one study used 600,000 (Sasse et al., 2015), and

one study utilized 1,250,000 cycles (Angerame and De Biasi, 2019). A 6.0 mm steatite

ceramic ball was used as an antagonist in 4 studies (Al-Akhali et al., 2017; Al-Akhali et al.,

2019; Yazigi et al., 2018; Yazigi et al., 2017), whilst 2 studies (Angerame and De Biasi,

2019; Sasse et al., 2015) used a 5.0 mm steatite ceramic ball. An 8.0 mm corrosion-free steel

indenter with a rounded tip was used as an antagonist in 2 studies (Ioannidis et al., 2019;

Maeder et al., 2019). A summary of different aging methodologies utilized in the included

studies is presented in Table 6.

18
Table 6 Summary of different aging methodologies utilized in the included studies

Study Thermomechanical Number of thermomechanical Antagonist Results


aging machine cycles material and
diameter
Al-Akhali et Thermomechanical loading
al. (2019) increased the fracture
Dual-axis computerized 1,200,000 chewing cycles at a A 6.0 mm steatite resistance for ZLS, PIC and
chewing simulator loading frequency of 2.4 Hz, with ceramic balls were PM.
Al-Akhali et (Willytech, simultaneous thermocycling between used. With thermo-mechanical
al. (2017) Feldkirchen- 5 and 55 °C at a total of 5500 cycles. fatigue, no statically
Westerham, Germany) significant difference was
found in the fracture
strength values of the tested
groups.
Yazigi et al. All samples survived
(2018) and thermo-dynamic loading
Yazigi et al. except one was rated as a
(2017) failure
Ioannidis et al. Custom-made chewing 1,200,000 cycles at 1.67 Hz A corrosion-free All tested specimens
(2019) and simulator frequency combined with steel indenter with a survived the thermo-
Maeder et al. thermocycling between 5 and 50 °C) rounded tip (8.0 mechanical aging
(2019) with 120 seconds dwelling time. mm) was used as an procedures.
antagonist.

Angerame et Computerized 1,250,000 cycles at 1.0 Hz combined A steatite ceramic All samples survived after
al. (2019) masticatory simulator with thermocycling between 5 ± 3 sphere with a 5.0 thermo-mechanical loading.
(Willytech, Munich, and 55 ± 3 °C mm diameter was
Sasse et al. Germany) 600,000 mechanical cycles used as an Some specimens did not
(2015) combined with thermocycling antagonist. withstand dynamic loading
between 5 and 55 °C with a loading and were rated as a failure
frequency of 2Hz. when chipping occurs and
as a partial failure when
crack took place.
Andrade et al. Universal testing 1,000,000 cycles at 1 Hz frequency. NA All samples survived
(2018) machine (ER-11000, No thermal cycling was performed. mechanical cyclic loading
Erios, São Paulo, SP, with no chips, cracks, or
Brazil) fractures.
Abbreviations: ZLS: zirconia-containing lithium silicate glass ceramics; PIC: polymer-infiltrated
ceramics; PM: poly(methyl)methacrylate; NA: Not available

19
3.5 Assessment of mechanical test parameters

Ten out of 12 studies (Al-Akhali et al., 2017; Al-Akhali et al., 2019; Andrade et al.,

2018; Angerame and De Biasi, 2019; Ioannidis et al., 2019; Johnson et al., 2014; Maeder et

al., 2019; Sasse et al., 2015; Yazigi et al., 2018; Yazigi et al., 2017) utilized fracture strength

test. Whereas the remaining 2 (Magne et al., 2010; Schlichting et al., 2011) used fatigue

testing. Among the studies that evaluated fracture resistance, 8 studies used a crosshead speed

of 1.0 mm/min (Al-Akhali et al., 2017; Al-Akhali et al., 2019; Andrade et al., 2018;

Angerame and De Biasi, 2019; Ioannidis et al., 2019; Maeder et al., 2019; Yazigi et al., 2018;

Yazigi et al., 2017). One study (Sasse et al., 2015) used a crosshead speed of 2.0 mm/min and

another one (Johnson et al., 2014) used a crosshead speed of 0.5 mm/min. A 6.0 mm steel

ball-ended bar was used to deliver the compressive load in 5 studies (Al-Akhali et al., 2017;

Al-Akhali et al., 2019; Sasse et al., 2015; Yazigi et al., 2018; Yazigi et al., 2017). One study

(Johnson et al., 2014) used a 3.5 mm spherical stainless steel tip to simulate an opposing

cusp. Two studies (Ioannidis et al., 2019; Maeder et al., 2019) utilized 8 mm diameter round-

tip indenters in the compression mode. In the remaining two, 1 study (Angerame and De

Biasi, 2019) used a 5.0 mm round stainless steel stylus and the other one (Andrade et al.,

2018) utilized a metal sphere with a diameter of 6.0 mm. In the 2 studies (Magne et al., 2010;

Schlichting et al., 2011) that utilized fatigue testing, the specimens were subjected to cyclic

loading at a frequency of 5Hz at a maximum of 30,000 cycles. Test methodologies and the

machine setup are summarized in Table 7.

20
Table 7. Assessment of test methodologies

Study Test Crosshead Test set-up Testing machine


speed
Al-Akhali et al., 2017; Al- 1.0 mm/min A 6-mm-diameter stainless steel Universal testing machine
Akhali et al., 2019; Yazigi ball-ended bar was aligned and (Zwick Z010/TN2A, Zwick,
et al., 2018; Yazigi et al., descended at the fissure along the Ulm, Germany)
2017 long axis of the specimens.
Sasse et al., 2015 2.0 mm/min
Ioannidis et al., 2019; 1.0 mm/min An 8 mm diameter round-tip
Maeder et al., 2019 Fracture resistance indenter was used to deliver the
load at a perpendicular direction
to the occlusal surface.

Andrade et al., 2018 1.0 mm/min. A metal sphere with a diameter of Universal testing machine DL-
6.0 mm positioned to achieve 2000 (EMIC, Sa˜o Josedos
tripodization of contacts along the Pinhais, PR, Brazil
cuspal inclines over the central
fossa.
Angerame and De Biasi, 1.0 mm/min A 5.0 mm wide stainless-steel Universal testing machine
2019 stylus was positioned over the (Quasar, Galdabini, Cardano al
central fossa to achieve Campo, Italy)
tripodization of contacts along the
cuspal inclines.
Johnson et al., 2014 0.5 mm/min A custom 3.5 mm diameter Universal testing machine
spherical stainless-steel tip was (INSTRON 5567, NORWOOD,
used to achieve equalized tripod MA)
contacts along the cuspal inclines
surrounding the central fossae.
Schlichting et al., 2011 Fatigue resistance Not A 7.0 mm diameter resin Masticatory forces were
Magne et al., 2010 available composite sphere post simulated using closed-loop
o
polymerized at 100 C for 5 servo hydraulics (Mini Bionix
minutes was used. The sphere was II; MTS Systems Corp, Eden
positioned to achieve tripod Prairie, Minn)
contact.

21
3.6 Assessment of bonding substrate

In 8 studies, all occlusal veneers were adhesively bonded to sound dentin (Andrade et

al., 2018; Angerame and De Biasi, 2019; Johnson et al., 2014; Maeder et al., 2019; Magne et

al., 2010; Schlichting et al., 2011; Yazigi et al., 2018; Yazigi et al., 2017). Whereas in 3

studies (Al-Akhali et al., 2017; Al-Akhali et al., 2019; Ioannidis et al., 2019), the restorations

were bonded to enamel. In the remaining study, occlusal veneers were bonded to enamel,

dentin, or composite (Sasse et al., 2015). Cementation protocols for all the included studies

are presented in Table 8.

Table 8 Bonding protocols

Study Bonding Bonding protocol Resin cement


Substrate
(Al-Akhali et al., Sound enamel A self-etching primer (Multilink Primer A/B; Ivoclar Dual-cure composite resin cement
2019)
Vivadent AG) was used. (Multilink Automix; Ivoclar
Vivadent AG).
(Al-Akhali et al., Sound enamel The prepared enamel surface was etched with 37%
2017)
phosphoric acid followed by application of Multilink 7
Primer A/B (Ivoclar Vivadent)
(Angerame and Sound dentin 1.Immediate dentin sealing using a self-etch adhesive Dual-cure resin cement (Variolink
De Biasi, 2019)
system (Clearfil SE Bond 2, Kuraray, Osaka, Japan). II, shade A3, Ivoclar-Vivadent).
2. Pre-cementation protocol involved etching of
enamel with 37% orthophosphoric acid followed by
adhesive resin application (Clearfil SE Bond 2,
Kuraray).
(Ioannidis et al., Sound enamel The enamel surface of all prepared teeth was etched 1. For e.max: dual-cure resin
2019)
with 35% phosphoric acid followed by the application cement (Variolink II, shade A3,
(Maeder et al., Sound dentin
2019) of the adhesive system. Ivoclar-Vivadent).
1. For e.max and Vita Enamic: Heliobond; 2. For Vita Enamic: Tetric Flow;
Ivoclar Vivadent. Ivoclar Vivadent.
2. For Lava Ultimate: Scotchbond Universal Adhesive, 3. For Lava Ultimate: RelyX
3M ESPE. Ultimate cement; 3M
3. For zirconia: ED Primer A and B; Kuraray, Tokyo, ESPE
Japan. 4. For zirconia Panavia 21;
Kuraray

22
Table 8 Bonding protocols (cont’d)

(Andrade et al., Sound dentin 1. For teeth restored with e.max CAD and Vita 1. For e.max and Vita Enamic,
2018)
Enamic, both the enamel and dentin surfaces were Variolink N (Ivoclar Vivadent,
etched with 37% phosphoric acid followed by Schaan, Liechtenstein).
application of the adhesive system (ExciTE F DSC,
Ivoclar Vivadent, Schaan, Liechtenstein) 2. For Lava Ultimate, RelyX
2. For teeth restored with Lava Ultimate, the enamel Ultimate resin cement (3M, St
surface with etched with 35% phosphoric acid Paul, MN, USA).
followed by application of Single Bond Universal
adhesive system (3M, St Paul, MN, USA).
(Yazigi et al., Sound dentin The specimens were assigned into 3 main groups: with Dual-cure resin cement (Variolink
2018; Yazigi et no immediate dentin sealing, immediate dentin sealing Esthetic DC, Ivoclar Vivadent).
al., 2017) with a total-etch protocol, and immediate dentin
sealing with selective etching. Adhese Universal
(Ivoclar Vivadent) was used.

(Sasse et al., The restorations A self-etching primer (Multilink Primer A and B, A self-curing luting composite
2015)
where either Ivoclar Vivadent, Schaan, Liechtenstein) was used. (Multilink Automix, Ivoclar
bonded solely to Vivadent)
enamel, within
enamel and
dentin or to an
occlusal
composite filling
(Johnson et al., Sound dentin The prepared tooth surface was etched with 37.5% Self-adhesive resin cement
2014)
phosphoric acid without dentin desiccation. (RelyX Unicem, 3M ESPE, St.
Paul, Minnesota)
(Schlichting et Sound dentin Immediate dentin sealing using a 3-step etch-and-rinse A luting material (Filtek Z100;
al., 2011)
protocol (OptiBond FL; Kerr Corp, Orange, Calif). 3M ESPE) preheated at 68oC in
(Magne et al.,
2010) Pre-cementation protocol involved etching with 37.5% Calset (AdDent, Danbury, Conn)
phosphoric acid followed by OptiBond adhesive was used.
application.

23
3.7 Assessment of failure

Table 9 lists the failure mode assessment methods among these studies. Failure

assessment methodologies showed variations between different studies. In 2 studies (16.6%),

the failure mode was evaluated visually (Andrade et al., 2018; Angerame and De Biasi,

2019). An optical microscope with a light source (LED or transillumination) with variable

magnification power was used in 4 studies (Al-Akhali et al., 2017; Al-Akhali et al., 2019;

Magne et al., 2010; Schlichting et al., 2011). Digital photographs with the aid of x2.5 loupes

were utilized in failure detection in 2 studies(Ioannidis et al., 2019; Maeder et al., 2019),

while 3 studies (Johnson et al., 2014; Sasse et al., 2015; Yazigi et al., 2017) used

stereomicroscope with variable magnification. One study (Yazigi et al., 2017) has not clearly

reported the failure detection methodology.

Irreparable and reparable fracture criteria were implemented in 1 study (Andrade et

al., 2018) (8.3%). The fracture was considered reparable if it included only the restoration or

part or all the cusps. When the line of fracture divided the tooth into two parts at the level of

the pulp chamber floor, it was regarded as irreparable. In 3 studies (Al-Akhali et al., 2017;

Al-Akhali et al., 2019; Yazigi et al., 2018) (25%), the failure mode was classified into 4

categories according to predetermined criteria with regards to the fracture features on the

restoration and tooth. Two studies (Magne et al., 2010; Schlichting et al., 2011) (16.6%) had

specific criteria for failure, the specimen had to demonstrate 1 or more surface cracks more

than or equal to 2.0 mm in length. The crack detection process proceeded until catastrophic

failure or completion of 185,000 cycles. Out of the remaining 6 studies, 2 studies (Ioannidis

et al., 2019; Maeder et al., 2019) (16.6 %) classified failure mode using 0 to 3 scoring marks

to indicate the severity of visual fractures at different substrates. One study (Angerame and

De Biasi, 2019) (8.3%) distinguished between failures superior or inferior to the

24
cementoenamel junction, and between fractures and cracks (fracture without fragment

detachment). Also, one study (Johnson et al., 2014) (8.3%) categorized failure modes

according only to the occurrence of the locations. In the remaining 2 studies (Sasse et al.,

2015; Yazigi et al., 2017), failure mode was not clearly stated.

25
Table 9 Assessment of failure mode

Study Failure mode classification Predominant failure mode


(Al-Akhali et al., 2019) I = extensive crack formation within The most observed failure patterns
the restoration were class (I) and class (III) for all
II = cohesive fracture the groups.
(Al-Akhali et al., 2017) III = adhesive fracture The most common failure modes
IV = longitudinal fracture of the were class (I) and class (III) for all
restoration and the tooth the studied groups.
(Yazigi et al., 2017) The predominant types of failure
(66%) were modes (III) and (IV).
(Ioannidis et al., 2019) Score 0 = no visible fracture In control groups, score 2 fracture
Score 1 = cohesive fracture was predominant (80%).
(restoration only) In the test groups, failure modes
Score 2 = cohesive fracture were variable with score 2 being the
(restoration cement) most common.
(Maeder et al., 2019) Score 3 = fracture of restoration- Score 2 was the most common
cement-tooth complex among all groups.
(Angerame and De Biasi, Failure was classified into The predominant failure mode was
2019) fracture above CEJ/fracture below fracture above CEJ.
CEJ
(Andrade et al., 2018) Repairable and irreparable Repairable fracture was the
Repairable: when the fracture predominant failure mode.
involved only the restoration or part
of the cusps.
Irreparable: when the fracture line
divided the tooth into two parts at
the floor level of the pulp chamber
(Johnson et al., 2014) Mode 1 = restoration only About 36.6% of the samples showed
Mode 2 = restoration and enamel mode 2 failure,38.3% showed mode
Mode 3 = restoration, enamel, and 3 and 25% showed mode 1.
dentin.
(Schlichting et al., 2011) To meet the criteria of failure Failure modes showed variation
specimens had to exhibit 1 or more among the groups with none of the
surface cracks greater than 2mm in groups demonstrating catastrophic
length. failure *.
(Magne et al., 2010)

*catastrophic failure = loss of a restoration fragment

26
4. Discussion

Systematic reviews provide practitioners with access to prefiltered evidence as they

summarize the available knowledge on a particular topic with implementing strategies that

decrease bias (Cook et al. 1997). In the same context, systematic reviews decrease the time

and expertise it would take to identify, locate, and appraise individual studies while adhering

to scientific methodology (Petticrew and Roberts, 2008).

Nowadays, preservation of the remaining tooth structure is an important aspect of

restorative dentistry (Angerame and De Biasi, 2019). To maintain the balance between

biological, functional, and aesthetic parameters, conservation of remaining tooth structure is

pivotal (Vailati and Belser, 2008b, c). In an attempt to shift towards conservative approaches,

occlusal veneers have been proposed as a substitute for onlays and full-coverage crowns in

the management of extensive erosive/abrasive defects. In the last decade, several trials have

been made towards the laboratory assessment of occlusal veneers before adopting it as a

clinically viable treatment option. The rationale for conducting this systematic review was to

assess in vitro fracture and fatigue resistance of occlusal veneers under variable thicknesses,

materials, and artificial aging procedures.

Fracture strength of all-ceramic restorations is affected by multiple factors as internal

microstructure, the technique of fabrication, preparation design, and luting method (Lima et

al., 2013). The results of this systematic review showed that glass ceramic occlusal veneers

with a thickness of 0.7-1.0 mm achieved strength values exceeding the recommended

minimum fracture strength values for posterior restorations ranging from 500-700 N (Körber

and Ludwig, 1983; Sasse et al., 2015). These results came in agreement with a previous study

(Bakeman et al., 2015) in which decreasing ceramic thickness to 1.0 did not influence

fracture resistance of posterior ceramic partial coverage restorations. Moreover, the outcomes

27
of 2 recent studies (Ioannidis et al., 2019; Maeder et al., 2019) revealed that 1.0 mm lithium

disilicate occlusal veneers demonstrated fracture resistance values not different from full-

coverage crowns. However, when the thickness of lithium disilicate glass ceramic was

decreased to 0.5 mm, fracture resistance values were significantly inferior to full-coverage

porcelain fused to metal crowns. It is important to mention that this result was obtained when

the restorations were bonded to dentin (Maeder et al., 2019). The stability of ceramic

restorations could be increased by decreasing the mismatch between ceramics and supporting

structures (Ma et al., 2013). With minimal thickness restorations, it was found that the load-

bearing capacity of 1.0 mm thick lithium disilicate occlusal onlay supported by enamel was

not significantly different from 1.4 mm thick lithium disilicate occlusal onlay supported by

dentin. This could be explained by the smaller mismatch ratio between lithium disilicate and

enamel in comparison to lithium disilicate and dentin (Ma et al., 2013). Moreover, at a

thickness of 0.5 mm lithium disilicate glass ceramic demonstrated cracks at 450 N (Maeder et

al., 2019). This could indicate that such minimal thickness is not ideal for restoration in

patients with high loading forces.

The type of bonding substrate as well as adhesive protocols should not be neglected

when evaluating the fracture strength of ceramic restorations (Bakeman et al., 2015). Lithium

disilicate has relatively high flexural strength of nearly 400 MPa and fracture toughness value

of 3.3MPa● m1/2 (Albakry et al., 2003a; Albakry et al., 2003b). In comparison, resin

composite blocks (e.g. Lava Ultimate) and polymer-infiltrated ceramics blocks (e.g. Vita

Enamic) have flexural strengths of 200 MPa (Lauvahutanon et al., 2014) and 110-160 MPa

(Argyrou et al., 2016; Coldea et al., 2013a; Johnson et al., 2014), respectively. Lithium

disilicate glass ceramic has the crystalline phase that provides a tight interlocking-like matrix

preventing the propagation of microcracks (Della Bona et al., 2004; Holand et al., 2000).

However, due to the brittle nature of glass ceramic, adhesion should be done with resin

28
cement and the presence of an enamel substrate can favorably affect the predictability of

ceramic restorations (Fleming and Narayan, 2003; Layton and Walton, 2007; Peumans et al.,

2004). The influence of bonding substrate and luting protocol on the survival of glass ceramic

occlusal veneers is emphasized by Sasse et al. (Sasse et al., 2015), where lithium disilicate

occlusal veneers bonded to enamel with a thickness of 0.5-0.8 mm demonstrated cracks after

thermomechanical aging. This could be related to the use of a self-etching primer on enamel

rather than a separate etching step.

The quality of the bond between resin cement and ceramic is pivotal for the clinical

success of any ceramic restoration (Soares et al., 2005). Surface treatment of ceramics can

affect the bonding mechanism by promoting micromechanical and/ or chemical retention in

the ceramic surface (Blatz, 2014; Blatz et al., 2003; Johnson et al., 2014). The optimal

surface treatment for silica-based ceramics is hydrofluoric acid etching followed by silane

application (Matinlinna et al., 2018; Tian et al., 2014). As for polymer-infiltrated ceramics,

the International Academy for Adhesive Dentistry (IAAD) recommends pre-treatment with

hydrofluoric acid and subsequent silane application (Ozcan and Volpato, 2016). Furthermore,

the IAAD recommends pre-treatment of CAD/CAM resin composite with air abrasion with

50 μm aluminum oxides or 30 μm of silicon dioxide at a pressure of 2 bar (Ozcan and

Volpato, 2016). In this systematic review, different concentrations of hydrofluoric acid were

used to etch the surface of glass ceramic between 5% and 9% for 20 seconds. As for polymer-

infiltrated ceramics, etching time was 60 seconds. Indeed, bonding protocols and types of

resin cement varies among different manufacturer instructions for various materials. Our

included studies have shown the role of a reliable bonding and protocol can enhance the

mechanical performance of thin occlusal restorations (Yazigi et al., 2018; Yazigi et al., 2017).

In addition, a significantly higher fracture resistance for glass ceramic occlusal veneers was

obtained when immediate dentin sealing was performed, regardless of the pre-cementation

29
etching protocol (total-etch or self-etch). This could be explained on the grounds that freshly

exposed dentin without contamination by temporary cement or impression can represent the

most favorable substrate for bonding, and dentin bonding needs to be developed without the

effect of occlusal stresses from restoration placement (Magne, 2005).

In contrary to glass ceramic, polymer-infiltrated ceramics and resin composite blocks

were successfully used in a thickness of 0.5 mm and even up to 0.3 mm. In one study

(Johnson et al., 2014), CAD/CAM resin composite with a thickness of 0.3, 0.6 and 1.0 mm

showed fracture resistance values above the achievable human masticatory forces. This could

indicate that such minimalistic restorations could achieve clinical success in ideal

circumstances and any complications could be attributed to factors beyond normal maximum

occlusal loading (Johnson et al., 2014). However, the results of this study should be carefully

assessed considering that thermal cycling and dynamic loading were not performed which

render it less clinically relevant.

Determining the impact of minor changes in thickness on fracture resistance of

anatomic restorations can be challenging due to complex geometry and cusp height as

opposed to material discs. In 2 studies (Al-Akhali et al., 2017; Al-Akhali et al., 2019),

restoration thickness was adjusted according to fossa/cusp thickness (0.5-0.8 mm). Intra-oral

scanner was used in 2 studies (Ioannidis et al., 2019; Maeder et al., 2019) to measure the

thickness by taking two scans for each of the two groups of thicknesses (0.5 and 1.0 mm). For

the group with 0.5 mm thickness, the tooth was scanned twice, with the second scan in 0.5

mm infra-position. The difference between the initial scan and the scan in infra-position

served as the source of information for the software to design. Nevertheless, the accuracy of

using an intra-oral scanner to measure the thickness is vulnerable (Nedelcu et al., 2018).

Furthermore, in terms of shape, Sasse et al. have used occlusal veneers with a semi-anatomic

30
shaping to achieve a constant ceramic thickness (Sasse et al., 2015). The tooth was virtually

elevated and reduced again in the fissure area until the desired thickness was obtained.

Therefore, caution should be taken in reading the data, and standardization of thickness

measurement is necessary and important to accurately determine its effect on fracture

resistance values.

Laboratory investigations including thermomechanical aging are important in testing

dental materials with conditions simulating the oral environment. Thermomechanical stresses

occur at the adhesive interface owing to variations in the coefficient of thermal expansion of

different dental materials that can affect their mechanical properties (Oyafuso et al., 2008).

The results of this systematic review seem to be conflicting regarding the impact of

thermomechanical aging on the fracture strength of occlusal veneers. In a study by Al-Akhali

et al., 2019, thermomechanical fatigue significantly decreased the final fracture strength of

occlusal veneers manufactured from polymeric materials, while no significant reduction in

strength was noticed in lithium disilicate glass ceramic. This could be attributed to thermal

expansion and shrinkage in polymeric contents of these materials, which could have

accelerated fatigue during thermomechanical aging and consequently decreased their fracture

strength. Interestingly, in a previous study by the same author (Al-Akhali et al., 2017),

thermomechanical aging significantly increased the fracture resistance of polymeric materials

(polymer-infiltrated ceramics and PMMA). The authors attributed this result to the post-

curing effect and stress relaxation due to the warm temperature of thermocycling water (Par

et al., 2014; Vouvoudi and Sideridou, 2013). These conflicting results may be attributed to

different bonding protocols followed in the 2 studies. One study followed an etch-and-rinse

technique (Al-Akhali et al., 2017), and the other (Al-Akhali et al., 2019) used a self-etching

primer. Both studies used enamel as a bonding substrate.

31
Multiple factors might influence the validity of thermomechanical testing. For

example, the specimen stiffness effect on contact forces and impact velocity were shown to

be correlated with the relative overshot force (Conserva et al., 2008; Rues et al., 2011).

Fluctuations in contact forces over the test chambers in the widely-used chewing simulator

and stiffness of the specimens can lead to variations in contact forces. Furthermore, the

material of the antagonist might play a role in decreasing fracture loads (Nawafleh et al.,

2020). A recent study (Nawafleh et al., 2020) has shown steatite ceramic and steel indenters

were found to decrease the load of zirconia crowns when compared to tungsten carbide.

Therefore, test parameters should be standardized in future studies to obtain a reliable

estimate of clinical performance. In fracture resistance testing, the shape, diameter, and

position of the load application device define the contact with the tooth or the restorative

material (Silva et al., 2012). In this systematic review, studies employed a tripod contact

along the cuspal inclines to distribute stresses more evenly which can substantiate the

reliability of the data.

Furthermore, failed specimens during thermomechanical fatigue manufactured from

polymeric materials (polymer-infiltrated ceramics and PMMA) were capable of surviving

more masticatory cycles when compared to glass ceramic and zirconia-containing glass

ceramic (Al-Akhali et al., 2019). These outcomes came in agreement with previous

laboratory studies in which thin CAD/CAM resin composite occlusal veneers had a

significantly higher stepwise fatigue resistance compared to lithium-disilicate glass ceramic

(Magne et al., 2010; Schlichting et al., 2011). However, these studies have not used

thermocycling that makes the comparison with the aforementioned study difficult and

limited.

32
Evaluating fatigue strength of occlusal veneers has clinical relevance as restorations

are subjected to millions of masticatory cycles in the oral cavity which can result in a

significant reduction of strength due to fatigue (Magne et al., 2012). In the 2 studies (Magne

et al., 2010; Schlichting et al., 2011) that evaluated fatigue strength of occlusal veneers,

CAD/CAM resin composite had higher fatigue resistance in comparison with glass ceramic.

This could be attributed to the fact that failure induced by tensile stresses is more sensitive to

the ratio of elastic modulus between restoration, luting agent, and dentin than it is to uniaxial

flexural strength of the material and restoration thickness (Magne et al., 2010; Schlichting et

al., 2011). Thus, the higher performance of CAD/CAM resin composite restorations could be

related to the relative similarity between the elastic modulus of composite (approximately 16-

20 GPa), and dentin (approximately 17.7–29.8 GPa) (De Munck et al., 2005; Lauvahutanon

et al., 2014).

Failure pattern evaluation showed considerable variation among the included studies.

In two studies (Ioannidis et al., 2019; Maeder et al., 2019) that followed the same criteria for

failure evaluation, a cohesive fracture that involved the restoration and cement layer with no

damage to the underlying tooth structure was the most common. Furthermore, (Magne et al.,

2010) and (Schlichting et al., 2011) reported that fractures or cracks in occlusal veneers were

limited to restorative materials. These positive outcomes come in line with the principles of

minimally invasive dentistry. Failures that don not involve damage to tooth structure increase

the longevity and prognosis of the restored teeth because the veneer can be replaced.

Conversely, when the underlying tooth structure is involved, invasive approaches including

endodontic treatment and extractions might be required.

This systematic review presents some limitations as the exclusion of non-English

manuscripts and variations in methodologies with high heterogenicity among the studies.

33
Regarding the risk of bias, 50% of the studies showed a high risk of bias. In addition,

differences in type (premolar or molars) and dimensions of extracted teeth among studies

eventually led to variations in fracture resistance values. Extracted teeth, even if carefully

selected, show morphological variations, flaws, and irregularities which might have resulted

in variations in fracture resistance values. The impact of bonding substrate and adhesion on

the longevity of occlusal veneers needs further investigations. While durable bond to enamel

is well established, in clinical situations in which teeth are affected by erosion, dentin

exposure is inevitable. Also, variations in cement type and their adhesion to dentin need

further evaluation. Furthermore, due to wide variations in methodologies and materials, a

meta-analysis could not be conducted.

5. Conclusions

Occlusal veneers could potentially represent a reliable treatment option for patients

with extensive tooth wear. Lithium disilicate glass ceramic showed more favorable results in

terms of fracture strength at a thickness of 0.7-1.0 mm. The presence of an enamel substrate

in addition to the use of etch-and-rinse bonding protocol can favorably affect the mechanical

behavior of lithium disilicate glass ceramic. Polymer-infiltrated ceramics and resin-nano

ceramics could be used successfully with a thickness of less than 1.0 mm. This could

potentially be a conservative treatment option for patients with extreme tooth wear with a

need to increase the vertical dimension of occlusion without sacrificing sound tooth structure.

The scientific database is still lacking in clinical trials that evaluate the longevity and clinical

performance of this type of restorations.

34
References:

Al-Akhali, M., Chaar, M.S., Elsayed, A., Samran, A., Kern, M., 2017. Fracture resistance of ceramic
and polymer-based occlusal veneer restorations. J Mech Behav Biomed Mater 74, 245-250.
Al-Akhali, M., Kern, M., Elsayed, A., Samran, A., Chaar, M.S., 2019. Influence of thermomechanical
fatigue on the fracture strength of CAD-CAM-fabricated occlusal veneers. J Prosthet Dent 121, 644-
650.
Albakry, M., Guazzato, M., Swain, M.V., 2003a. Biaxial flexural strength, elastic moduli, and x-ray
diffraction characterization of three pressable all-ceramic materials. The Journal of prosthetic
dentistry 89, 374-380.
Albakry, M., Guazzato, M., Swain, M.V., 2003b. Fracture toughness and hardness evaluation of three
pressable all-ceramic dental materials. J Dent 31, 181-188.
Andrade, J.P., Stona, D., Bittencourt, H.R., Borges, G.A., Burnett, L.H.J., Spohr, A.M., 2018. Effect
of Different Computer-aided Design/Computer-aided Manufacturing (CAD/CAM) Materials and
Thicknesses on the Fracture Resistance of Occlusal Veneers. Oper Dent 43, 539-548.
Angerame, D., De Biasi, M., 2019. Influence of preparation designs on marginal adaptation and
failure load of full-coverage occlusal veneers after thermomechanical aging simulation.
Argyrou, R., Thompson, G.A., Cho, S.H., Berzins, D.W., 2016. Edge chipping resistance and flexural
strength of polymer infiltrated ceramic network and resin nanoceramic restorative materials. J
Prosthet Dent 116, 397-403.
Bakeman, E.M., Rego, N., Chaiyabutr, Y., Kois, J.C., 2015. Influence of ceramic thickness and
ceramic materials on fracture resistance of posterior partial coverage restorations. Oper Dent 40, 211-
217.
Beuer, F., Schweiger, J., Edelhoff, D., 2008. Digital dentistry: an overview of recent developments for
CAD/CAM generated restorations. Br Dent J 204, 505-511.
Bindl, A., Luthy, H., Mormann, W.H., 2006. Strength and fracture pattern of monolithic CAD/CAM-
generated posterior crowns. Dent Mater 22, 29-36.
Blatz, M.B., 2014. Bonding protocols for improved long-term clinical success. Compend Contin Educ
Dent 35, 276-277.
Blatz, M.B., Sadan, A., Kern, M., 2003. Resin-ceramic bonding: a review of the literature. J Prosthet
Dent 89, 268-274.
Chen, C., Trindade, F.Z., de Jager, N., Kleverlaan, C.J., Feilzer, A.J., 2014. The fracture resistance of
a CAD/CAM Resin Nano Ceramic (RNC) and a CAD ceramic at different thicknesses. Dent Mater
30, 954-962.
Coldea, A., Swain, M.V., Thiel, N., 2013a. In-vitro strength degradation of dental ceramics and novel
PICN material by sharp indentation. J Mech Behav Biomed Mater 26, 34-42.

35
Coldea, A., Swain, M.V., Thiel, N., 2013b. Mechanical properties of polymer-infiltrated-ceramic-
network materials. Dent Mater 29, 419-426.
Conserva, E., Menini, M., Tealdo, T., Bevilacqua, M., Pera, F., Ravera, G., Pera, P., 2008. Robotic
chewing simulator for dental materials testing on a sensor-equipped implant setup. Int J Prosthodont
21, 501-508.
De Munck, J., Van Landuyt, K., Peumans, M., Poitevin, A., Lambrechts, P., Braem, M., Van
Meerbeek, B., 2005. A critical review of the durability of adhesion to tooth tissue: methods and
results. J Dent Res 84, 118-132.
Della Bona, A., Shen, C., Anusavice, K.J., 2004. Work of adhesion of resin on treated lithia disilicate-
based ceramic. Dent Mater 20, 338-344.
Dirxen, C., Blunck, U., Preissner, S., 2013. Clinical performance of a new biomimetic double network
material. The open dentistry journal 7, 118-122.
Egilmez, F., Ergun, G., Cekic-Nagas, I., Vallittu, P.K., Lassila, L.V.J., 2018. Does artificial aging
affect mechanical properties of CAD/CAM composite materials. J Prosthodont Res 62, 65-74.
El-Damanhoury, H.M., Haj-Ali, R.N., Platt, J.A., 2015. Fracture resistance and microleakage of
endocrowns utilizing three CAD-CAM blocks. Oper Dent 40, 201-210.
Fleming, G.J., Narayan, O., 2003. The effect of cement type and mixing on the bi-axial fracture
strength of cemented aluminous core porcelain discs. Dent Mater 19, 69-76.
Guess, P.C., Schultheis, S., Wolkewitz, M., Zhang, Y., Strub, J.R., 2013. Influence of preparation
design and ceramic thicknesses on fracture resistance and failure modes of premolar partial coverage
restorations. J Prosthet Dent 110, 264-273.
Habib, S.R., Otaibi, A.K.A., Anazi, T.A.A., Anazi, S.M.A., 2018. Evaluation of thickness of
CAD/CAM fabricated zirconia cores by digital microscope. Technol Health Care 26, 181-185.
Holand, W., Schweiger, M., Frank, M., Rheinberger, V., 2000. A comparison of the microstructure
and properties of the IPS Empress 2 and the IPS Empress glass-ceramics. J Biomed Mater Res 53,
297-303.
Homaei, E., Farhangdoost, K., Pow, E.H.N., Matinlinna, J.P., Akbari, M., Tsoi, J.K.H., 2016a.
Fatigue resistance of monolithic CAD/CAM ceramic crowns on human premolars. Ceram Int 42,
15709-15717.
Homaei, E., Farhangdoost, K., Tsoi, J.K.H., Matinlinna, J.P., Pow, E.H.N., 2016b. Static and fatigue
mechanical behavior of three dental CAD/CAM ceramics. J Mech Behav Biomed 59, 304-313.
Homaei, E., Jin, X.Z., Pow, E.H.N., Matinlinna, J.P., Tsoi, J.K.H., Farhangdoost, K., 2018. Numerical
fatigue analysis of premolars restored by CAD/CAM ceramic crowns. Dent Mater 34, E149-E157.
Ioannidis, A., Muhlemann, S., Ozcan, M., Husler, J., Hammerle, C.H.F., Benic, G.I., 2019. Ultra-thin
occlusal veneers bonded to enamel and made of ceramic or hybrid materials exhibit load-bearing
capacities not different from conventional restorations. J Mech Behav Biomed Mater 90, 433-440.

36
Johnson, A.C., Versluis, A., Tantbirojn, D., Ahuja, S., 2014. Fracture strength of CAD/CAM
composite and composite-ceramic occlusal veneers. J Prosthodont Res 58, 107-114.
Körber, K., Ludwig, K., 1983. Maximum biteforces as a basis for the design of dental restorations.
Dent Labor 31, 55-60.
Kassem, A.S., Atta, O., El-Mowafy, O., 2012. Fatigue resistance and microleakage of CAD/CAM
ceramic and composite molar crowns. J Prosthodont 21, 28-32.
Lauvahutanon, S., Takahashi, H., Shiozawa, M., Iwasaki, N., Asakawa, Y., Oki, M., Finger, W.J.,
Arksornnukit, M., 2014. Mechanical properties of composite resin blocks for CAD/CAM. Dent Mater
J 33, 705-710.
Layton, D., Walton, T., 2007. An up to 16-year prospective study of 304 porcelain veneers.
International Journal of Prosthodontics 20, 389.
Leung, B.T.W., Tsoi, J.K.H., Matinlinna, J.P., Pow, E.H.N., 2015. Comparison of mechanical
properties of three machinable ceramics with an experimental fluorophlogopite glass ceramic. J
Prosthet Dent 114, 440-446.
Lima, J.M., Souza, A.C., Anami, L.C., Bottino, M.A., Melo, R.M., Souza, R.O., 2013. Effects of
thickness, processing technique, and cooling rate protocol on the flexural strength of a bilayer ceramic
system. Dent Mater 29, 1063-1072.
Ma, L., Guess, P.C., Zhang, Y., 2013. Load-bearing properties of minimal-invasive monolithic
lithium disilicate and zirconia occlusal onlays: finite element and theoretical analyses. Dent Mater 29,
742-751.
Maeder, M., Pasic, P., Ender, A., Ozcan, M., Benic, G.I., Ioannidis, A., 2019. Load-bearing capacities
of ultra-thin occlusal veneers bonded to dentin. J Mech Behav Biomed Mater 95, 165-171.
Maghami, E., Homaei, E., Farhangdoost, K., Pow, E.H.N., Matinlinna, J.P., Tsoi, J.K.H., 2018. Effect
of preparation design for all-ceramic restoration on maxillary premolar: a 3D finite element study. J
Prosthodont Res 62, 436-442.
Magne, P., 2005. Immediate dentin sealing: a fundamental procedure for indirect bonded restorations.
J Esthet Restor Dent 17, 144-154; discussion 155.
Magne, P., Hanna, J., Magne, M., 2013. The case for moderate "guided prep" indirect porcelain
veneers in the anterior dentition. The pendulum of porcelain veneer preparations: from almost no-prep
to over-prep to no-prep. Eur J Esthet Dent 8, 376-388.
Magne, P., Knezevic, A., 2009a. Influence of overlay restorative materials and load cusps on the
fatigue resistance of endodontically treated molars. Quintessence Int 40, 729-737.
Magne, P., Knezevic, A., 2009b. Simulated fatigue resistance of composite resin versus porcelain
CAD/CAM overlay restorations on endodontically treated molars. Quintessence Int 40, 125-133.
Magne, P., Schlichting, L.H., Maia, H.P., Baratieri, L.N., 2010. In vitro fatigue resistance of
CAD/CAM composite resin and ceramic posterior occlusal veneers. J Prosthet Dent 104, 149-157.

37
Magne, P., Stanley, K., Schlichting, L.H., 2012. Modeling of ultrathin occlusal veneers. Dent Mater
28, 777-782.
Mainjot, A.K., Dupont, N.M., Oudkerk, J.C., Dewael, T.Y., Sadoun, M.J., 2016. From Artisanal to
CAD-CAM Blocks: State of the Art of Indirect Composites. J Dent Res 95, 487-495.
Matinlinna, J.P., Lung, C.Y.K., Tsoi, J.K.H., 2018. Silane adhesion mechanism in dental applications
and surface treatments: A review. Dent Mater 34, 13-28.
Morimoto, S., Rebello de Sampaio, F.B., Braga, M.M., Sesma, N., Ozcan, M., 2016. Survival Rate of
Resin and Ceramic Inlays, Onlays, and Overlays: A Systematic Review and Meta-analysis. J Dent Res
95, 985-994.
Nawafleh, N., Bibars, A.R., Al Twal, E., Ochsner, A., 2020. Influence of Antagonist Material on
Fatigue and Fracture Resistance of Zirconia Crowns. Eur J Dent.
Nedelcu, R., Olsson, P., Nystrom, I., Ryden, J., Thor, A., 2018. Accuracy and precision of 3 intraoral
scanners and accuracy of conventional impressions: A novel in vivo analysis method. J Dent 69, 110-
118.
Oyafuso, D.K., Ozcan, M., Bottino, M.A., Itinoche, M.K., 2008. Influence of thermal and mechanical
cycling on the flexural strength of ceramics with titanium or gold alloy frameworks. Dent Mater 24,
351-356.
Ozcan, M., Volpato, C.A., 2016. Surface Conditioning and Bonding Protocol for Polymer-infiltrated
Ceramic: How and Why? J Adhes Dent 18, 174-175.
Par, M., Gamulin, O., Marovic, D., Klaric, E., Tarle, Z., 2014. Effect of temperature on post-cure
polymerization of bulk-fill composites. J Dent 42, 1255-1260.
Petticrew, M., Roberts, H., 2008. Systematic reviews in the social sciences: A practical guide. John
Wiley & Sons.
Peumans, M., De Munck, J., Fieuws, S., Lambrechts, P., Vanherle, G., Van Meerbeek, B., 2004. A
prospective ten-year clinical trial of porcelain veneers. The journal of adhesive dentistry 6, 65-76.
Pjetursson, B.E., Sailer, I., Zwahlen, M., Hammerle, C.H., 2007. A systematic review of the survival
and complication rates of all-ceramic and metal-ceramic reconstructions after an observation period of
at least 3 years. Part I: Single crowns. Clin Oral Implants Res 18 Suppl 3, 73-85.
Rosa, W.L., Piva, E., Silva, A.F., 2015. Bond strength of universal adhesives: A systematic review
and meta-analysis. J Dent 43, 765-776.
Rues, S., Huber, G., Rammelsberg, P., Stober, T., 2011. Effect of impact velocity and specimen
stiffness on contact forces in a weight-controlled chewing simulator. Dent Mater 27, 1267-1272.
Sarkis-Onofre, R., Skupien, J.A., Cenci, M.S., Moraes, R.R., Pereira-Cenci, T., 2014. The role of
resin cement on bond strength of glass-fiber posts luted into root canals: a systematic review and
meta-analysis of in vitro studies. Oper Dent 39, E31-44.

38
Sasse, M., Krummel, A., Klosa, K., Kern, M., 2015. Influence of restoration thickness and dental
bonding surface on the fracture resistance of full-coverage occlusal veneers made from lithium
disilicate ceramic. Dent Mater 31, 907-915.
Schlichting, L.H., Maia, H.P., Baratieri, L.N., Magne, P., 2011. Novel-design ultra-thin CAD/CAM
composite resin and ceramic occlusal veneers for the treatment of severe dental erosion. J Prosthet
Dent 105, 217-226.
Silva, G.R.d., Silva, N.R.d., Soares, P.V., Costa, A.R., Fernandes-Neto, A.J., Soares, C.J., 2012.
Influence of different load application devices on fracture resistance of restored premolars. Brazilian
Dental Journal 23, 484-489.
Skouridou, N., Pollington, S., Rosentritt, M., Tsitrou, E., 2013. Fracture strength of minimally
prepared all-ceramic CEREC crowns after simulating 5 years of service. Dent Mater 29, e70-77.
Soares, C.J., Soares, P.V., Pereira, J.C., Fonseca, R.B., 2005. Surface treatment protocols in the
cementation process of ceramic and laboratory-processed composite restorations: a literature review. J
Esthet Restor Dent 17, 224-235.
Spitznagel, F.A., Scholz, K.J., Strub, J.R., Vach, K., Gierthmuehlen, P.C., 2018. Polymer-infiltrated
ceramic CAD/CAM inlays and partial coverage restorations: 3-year results of a prospective clinical
study over 5 years. Clin Oral Investig 22, 1973-1983.
Tekce, N., Pala, K., Demirci, M., Tuncer, S., 2016. Changes in surface characteristics of two different
resin composites after 1 year water storage: An SEM and AFM study. Scanning 38, 694-700.
Tian, T., Tsoi, J.K.H., Matinlinna, J.P., Burrow, M.F., 2014. Aspects of bonding between resin luting
cements and glass ceramic materials. Dent Mater 30, E147-E162.
Vailati, F., Belser, U.C., 2008a. Full-mouth adhesive rehabilitation of a severely eroded dentition: the
three-step technique. Part 1. Eur J Esthet Dent 3, 30-44.
Vailati, F., Belser, U.C., 2008b. Full-mouth adhesive rehabilitation of a severely eroded dentition: the
three-step technique. Part 2. Eur J Esthet Dent 3, 128-146.
Vailati, F., Belser, U.C., 2008c. Full-mouth adhesive rehabilitation of a severely eroded dentition: the
three-step technique. Part 3. Eur J Esthet Dent 3, 236-257.
Vouvoudi, E.C., Sideridou, I.D., 2013. Effect of food/oral-simulating liquids on dynamic mechanical
thermal properties of dental nanohybrid light-cured resin composites. Dent Mater 29, 842-850.
Wong, A.C.H., Tian, T., Tsoi, J.K.H., Burrow, M.F., Matinlinna, J.P., 2017. Aspects of adhesion tests
on resin-glass ceramic bonding. Dent Mater 33, 1045-1055.
Yazigi, C., Franzo, A., Marchesi, G., Yazigi, C., Schneider, H., Chaar, M.S., Ruger, C., Haak, R.,
Kern, M., 2018. Effects of artificial aging and progression of cracks on thin occlusal veneers using
SD-OCT. J Esthet Restor Dent 88, 231-237.
Yazigi, C., Kern, M., Chaar, M.S., 2017. Influence of various bonding techniques on the fracture
strength of thin CAD/CAM-fabricated occlusal glass-ceramic veneers. J Mech Behav Biomed Mater
75, 504-511.

39
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.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

You might also like