You are on page 1of 68

Survival rate of bonded composite

veneer versus ceramic veneer


restorations: A systematic review

Dr. Neethu Rakesh

A dissertation submitted in partial fulfilment


for the degree of

Master of Dentistry
with endorsement in Aesthetic Dentistry (MDent)

University of Otago

Dunedin, New Zealand

August 2019

i
Principal Investigator: Neethu Rakesh
Programme: Master of Dentistry with endorsement in Aesthetic Dentistry

University of Otago
Dunedin
New Zealand
rakne071@student.otago.ac.nz

Supervisors

Dr. Vincent Bennani


Associate Professor
Department of Oral Rehabilitation, University of Otago

Mr. John Aarts


Senior Lecturer
Department of Oral Rehabilitation, University of Otago

Professor Paul Brunton


Pro-Vice-Chancellor
Division of Health Sciences, University of Otago

ii
Abstract

Background: Developments in veneer materials offer clinicians multiple options for creating
aesthetic and functional laminate veneers. Despite the obvious advantage of the minimum
invasive aspect of veneer restorations, the best material to use is still debatable.

Aim: To determine the survival rate of composite veneer versus ceramic veneer restorations.

Methodology: The ‘Preferred reporting items for systematic review and meta-analysis
protocols’ (PRISMA-P) was used for this systematic review. The articles were searched in the
following electronic databases of Medline, PubMed, Web of Science and Scopus from 2004 to
present. All studies involving survival rate of composite veneers and ceramic veneers between
January 2004 to January 2019 were included. The articles were then scored using the GRADE
system (British Medical Journal, clinical evidence for assessing the quality of studies) allowing
a holistic assessment.

Results: A total of 808 full publications as being potentially applicable to the review was
identified. After applying the inclusion and exclusion criteria, a shortlist of 32 articles for full
text review was identified. After reading abstracts and full screening of articles only 19 articles
were found to meet the review’s inclusion criteria. Within the included articles, 3 studies
provided survival rates of composite veneers, 14 studies provided survival rates of ceramic
veneers and 2 studies compared both composite and ceramic veneer survival rate. Due to the
heterogeneity of the study designs, interventions, and survival assessment methods, no meta-
analysis was performed.

Conclusion: Within the limitations of this research, ceramic veneers performed significantly
better compared to composite laminate veneers.

iii
Acknowledgements

It would be absolutely perfect to start by thanking my thesis advisor Associate Professor Dr.
Vincent Bennani who has been a strong driving force for this project right from the word Go.
The door to his office was always open whenever I ran into trouble or had a question about my
research. He consistently allowed this paper to be my own work but steered me in the right
direction whenever he thought I needed it. I would also like to thank the expert who was
involved in scrutinizing this research project, Senior Lecturer Mr. John Aarts, whose timely
interventions and advice have gone a long way in making this thesis a reality. I would like to
express my gratitude to Professor Paul Brunton, for giving me this opportunity to do this
research in the Department of Aesthetic Dentistry, University of Otago. Despite his busy
schedule he made himself available to me when I had doubts regarding my study.

I have also received generous support from Ms. Trish Leishman regarding access to the various
scientific databases, which was critical to this project. Ms Sandi Jull and Mr David Jull have
been a tremendous help throughout this endeavour through their technical know-how. I greatly
appreciate the assistance I received from my colleague Dr. Shweta Gautam. She has always
been there for me as a bouncing pad for ideas and a constant source of helpful advice.

Finally, I must express my very profound gratitude to my parents Mr. Manohar and Mrs Sudha
Manohar, my sister Ms. Nithya Manohar, my partner Dr. Rakesh Mohan and my son Rachit
Rakesh for providing me with unfailing support and continuous encouragement throughout my
years of study and through the process of researching and writing this thesis. This
accomplishment would not have been possible without them. Thank you.

iv
Table of Contents

Abstract ..................................................................................................................................... iii


Acknowledgements ................................................................................................................... iv
List of Tables .......................................................................................................................... viii
List of Figures ......................................................................................................................... viii

1. Introduction ............................................................................................................................ 1

2. Background ............................................................................................................................. 2

3. Veneer Material ...................................................................................................................... 4


3.1 Feldspathic Porcelain .................................................................................................. 4
3.2 Leucite-Reinforced Glass Ceramic ............................................................................. 5
3.3 Lithium Disilicate ....................................................................................................... 5
3.4 Zirconium-based Veneers ........................................................................................... 6
3.5 Composite Resin Material .......................................................................................... 7
3.5.1 Microfilled composites....................................................................................................... 7
3.5.2 Nanofilled composites........................................................................................................ 7

4. Veneer Preparation Design ..................................................................................................... 9


4.1 Incisal Preparation Designs ........................................................................................ 9
4.2 Nonoverlap Incisal Preparation Design ...................................................................... 9
4.2.1 Window Preparation Design .............................................................................................. 9
4.2.2 Feathered Edge Preparation ............................................................................................... 9
4.3 Overlap Incisal Preparation Design .......................................................................... 10
4.3.1 Palatal Chamfer Preparation............................................................................................. 10
4.3.2 Butt Joint Preparation....................................................................................................... 10
4.4 Veneers on Vital and Non-Vital Teeth ..................................................................... 11
4.5 Curing system ........................................................................................................... 12

5. Challenges of The Veneers ................................................................................................... 13


5.1 Direct and indirect Composite veneer failure ........................................................... 13
5.1.1 Debonding of composite veneers ..................................................................................... 13
5.1.2 Microleakage .................................................................................................................... 13
5.1.3 Aging of composite veneers ............................................................................................. 14
5.1.4 Wear of composite veneers .............................................................................................. 14
v
5.2 Ceramic veneer failures ............................................................................................ 15
5.2.1 Veneer Fracture ................................................................................................................ 15
5.2.2 Aging of luting agent ....................................................................................................... 15
5.2.3 Veneer debonding ............................................................................................................ 16
5.2.4 Veneer cracks ................................................................................................................... 16
5.2.5 Postoperative Sensitivity .................................................................................................. 17
5.3 Marginal Adaptation ................................................................................................. 17

6. Aim ....................................................................................................................................... 18
6.1 Objective ................................................................................................................... 18
6.2 Hypothesis ................................................................................................................ 18

7. Material and Methods ........................................................................................................... 19


7.1 Study Design............................................................................................................. 19
7.2 Data Collection ......................................................................................................... 19
7.3 Data Sampling .......................................................................................................... 20
7.4 Keywords .................................................................................................................. 20

8. Grading of the Studies .......................................................................................................... 22


8.1 Survival Rate Analysis ............................................................................................. 23

9. Results .................................................................................................................................. 25
9.1 Study Characteristics ................................................................................................ 25
9.2 Veneer overall survival rate ...................................................................................... 29
9.3 Absolute Veneer Failures ......................................................................................... 32
9.4 Relative Veneer Failure ............................................................................................ 33
9.5 Veneer luting agents ................................................................................................. 34
9.5.1 Composite veener luting agents ....................................................................................... 34
9.5.2 Ceramic veneer luting agent............................................................................................. 35

10. Discussion........................................................................................................................... 36
10.1 Survival rate of composite veneers....................................................................... 36
10.1.1 Direct versus indirect composite veneer……………………………………………...37

10.2 Survival rate of ceramic veneers .......................................................................... 38


10.2.1 Feldspathic veneers…………………………………………………………………..38
10.2.2 Leucite reinforced glass ceramic veneers……………………………………………39

vi
10.3 Relative failure of composite and ceramic veneers .............................................. 40
10.3.1 Surface roughness……………………..…………………………………………..…41
10.3.2 Colour match………………………..………………………………………………..41
10.3.3 Marginal discolouration and wear of the restoration……………………………...…41

10.4 Effect of bonding agent on ceramic veneer failure .............................................. 42


10.4.1 Effect of parafunctional activity on ceramic veneer failure……...…………...43
10.4.2 Effect of substrate on ceramic veneer failure ……………………...…………43
10.5 Composite veneers versus ceramic veneers ......................................................... 44

11. Conclusions ........................................................................................................................ 45

12. Recommendations .............................................................................................................. 46

13. Clinical Significance........................................................................................................... 46

References ................................................................................................................................ 47

vii
List of Tables

Table 8.1: The GRADE scoring system used for reviews ........................................................ 22
Table 9.1: Grading of the reviewed articles ............................................................................. 25
Table 9.2: Description of studies and types of failure for composite veneers.......................... 27
Table 9.3: Description of studies and types of failure for ceramic veneer studies ................... 29
Table 9.4: Survival rates for studies that included ceramic veneers ........................................ 33

List of Figures

Figure 7.1: Flow chart illustrating the data selection process and number of studies that
were sourced from the various databases ................................................................................. 22

viii
Introduction

More than one-third of the adult population is dissatisfied with the colour or shape of their
natural teeth. Therefore, adhesive technologies make it possible to conserve tooth structure
while satisfying patients aesthetic desires (Pini et al., 2012). Improvements in technologies have
contributed in treating unaesthetic teeth with veneers (Christensen, 2008). In the early 20th
century, the veneer restoration technique was developed in the United States of America and
was quickly adopted world-wide. However, bonding veneers to natural teeth proved initially
difficult for dental practitioners but with the improvement of restorative dentistry, veneer
restorations are now a reliable option (Chen et al., 2005). The continued development of veneer
materials offered clinicians with a variety of options for creating aesthetic and functional
laminate veneers (Pini et al., 2012).

Veneer restorations bonded to enamel require minimal tooth reduction. The tooth surface is
etched with phosphoric acid, and then a thin a layer of composite or ceramic material is bonded
on the buccal surface to improve the aesthetic outcome (Burke, 2012). For the composite
veneers the material could be directly build free hand in the patient’s mouth it is known as the
direct technic or the composite veneer restorations could be first fabricated in the dental
laboratory and then bonded in situ, it is known as the indirect technic.

Direct composite veneers are becoming popular because of its quick, inexpensive and easy to
repair qualities as compared to the ceramic veneers (M. M. Gresnigt et al., 2012). Furthermore,
the improvement of adhesive systems and resin composite materials has drastically increased
their uses in dental practices over the last two decades (De Munck et al., 2005). The composite
materials used originally produced a dull restoration, however, recently developed techniques
produce aesthetically suitable composite veneers. Nevertheless, the survival rates of these
restorations are still in debate with problems such as discolouration, wear, and fractures (Bagis
et al., 2008).

The clinical success of these technics depends on technical aspects, aesthetic results and
performance over time, but the benefit for the patient in terms of function, comfort and survival
time is still debatable (Wakiaga et al., 2004). Thus, the purpose of this review is to discern the
survival rate of composite and ceramic veneers.

1
Background

Veneers restorations were designed by California dentist Charles Pineus in 1928. Dr Pineus
used veneers for a film shoot – temporarily changing actors’ teeth (Pincus, 1937). Denture
adhesive was used to retain the veneers temporarily in place. In 1983, Simonson and Calamia
(1983) revealed that porcelain could be etched by hydrofluoric acid and then bonded
permanently to the surface of teeth. This major breakthrough opened a new field in veneer
restoration.

Despite the obvious advantage of the minimum invasive aspect of veneer restorations, the
benefit for the patient in terms of function, comfort, and survival rates is still debatable. This is
further complicated due to the uncertainty about the best choice of material (composite versus
ceramic) (Korkut et al., 2013; Wakiaga et al., 2004).

Composite veneer restorations present microleakage problems leading to discolouration,


postoperative sensitivity, caries and pulpal inflammation. The major cause for microleakage is
poor adaptation between the composite veneer and tooth structure (Priyalakshmi & Ranjan,
2014). Walls et al.,(1988) conducted a two year trial on 320 composite laminate veneers for
discolouration or hypoplasia. Of these 273 veneers were followed over a period of two years.
However, the veneers reported chipping (52%) and marginal discolouration (72%). Therefore,
there was a greater frequency of failure which was associated with marginal staining and
chipping.

Ceramic veneer restorations also presents a number of problems. Dunne et al.,(1993) conducted
a longitudinal study on 315 ceramic veneers on 96 patients that were evaluated up to 5 years 3
months. During the evaluation period there was a failure of 59 restorations in which 25 were
minor and 34 veneers showed debonding. Similarly, another retrospective study done on 191
veneers with an observational time period of one to ten years which resulted in a failure of 4%
with a survival probability of 91% over a period of 10 years (Dumfahrt & Schäffer, 2000).

A systemic review conducted by Wakiaga et al.,(2004) included a 2.5 years interim analysis on
112 participants requiring 1 and 6 veneer restorations for anterior teeth. Follow up was done
approximately 2 years later. The participants were divided into three groups constituting of a
total number of 263 restorations; Group A was composed of restoration using direct resin
composite (69 restorations), Group B was restoration using indirect resin composite (54
restorations), and Group C contained ceramic restoration (56 restorations). The outcome of this
study reported in terms of absolute failure, relative failure, and survival of original restorations.

2
The authors of the systematic review reported that the evidence was not reliable enough to show
benefit of one type of restoration over the other, and hence recommended more randomised
control studies to examine the effectiveness of veneer restorations. Therefore, a strong
evidence-based review comparing the survival rate of composite veneer versus ceramic veneer
restorations would be very helpful for the dental community.

3
Veneer Material

The most important factor for the clinical success of veneers is the proper selection of the
material used (Sadaqah, 2014). Composite resin and ceramic materials are the materials
available for the anterior aesthetic treatment (Ali et al., 2015). Composite resins are used as
veneer material to mask discolouration and to correct the unaesthetic tooth shape and/ or
position (Korkut et al., 2013).They have been continuously progressing ever since Bis-GMA
was introduced by Bowen in 1962. Recent technologies have improved the composite material
and their physical properties. Therefore, dental composite material can be divided into direct
and indirect resin composite (IRC). IRCs are also known as prosthetic composite and laboratory
composites (Nandini, 2010).

Ceramic materials are commonly used for morphological modifications related to tooth shape,
size, and colour (Radz, 2011). There are several ceramic materials available for veneers
including feldspathic porcelain, glass-based ceramic and zirconia-based material (Pini et al.,
2012). The success of veneer restorations depends on numerous factors, but the appropriate
selection of the most appropriate veneering material for the case is key.

3.1 Feldspathic Porcelain


Feldspathic porcelain was the most common material used for ceramic veneers in the 1980s
because of its translucency and the ‘contact lens effect’, which is described as the relative
thickness of the veneer margin (approximately 0.1mm thick) that, creates an optical blend
between the tooth and the veneer margin. This effect is similar to perception of the human eye
with a clear contact lens (Friedman, 1998; Simonsen & Calamia, 1983). Feldspathic porcelain
material is inherently brittle and friable, however it can be used to restore tooth shape with good
control over the marginal accuracy, emergence profile and translucency (Fleming et al., 2006).
Due to its minimal thickness requirement, this type of ceramic material is useful for the aesthetic
veneer. Feldspathic porcelain has a high degree of translucency and an outstanding aesthetic
appearance that can optically approximate natural teeth and it enables attainment of thicknesses
of less than 0.5mm. This enables a slight reduction of only 0.3mm to 0.5mm of the tooth surface
to be achieved that preserves the health of the gingival tissues and limits over contouring
(Magne & Belser, 2004). The limitation of this material is its low flexural strength prior to
cementation (Spazzin et al., 2016), and the required investment of time and effort in order to
produce accurately fitting restorations (Taskonak et al., 2004). Furthermore, feldspathic
porcelain veneers are manufactured using liquid sculpting powder by a ceramist, therefore the

4
aesthetic properties of such veneers is reliant upon the ceramist ability to build the anatomy,
colour and translucency of the veneers (McLaren & LeSage, 2011).

There are several clinical studies that support feldspathic porcelain for laminate veneers as their
survival rate has been recorded to be more than 90% over 10 years of clinical service. Layton
and Walton (2007) conducted a clinical study on 155 patients with 499 veneers over a period
of 21 years to evaluate the survival rate of feldspathic porcelain. The results showed that
feldspathic porcelain veneers have an excellent survival rate.

3.2 Leucite-Reinforced Glass Ceramic


The strength of glass ceramic is achieved by suitable micro-sized crystals that are uniformly
disseminated through the glass, such as leucite reinforced and lithium disilicate ceramic (Guess
et al., 2011). They are indicated for veneers due to its optical properties (Conrad et al., 2007).

Leucite crystals present in leucite reinforced glass ceramic (IPS Empress) make up 50-55% of
the material, therefore the refractive index is closer to feldspathic ceramic (Kelly & Benetti,
2011). Numerous studies have firmly established that leucite reinforced glass ceramic (IPS
Empress) fulfils the high standard demands of aesthetic restoration, such as veneers, delivering
colour, translucency, fluorescence and opalescence similar to natural teeth (Giordano &
McLaren, 2010). Needle-like morphology of apatite is unusual in glass-based ceramic (Reich
et al., 2004), however, later, leucite apatite and multiple apatite glass-ceramic developed
needle-like morphology (Attia & Kern, 2004). In addition, IPS Pro CAD, which is a leucite
reinforced ceramic similar to IPS Empress, was introduced in 1998 to be used in CEREC in lab
system. Fradeani (1998) conducted a study report IPS empress laminate veneers over the period
of 6 years. In this study, a total of 83 anterior empress veneers were placed in private practice.
Colour, marginal discolouration, recurrent caries, contour and marginal integrity were
evaluated. There was only one failure with a success rate of 98.8%.

3.3 Lithium Disilicate


Lithium disilicate material was introduced in the 1990s (Kelly & Benetti, 2011). Most veneers
are made up of lithium disilicate material; when compared to other materials, they have greater
biaxial strength and fracture toughness (Gonzaga et al., 2011). Lithium disilicate is processed
either as pressed ceramic or as lithium metasilicate. During firing, there is a complete
crystallization of lithium meta-silicate that is related to the final strength of the veneers
(Schmitter, 2012). Lithium disilicate (IPS Empress 2) in addition to its translucency also
exhibits good chemical durability. In comparison with IPS Empress, the mechanical properties
5
of the IPS Empress 2 framework material were enhanced. The results of this was a higher degree
of crystallinity. The framework material of lithium disilicate glass-ceramic satisfies the
chemical longevity of the veneer materials (Anusavice, 1996). In addition, IPS Emax press, a
lithium disilicate glass-ceramic introduced in 2005 is the most preferred material for porcelain
laminate veneers in recent years (Turgut & Bagis, 2011). They are translucent with different
crystalline structures; when compared to other materials such as IPS empress, they have a
different crystalline volume and reactive index, with the IPS Emax having more translucency
(Heffernan et al., 2002). However, this advantage of the IPS Emax becomes a disadvantage if
the discolouration of the underlying cement affects the final veneer (Karaagaclioglu & Yilmaz,
2008).

3.4 Zirconium-based Veneers


Zirconium-based ceramic veneers are polycrystalline material with no glass content present.
Since the atoms are placed in irregular fashion, propagation of cracks is more difficult than
compared to glass-based ceramic. Therefore, they are much harder and stronger than glass-
based ceramic (Kelly, 2008).

Yttrium oxide stabilized zirconia due to its high fracture toughness and mechanical strength is
also a material of choice in restorative dentistry (Sundh et al., 2005). They can be manufactured
via two methods. In the first method zirconium can be milled, the restorations will have a linear
shrinkage of about 20-25% during sintering. Therefore, using a pre-sintered blank can decrease
the wear on milling. In the second method, fabrication can be done using CAD/CAM technique
using a presintered blank. However, the disadvantage of this would be compromise in the
microstructure and strength of the veneers (Luthardt et al., 2002). A method that mills
zirconium core in the softer “green” state and then sinter it, is higher to one that mills the core
in the sintered state. This is because it requires a high temperature milling and robust milling
machine that will significantly shorten the life span of the restoration (Rekow & Thompson,
2005). Therefore, milling in green state followed by sintering decreases the temperature and
heals the milling induced defects (Donovan, 2008).

The advantage of zirconium veneers over feldspathic veneers and glass veneers is they offer a
opaque nature in masking unpleasant colour of the tooth. The opaque nature of zirconium is
due to its elemental chemistry, high crystallinity and density (Heffernan et al., 2002). If the
zirconium veneers are thin, then the final colour and translucency of the restoration may be of
concern. Hence, colour of the luting cement and the region of the veneer will be important
(Baldissara et al., 2010). Unfortunately, bonding zirconium veneers with resin cements is a

6
major drawback as it cannot be etched by hydrofluoric acid and therefore will produce poor
micromechanical retention to dentin (Thompson et al., 2011). Furthermore, zirconium core
interface is one of the drawbacks of the restoration due to ceramic chipping or cracking (M.
Aboushelib et al., 2007). The factors that influence veneer cracking is firing shrinkage of
ceramic, flaws on veneering, poor wetting by veneering on core, thermal expansion co-efficient
between the core and ceramic (Aboushelib et al., 2006).

However, in clinical situations where there is a higher tensile and shear stress, zirconium-based
ceramic should be considered (Font et al., 2006).

3.5 Composite Resin Material


The introduction of light-cured composite resin materials and their advancement has led to the
development of several conservative aesthetic techniques for the correction of defects.
Composite resin can be used as directly or indirect fabricated veneers. Composite resin,
however, have inherent disadvantages; including discolouration, excessive wear, and
polymerization shrinkage. The mechanical properties of the composite resin is subject to its
curing, which is achieved using the visible spectrum light combination with heat. The
mechanical properties of the composite resin also depend on the particle size and concentration
of fillers. The amount of filler present increases the elastic modulus, flexural strength, hardness
and compressive strength but decreases the polymerization shrinkage (Ikejima et al., 2003). The
size of the particle is 8-30um in hybrid composites, 0.7-3.6um in microfilled and nanofilled
composites were between 5-100 nanometers (Moszner & Klapdohr, 2004). All these factors
influence the properties of the composite resin material, such as their hardness, colour stability,
and biocompatibility (Ikejima et al., 2003).

3.5.1 Microfilled composites


Microfilled composites are used in veneer composite materials such as the opaque resin, dentine
resin, enamel resin, and transparent resin. Microfilled composites produce a highly polished
abrasion resistance surface and have a higher wear resistance compared to unfilled resin. In
addition, colour stability is of paramount importance in veneer restorations, which is the main
advantage of microfilled composites (Welbury, 1991).

3.5.2 Nanofilled composites

Nanofilled composites has significantly further improved the clinical performance of


composites in terms of wear resistance, strength and enhanced colour stability. They also show

7
high translucency and polishability which is similar to microfilled composites and wear
resistance equal to hybrid composites (Beun et al., 2007). Due to the wide filler distribution,
decreased dimension and high filler load, there is a reduced polymerization shrinkage and
increase in mechanical properties, such as resistance to fracture, compressive strength and
tensile strength (Moszner & Salz, 2001). The nanofilled particle size is a fraction of the visible
light wavelength spectrum and therefore, are unable to be detected by the human eye, resulting
in increased optical properties (Mitra et al., 2003). Furthermore, there is an increase in the wear
resistance and better retention of the gloss finish (Terry, 2004).

8
Veneer Preparation Design

4.1 Incisal Preparation Designs

According to LeSage (2013), the preparation of ceramic veneers can be classified into buccal
surface preparation (no preparation, minimal preparation, conservative or conventional
preparation), proximal preparation (slice or chamfer margin), cervical definition (chamfer or
knife-edge) and incisal preparation (overlap or non-overlap). Incisal preparation is classified
into two categories: overlap and non-overlap. The designs are described as the window
preparation or intra-enamel, the feathered edge, the palatal chamfer or overlapped, and the butt
joint or incisal level (Clyde, 1988). The non-overlap category has the window and feathered
edge preparation designs, the overlap category includes the butt joint and the palatal chamfer
(Walls et al., 2002).

4.2 Nonoverlap Incisal Preparation Design

4.2.1 Window Preparation Design


The window preparation designs result in the suitable thickness of 0.4mm to 0.7mm close to
the incisal edge, reduces the risk of veneer fracture as well as the wear of the antagonist's teeth,
and does not interfere with the incisal guidance (Ben-Amar, 1989). However, this was not
acceptable as there is trouble in masking the ceramic finish lines, moreover, the chipping of the
unsupported enamel posed a high risk (Walls et al., 2002).

A study conducted by Hui and his colleagues in 1991 reported that window preparation designs
showed the least stress compared to the feathered edge and palatal designs. They concluded
that window preparation was a conservative type of veneer and hence, was considered as the
design of choice. Another study by Seymor et al.,(2001) investigated stresses within porcelain
veneers using window preparation and incisal overlap preparation designs. Maximum tensile
and compressive stresses were recorded in the labial marginal region of the porcelain veneers.
They reported that the window preparation showed lower maximum stress at the labial margin
of the porcelain and the labial composite lute for veneers when compared to the incisal overlap
preparation designs.

4.2.2 Feathered Edge Preparation

Feathered edge preparations have been recommended for patients with a normal overbite to
prevent contact of the veneers with their opposing teeth structures (Nomoto et al.,1994). On the

9
other hand, the disadvantages of feathered edge designs include weak veneers as they suffer
from ceramic chipping and difficulties in seating of the veneers (Gilmour & Stone, 1993),
marginal discolouration, and weak marginal adaptations (Calamia, 1989). Porcelain with
feathered edged designs during protrusive guidance are subjected to peel and shear forces
(Walls et al., 2002).

A study conducted by Bergoli and colleagues in 2014 reported that feathered edge design had
a higher fracture strength compared to palatal chamfer preparation designs. They also deducted
that the veneers with feathered edge design generate lower principle stress concentration as
compared to palatal chamfer preparation designs (Bergoli et al., 2014). Smales et al.,(2004)
conducted a follow up clinical study of veneers with feathered edge design that depicted a
survival rate of 80% in situ.

4.3 Overlap Incisal Preparation Design

4.3.1 Palatal Chamfer Preparation

Palatal chamfer preparation is done if there is a thin incisal edge buccolingually or when the
length of the crown is to be increased (Garber, 1993). It provides adequate veneer thickness at
the incisal edge and increases the surface area for bonding by preventing the propagation of
cracks by avoiding sharp angles (Sheets & Taniguchi, 1990). According to Schmidt (2011),
palatal chamfer preparation has a higher failure load than butt joint incisal preparation.

Meta-analysis conducted by Da Costa and colleagues (2013) reported that palatal chamfer
preparation proved to be a risk for veneer fracture. In addition, Zarone and colleagues ( 2006)
conducted fine element analysis that showed that palatal chamfer preparation had the highest
stress tolerance and helps to distribute the stress throughout the preparation.

4.3.2 Butt Joint Preparation


According to Walls et al.,(2002), incisal preparation is advocated for butt joint preparation with
0.5 to 1mm. The advantage of the butt joint is the masking of incisal finish lines, thicker veneers
and strengthening of incisal edge and seating of the veneers. They have low fracture rates as
compared to other designs. The fracture strength did not differ from non-prepared teeth; Butt
joint designs are used in anterior crossbite (class II div 2 malocclusion) (Calamia, 1989). The
survival rate of butt joint design is higher than 92% in studies from 1.5 to 7 years (Smales &
Etemadi, 2004).

10
Castelnuovo et al.,(2000) conducted an in vitro study that evaluated the fracture load and failure
of porcelain laminate veneers. They reported that ceramic veneers with a butt joint and
feathered edge have the highest fracture resistance and remained intact. The study also showed
that butt joint designs offered several advantages compared to palatal chamfer preparation
designs, such as tooth preparation, faciopalatal path of insertion, greater fracture load, low risk
of fracture of veneers and improved aesthetics.

Metanalysis conducted by Da Costa and colleagues (2013) to evaluate the fracture strength of
three different types of tooth preparation designs concluded that butt joint designs have the least
strength of the tooth compared to the palatal chamfer designs. A photoelastic analysis which
compared the butt joint with feathered edge designs reported that butt joint has more favourable
stress distribution compared to feathered edge design (Highton & Caputo, 1987). Magne et al.,
(1999) recommended a butt joint preparation design with a slight bevel. Guess et al.,(2014)
conducted a prospective clinical study that showed a higher survival rate compared to palatal
chamfer designs.

Chai et al.,(2018) conducted a critical review of incisal preparation design on 40 ceramic


veneers. The author reported that butt joint and feathered edge were the two most common
preparation designs provided by clinicians. There was no clear evidence in the review regarding
the best incisal preparation. However, a butt joint design provided decreased ceramic fracture
rates and better aesthetics compared to other preparation types.

4.4 Veneers on Vital and Non-Vital Teeth


Meijering et al.,(1998) conducted a clinical trial on the survival of veneers and they reported a
higher failure risk for veneers on non-vital teeth compared to vital teeth. Veneers are advised
in non-vital teeth mainly because of discolouration, and therefore, a shoulder preparation is
generally indicated, to allow more room for the restorative material. The study concluded a
higher failure rate was observed with the combination of composite resin and non-vital teeth.
Another study done by Çöterta et al.,(2009) compared the survival rates of porcelain laminate
veneers on vital and non-vital teeth using different preparation designs. It was reported that the
clinical failure was higher in porcelain veneers for non-vital teeth.

A retrospective study by Coelho de Souza et al.,(2015) of veneer restorations on vital and non-
vital teeth, reported that survival rates of veneer restoration on non-vital teeth was 50% less
than on vital teeth (Demarco et al., 2012). Furthermore, vital teeth present better clinical
performance in the areas of fracture strength, retention, colour match, and surface lustre

11
(Skupien et al., 2013). In general, bonding to endodontically treated teeth is reduced when
compared to vital teeth. There is also lower resistance to fracture in non-vital teeth when the
pulp is removed, and endodontic therapy is carried out (Cauwels et al., 2014).

4.5 Curing system


The curing of the composite resin is associated with its colour stability, biocompatibility values,
and mechanical properties (Geurtsen & Leyhausen, 2001). Intensity of light is a factor
associated in determining the conversion of monomer into polymer, which is referred to as
‘Degree of conversion’ (Nomoto et al., 1994). The concept “Depth of cure” is also an important
factor. An increment of 2mm thickness at a minimum irradiance of 280 w/cm2 has been
recommended (Musanje & Darvell, 2006). The recommended radiation times, has been
reported to be inadequate for complete polymerization which leads to poor mechanical
properties. Another factor is overheating of the composite due to over exposure which might
result in serious clinical implication (Musanje & Darvell, 2003).

Polymerization shrinkage is due to the conversion of monomer in the composite resin into
polymer network (Peutzfeldt, 1997). This creates contraction and stress in the restoration, which
results in postoperative pain, and poor marginal adaptation (Ferracane & Mitchem, 2003).
Factors that affects the shrinkage are molecular weight and degree of conversion of monomers
(Peutzfeldt, 1997). However, not all the shrinkage is converted to stress because polymer can
absorb/relieve stress (Haralur, 2018).

12
Challenges of The Veneers

5.1 Direct and indirect Composite veneer failure

5.1.1 Debonding of composite veneers


Debonding of composite veneers is due to the degradation of the adhesive system
(Vaidyanathan & Vaidyanathan, 2009). The marginal seal is obtained when the tooth prepared
is surrounded by enamel. However, when there is excessive acid etching, excessive dryness of
the etched surface may occur and lead to denaturation of collagen fibers, which in turn may
also be susceptible to degradation. This leads to nanoleakage and causes failure in adhesive
bond and veneer detachment (Dietschi & Spreafico R.1997). The porosities caused by nanoleakage
are considered as the main pathway for degradation of the adhesive system (Okuda et al., 2001).
It also results in decrease bond strength as well as microstructure breakdown. However, in a
restoration where mechanical retention dominates degradation may not influence the clinical
consequence (De Munck et al., 2005). Lin et al.’s (2005) study on biodegradation of composite
resins reported that salivary enzymes may also cause hydrolysis of the resin. For instance,
salivary enterase may increase the hydrolysis of ester bond in the resin giving rise to the dual
bonding technique, which means application of bonding agent at the end of preparation. This
dual bonding technique can preserve adhesive interface and prevents the veneer detachments
(Woody & Davis, 1992).

5.1.2 Microleakage

Microleakage is related to the invasion of bacteria through the restoration tooth margin which
leads to marginal staining, post-operative sensitivity, secondary caries and pulpal damage. The
main reason for microleakage is poor adaptation between the restorative material and tooth
structure. In addition, volume changes which occurs by thermal factors can lead to
microleakage and gap formation (Shih, 2016). The factors that affect microleakage are
malocclusion, non-retentive tooth preparation, excessive forces of mastication and
polymerization (Bhandari et al., 2012). In composite material microleakage is due to
polymerization shrinkage that leads to tensile stresses between the restoration and the tooth
surface. This resulting disassociation of the restoration from the tooth leads to marginal gap
which enables the penetration of the bacteria (Vaidyanathan & Vaidyanathan, 2009).
Furthermore, polymerization shrinkage or removal of luting cements causes poor marginal
integrity. In addition, the mechanical properties and type of luting cements also has an effect
on microleakage (Dukic et al., 2010). Another risk factor is the thickness of luting cements,

13
thicker luting cement can cause leakage due to propping and subsequently ill-fitting veneers
(Magne et al., 1999).

Aschenbrenner et al.,(2012) conducted a microscopic analysis of marginal adaptation and


sealing to enamel and dentin using four self-adhesive resin cements and reported that small
marginal cement deterioration were detected. Another study by Gerdolle et al.,(2005) on
microleakage of indirect composite restoration with four luting composite restoration and with
four luting cements reported that there was a deterioration of bond between tooth and the luting
agent.

Tooth preparation of composite veneers is another important factor for microleakage. Pini et
al.,(2012) suggested that the preparation designs should

enable favourable marginal adaptation. In bonded restoration, leakage is more common to


enamel margin than dentin margin because resin-based material bonds to etched enamel (Erkut
et al., 2013). A study conducted by Hekimoglu et al.,(2004) reported there was microleakage
in the cervical margin rather than incisal margin due to exposure of dentin and this may increase
the risk for marginal seal.

5.1.3 Aging of composite veneers

The colour of composite veneers is affected by external factors such as smoking and dietary
habits. In addition, factors such as UV radiation, water, heat and food colourants also influence
the colour stability (Bagheri et al., 2005). Colour stability of composites is also influenced by
the duration and intensity of polymerization and the degree of conversion. The filler type, size
and the matrix composition affect the colour stability (Janda et al., 2005). Ferracane et al.,
(1985) suggested that oxide process of unreacted double carbon bond present in polymerized
resin gets accelerated by heat or UV light which leads to the yellowish coloured peroxide.

Rosentritt et al.,(1998) conducted a clinical study on the colour stability on resin composite
veneers and they reported that the material became yellow and darker after 18 months of clinical
performance. Similarly, another study conducted by Matsumura et al.,(2000) on 110 composite
veneers reported that there was only a colour stability of 67.3% after 6 years.

5.1.4 Wear of composite veneers

The wear of composite resin is a combination of chemical disintegration, abrasion, adhesion


and fatigue (Lutz et al.,1983). The initial wear of composite resin is most likely due to
inadequately finished veneers, softening due to chemical degradation and also the incomplete

14
conversion of composite resin due to the oxygen inhibition which results in reduced surface
hardness (Montes-G & Draughn, 1986). Composite resin fillers hardness and size influence the
wear of the veneers. When opposed to natural teeth the filler hardness of the composite should
be equal to the hydroxyapatite (Burgoyne et al., 1991).

Wendt and Leinfelder (1990) demonstrated that when composite resin undergoes both heat and
light curing, they show an increase in mechanical properties such as wear resistance,
microhardness and physical properties such as brittleness, elasticity module, solubility, thermal
expansion co-efficient when compared to only light cured composite.

5.2 Ceramic veneer failures

5.2.1 Veneer Fracture

The main reason for the failure of veneers is a fracture. Fracture and cracks are caused mainly
by the occlusal forces on the restoration (Heymamn et al., 1991). Feldspathic porcelain with
lower mechanical resistance has a higher risk for fracture than high resistance feldspathic
porcelain (Giordano et al., 1995). It is observed that fracture rates are high in feldspathic
porcelain, lithium disilicate, and leucite reinforced glass-ceramics. Fradeani et al.,(2005)
reported 5.6% veneer fracture after 12 years. Similarly, Peumans et al.,(2004) reported a
fracture rate of 4% after 5 years. Parafunctional activity and masticatory load on a veneer
increase its fracture rate by eight times. It has be shown that even after specific guidelines are
followed to establish the restoration, parafunctional habits can continue, and therefore, occlusal
splints are advised in these patients (Beier et al., 2012). Fracture also occurs due to a failure in
the adhesive technique used, which may accelerate the hydrolysis (Piemjai & Arksornnukit,
2007). Due to this hydrolysis, the development of gaps between blocks of cement and the veneer
may cause crack propagation within the veneer restoration (Aristidis & Dimitra, 2002).

5.2.2 Aging of luting agent

For the bonding of porcelain laminate veneers, light-cured resin is preferred mainly because of
their colour stability and polymerization (Myers et al., 1994). To attain long-term colour
stability, establishing an optimal polymerization is an important factor (Pires de Souza et al.,
2009). While intensity and the extent of polymerization could be the main cause for
discolouration of veneers. It also includes other aspects such as environmental factors, UV
radiation, and intrinsic factors, mainly the composition of the resin matrix percentage of carbon-
carbon double bond, and type of photo initiator (Papadopoulos et al., 2010). Turgut and Bagis

15
(2011) studied the colour stability of different shades of light-cured resin cement for bonding
IPS Emax press material porcelain veneers of 0.5mm thickness. The results revealed that the
colour of porcelain laminate veneers was altered by the ageing process of the resin cement.

5.2.3 Veneer debonding

Friedman (1998) conducted a clinical study on 3500 ceramic veneers over the period of 15
years and the author reported two third of the failure were associated with fracture and
debonding. Adhesive cementation is an important factor for the longevity of ceramic veneers.
Nevertheless, a lasting bond not only depends on the composite resin but also the bond interface
that is the dental substrates (Piemjai & Arksornnukit, 2007). However, mechanical interlocking
is more stable in the case of ceramic veneers where the tooth preparation is confined to the
enamel. The margins have little dentin exposure, and this limits flexion of the tooth which may
be related to the debonding of the veneers (Aykor & Ozel, 2009). Bond failures may also be
due to incomplete polymerization of the adhesive cement which increases the hydrolysis in the
short term (Piemjai & Arksornnukit, 2007).

5.2.4 Veneer cracks


Cracks are mainly formed in the palatal area and in the labial aspects of the veneers. Despite
the improvement of ceramic material, brittleness, low ductility and cracks remain as limitations
(Lawn et al., 2004). Many factors contribute to veneer cracking, such as, the sintering process
causes structural flaws and residual stresses in the veneers which leads to crack formation and
propagation (Anusavice, 1996). In addition, polymerization shrinkage of resin creates stress at
the ceramic subsurface (Peumans et al., 1999). Furthermore, exposure to temperature between
0 to 67c from food intake leads to thermal expansion of luting composites, which weakens the
bond quality between the resin and ceramic veneers (Örtengren et al., 2001).

The occurrence of cracking in veneers is well documented. Peumans et al.,(2004) conducted


clinical studies on feldspathic veneers over a period of 5 years and 10 years and they reported
that there was an increase in the visible crack palatally and buccally due to masticatory load.
Another study conducted by Dumfahrt and Schaffer (2000) on felspathic veneers over 10 years
reported cracks. In the study of Magne et al.,(2000) after 4.5 years, 12% of veneer restorations
were reported to develop cracks.

16
5.2.5 Postoperative Sensitivity
Postoperative sensitivity is the major problem faced by patients after the placement of the
restoration, even if a dentin liner is used (Wendt & Leinfelder, 1990). It mainly depends on the
technique used in the restoration rather than the adhesive used. It is caused by several factors
including the etching of the dentin and the penetration of bacteria into the pulpal canal, occlusal
variation, and distortion of the tooth (Bryant & Mahler, 1986). Tooth preparation of the veneers
deep into the dentin can lead to post-operative sensitivity and pulpal death (Christensen, 2008).
During the total-etch technique on a deeply cut dentin, there is a significant chance that the
dentinal canals will not be plugged adequately and cause pulpal irritation and post-operative
sensitivity (Casselli & Martins, 2006).

5.3 Marginal Adaptation


Marginal discrepancies are caused as a result of ageing of adhesive luting agent that leads to
polymerization shrinkage, microleakage, different co-efficient of thermal expansion of the
bonding material and composite washout (McKinney & Wu, 1985). Peumans et al.,(2004)
reported 36% of detectable margin of veneers over an observational period of 10 years.
However, better marginal adaptation of 97.6% and 92% for IPS Empress was observed in
Fradeani et al.’s (2005) and Probster et al.’s (1999) studies respectively.

Magne et al.,(2000) conducted a clinical study on extensive veneers and reported higher palatal
marginal discrepancies, which was due to masticatory loading in the palatal area. Specially, in
case of overlap veneers, marginal discrepancies were associated with increase in marginal
discolouration (Peumans et al., 2004). Another study by Dumfahrt & Schaffer et al.,(2000)
reported that the increase in marginal discolouration of the veneers was due to dual-
polymerizing luting composite, exogenous deposits, marginal dentin exposure and the viscosity
of luting composites.

17
Aim

To determine the survival rate of bonded composite veneers versus ceramic veneer
restorations.

6.1 Objective

To examine the survival rate of bonded composite veneers versus ceramic veneer restorations.

6.2 Hypothesis

HO: There is no difference in the survival rate of bonded composite veneers and ceramic veneers

HA: There is a difference in the survival rate of bonded composite veneers and ceramic veneers.

18
Material and Methods

7.1 Study Design

This study is a systematic review of the literature published between January 2004 and January
2019 on the survival rate of bonded composite versus ceramic veneers. The guidelines followed
for this systematic review are taken from the ‘Preferred reporting items for systematic reviews
and meta-analysis protocols’ (PRISMA-P) 2015 statement (Moher et al., 2015).

7.2 Data Collection


Extensive search was conducted for all articles in Medline, Scopus, PubMed and Web of
Science. All articles with the keywords mentioned below were selected for further screening.
Two individual investigators separately assessed all the shortlisted articles and screened them
further without any bias. The eligibility of the relevant articles was independently screened for
titles and abstracts to evaluate the articles for full-text reading. When the data in the abstract
was inadequate, titles were used to obtain full texts. Final selection of articles was after mutual
consensus with another reviewer. In case both reviewers had opposing ideas a third reviewer
was consulted for the same. The studies meeting specific inclusion criteria were selected after
second round of screening process.

The inclusion and exclusion criteria are:

Inclusion Criteria:
• Permanent vital anterior teeth suitable for veneer restorations
• Studies that include comparison between direct and indirect techniques for veneer
restorations
• Studies that include data on composite direct and or indirect veneer restorations
• Studies that include data on ceramic veneer restorations
• Studies that include survival rate of veneer restorations
• In vivo studies only

Exclusion Criteria:
• Non-English studies
• Studies that investigated Non vital teeth
• Multiple publications on the same patient cohorts
• Single case report

19
• In vitro studies

7.3 Data Sampling

All studies involving the survival rate of bonded composite veneers and ceramic veneer
restorations from January 2004 to January 2019 were included. The key words used in the
Medline, PubMed, Web of Science, Scopus and are mentioned under Annex 1.

7.4 Keywords
Concept 1: Esthetic dentistry, aesthetic dentistry, dental esthetics, dental aesthetics.

Concept 2: Survival rate, longevity, treatment failure, aesthetic outcome, esthetic outcome,
treatment outcome.

Concept 3: Composite resin, dental ceramic, dental porcelain, direct composite veneers, indirect
composite veneers, dental veneer, dental laminate.

The initial search was conducted in Medline, followed by the other database from which
relevant articles were added. A manual search was done to look for studies that adhered to the
inclusion criteria. The articles which contained any of the exclusion criteria were removed. The
following flow chart describes the distribution of the number of studies that were selected,
among the different databases used (Figure 7.1).

20
Records identified through Additional records identified
IDENTIFICATION
database searching through other sources

(n= 808) (n= 0)

Records after duplicates removed

and screened on title (n= 675)


SCREENING

Records screened Records excluded


(n=503) (n= 283)
ELIGIBILITY

Full-text articles assessed for Full-text articles excluded with


eligibility (n=220) reasons (n= 188)

Studies included in qualitative


synthesis (n=32 )
INCLUDED

Studies included in quantitative


synthesis (n=19 )

21
Figure 7.1: Flow chart illustrating the data selection process and number of studies that were
sourced from the various databases

Grading of the Studies

The quality of the studies included in the review were scored using the GRADE system (British
Medical Journal, clinical evidence for assessing the quality of studies) allowing a holistic
assessment (Guyatt et al., 2008). The factors taken into consideration in this scoring system are:
type of evidence, quality, consistency, directness and effect size. The final GRADE score was
determined by the sum of the individual scores in the aforementioned categories and
summarised as: high (at least 4 points overall), moderate (3 points), low (2 points), and very
low (1 point or less).

22
8.1 Survival Rate Analysis
Survival analysis was established on two types of failure ‘absolute failure’ and ‘relative failure’.
“An absolute failure was defined as loss, fracture or removal of the restoration requiring a new
restoration.” “A relative failure was defined as fracture, chipping, dislodgement or shade
correction of the veneer restoration allowing repair or correction with a direct resin composite
material or rebonding without the need for new veneer restorations or other
restoration”(Meijering et al., 1998). Most articles have considered only absolute failure as
failures, with only a few articles considering both absolute and relative failure, as failures.

23
24
Results

After removal of duplicates a total of 808 unique studies were identified.

After screening of all the titles and abstracts and apling the exclusion and inclusion criterias a
shortlist of 32 articles for full text review was identified. Following this, a total of 19 individual
studies were selected to be a part of the final review and graded using the GRADE score system.
The results of the grading system are depicted in Table 9.1.

Table 9.1: Grading of the reviewed articles


Author Type of Quality Consistency Directness GRADE Rank
evidence score
Gresnigt et al., 2013 a 4 0 0 0 4 High
Gresnigt et al., 2013 b 4 -1 +1 0 4 High
Gresnigt et al., 2019 4 -1 +1 0 4 High
Rapani et al., 2015 2 0 +1 0 3 Moderate
Guess et al., 2008 2 0 +1 0 3 Moderate
Coelho de Souza et al., 2015 2 -0 +1 0 3 Moderate
Layton and Walton, 2007 2 0 +1 0 3 Moderate
Rinke et al., 2013 2 0 +1 0 3 Moderate
Fradeani et al., 2005 2 0 +1 0 3 Moderate
D’Arcangelo et al., 2012 2 0 +1 0 3 Moderate
Gurel et al., 2012 2 0 +1 0 3 Moderate
Beier et al., 2012 2 0 +1 0 3 Moderate
Rinke et al., 2018 2 0 +1 0 3 Moderate
Gresnigt et al., 2012 2 0 +1 0 3 Moderate
Arif et al., 2019 2 -1 +1 0 2 Low
Chen et al., 2005 2 0 0 0 2 Low
Öztürk et al., 2014 2 -1 +1 0 2 Low
Ruiz et al., 2010 2 -1 +1 0 2 Low
Guess et al., 2014 2 -1 +1 0 2 Low

9.1 Study Characteristics


A total of 19 studies on veneers was identified, in which there were 3 studies on composite
veneers and 14 studies on ceramic veneers, with only 2 studies reporting on both composite and
ceramic veneers. All the studies in the systematic review were published between January 2004
to January 2019. The clinical information of the veneers is summarized in the Tables 9.4 and
9.5.

Out of the 3 composite veneer studies, 1 study reported on indirect composite veneers (Rapani
et al., 2015). The prefabricated veneers used in the study was Edelweiss composite veneers.

25
The 2 other studies reported on direct composite veneers and these studies used microfilled and
micro-hybrid composites (Coelho de Souza et al., 2015; Gresnigt et al., 2012).

Out of 14 studies that included ceramic veneers, 3 studies reported on ceramic veneers made
out of feldspathic porcelain (Fradeani et al., 2005; Gresnigt et al., 2013a; Layton & Walton
(2007). Six studies reported on IPS Empress Esthetic, Ivoclar (Arif et al., 2019; Fradeani et al.,
2005; Guess et al., 2014; Guess & Stappert, 2008; Gurel et al., 2012; Ruiz et al., 2010). With
the exception of 3 studies which only mentioned glass-ceramic material (Beier et al., 2012;
Rinke et al., 2013; Rinke et al., 2018). Three other studies did not report the type of ceramic
material used for the veneer restorations (Chen et al., 2005; D’arcangelo et al., 2012; Öztürk &
Bolay, 2014).

Most of the studies have not mentioned the type of preparation design used in the fabrication
of ceramic veneer (Arif et al., 2019; Beier et al., 2012; Coelho de Souza et al., 2015; Guess &
Stappert, 2008; Gurel et al., 2012; Öztürk & Bolay, 2014; Ruiz et al., 2010). One study reported
non-overlap preparation designs (Chen et al., 2005). Four of the included studies used butt joint
preparation (D’arcangelo et al., 2012; Fradeani et al., 2005; Gresnigt et al., 2013a; Guess et al.,
2014). Palatal overlap designs were used in 3 studies (Fradeani et al., 2005; Rinke et al., 2013;
Rinke et al., 2018). One study included both butt and palatal overlap designs (M Fradeani et
al., 2005).

In the composite veneer studies, the veneering treatment in Rapani et al.’s (2015) study
involved tooth preparation, veneer adjustment, adhesive procedures followed by cementation,
but no particular preparation designs were mentioned in the study. Similarly, no data on tooth
preparation was mentioned in the direct composite veneer study by Coelho de Souza et al.,
(2015). However, 1 study on direct composite veneers mentioned an incisal overlap preparation
design for the veneer treatment (Gresnigt et al., 2012).

Two randomised clinical studies reported the survival rate of indirect composite veneers versus
ceramic veneers. The choice of material in the studies for indirect composite veneers was
Estenia C and B and IPS Empress Esthetic (Ivoclar-Vivadent AG, Schaan Liechtenstein) for
ceramic veneers. The veneer preparation for both composite and ceramic veneers was done
using incisal overlap (Gresnigt et al., 2013b; Gresnigt et al., 2019).

26
Table 9.2: Description of studies and types of failure for composite veneers
Author / year Gresnigt et al., Rapani et al., Gresnigt et Coelho de Gresnigt et al.,
2013b 2015 al., 2012 Souza et al., 2019
2015
Description of Randomized Prefabricated Randomized Direct anterior Randomized
study clinical trial of composite controlled composite clinical trial on
indirect resin veneers: split mouth veneers in vital indirect resin
composite and
composite and Retrospective clinical trial teeth
ceramic
ceramic veneers clinical of direct laminate
evaluation with composite veneers:
a four year veneers
follow up
Observation
3 years 4 years 3.5 years 3.5 years 10 years
time

Retrospective Randomized Retrospective Randomized


Randomized
Type of study controlled controlled
controlled study study study
study study

# of patients 10 20 23 86 11
Number of 23-Indirect 90-Indirect 96-Direct 196-Direct 24-Indirect
restorations composite composite composite composite composite
Microfilled
Eldewis ,universal
Composite type Estenia Micro hybrid Estenia
composite
composite
Incisal
Preparation type Incisal Overlap NA NA NA
overlap

Location Maxillary Maxillary NA NA Maxillary

Light cured Light cured Light cured Light cured


Light cured
Luting agent composite composite composite composite
composite cement
cement cement cement cement

Types of failures

Fracture 2 (A) 4 (A) 5 (A) 30 (A) 3 (A)


Debonding 1 (A) 0 6 (A) 0 3 (A)

Color match 0 0 1 (R) 0 8 (R)

Marginal
3 (R) 0 1 (R) 7(R) 12 (R)
intergrity
Surface
18 (R) 0 3(R) 0 18 (R)
roughness

Voids 6 (R) 0 0 0 14 (R)

Chipping 0 0 0 0 6 (R)
Wear of
0 0 0 0 7 (R)
restoration

Absolute failure 3 4 11 30 6

Relative failure 27 0 5 7 65

Survival rate 87% 95.6% 87.5% 80.1% 75%


(A) = Absolute failure, (R) = Relative failure, NA = Not available

27
Table 9.3: Description of studies and types of failure for ceramic veneer studies
Author Rinke et al., Chen et al., Guess et al., Layton & Ozturk Rinke et al., Fradeani et Ruiz et al., Guess et al., D’Arcangelo Gresnigt et Gurel, 2012 Beier et al., Arif et al., Gresnigt et Gresnigt et
2018 2005 2014 Walton and Bolay, 2013 al., 2005 2010 2008 et al., 2012 al., 2013 a 2012 2019 al., 2013 b al., 2019
2007 2014
Description of Retrospective Clinical Prospective Prospective Survival Retrospective Porcelain A clinical Mid-term Clinical Clinical Clinical Clinical Retrospective Randomized Randomized
Study evaluation of evaluation clinical study of 304 of study of laminate longitudinal results of a 5- evaluation longevity of performance performance evaluation of clinical trial clinical trial
extended heat of 546 study of porcelain porcelain extensive heat veneers study 323 year on porcelain ceramic of porcelain of porcelain the clinical of indirect on indirect
pressed tetracycline press veneers laminate pressed clinical porcelain prospective laminate laminate laminate laminate performance resin
resin
ceramic stained ceramic veneers ceramic evaluation laminate clinical veneers veneer veneers: veneers and longevity composite
composite and ceramic
veneers teeth treated overlap and with veneers veneers investigation bonded with bonded to outcome of of porcelain
and ceramic laminate
with full veneer different of extended light cured teeth with the aesthetic laminate
porcelain restoration degree of ceramic composite and without pre-evaluated veneers
veneers veneers:
laminate dentin veneers existing temporary
veneers exposure composite technique
restoration
Observation time 7 years 2.5 year 7 years 16 years 2 year 3 years 12 years 11 years 5 years 7 years 3.3 years 12 years 20 years 7 - 14 years 3 years 10 years
Type of study Clinical study Clinical Prospective Prospective Clinical Retrospective Retrospective Clinical Prospective Clinical Clinical Clinical Retrospective Retrospective Randomized Randomized
study study study study study study study study study study study study study controlled controlled
study study
# of patients 31 54 25 100 28 37 46 70 25 30 20 66 84 26 10 11
Numbers of 101 546 42 – 304 125 130 182 323 66 119 92 580 318 114 23 24
restorations overlap
veneer
24- full
veneer
Ceramic type Glass ceramic NA IPS Feldspathic IPS Emax Glass ceramic IPS Empress IPS Empress Glass NA Feldspathic IPS Empress Glass ceramic NA IPS Empress
Empress porcelain Esthetic+ Esthetic Ceramic porcelain I, II,Esthetic Esthetic IPS Empress
Esthetic veneers feldspathic + feldspathic Esthetic
veneers porcelain
Preparation type Palatal Non- Butt joint Palatal NA Palatal Butt joint or 124- simple NA Butt joint Butt joint NA NA Not Incisal overlap Incisal
overlap overlap overlap overlap palatal designs mentioned overlap
designs overlap 199-
Functional
designs
Location Maxillary- 65 NA NA Anterior NA Maxillary -76 Maxillary – Maxillary - Maxillary NA Maxillary Maxillary - Maxillary Maxillary -83 Maxillary Maxillary
Mandibular- and Mandibular- 127 238 anteriors teeth 414 Mandibular Mandibular-
36 posterior 54 Mandibular - Mandibular- Mandibular- 31
55 85 166
Luting agent Dual cured Dual cured Dual cured Dual cured Light Dual cured Light cured Dual cured Dual cured Light cured Light cured Light cured Dual cured NM Light cured Light cured
cements cement cement cement cured cement cement cement cement composite composite composite cement composite composite
composite
Types of failure
Fracture 7 (A) 0 1 (A) 6 (A) 2 (A) 4 (A) 5 (A) 13 (A) 1 (A) 3(A) 4 (A) 20 (A) 21(A) 5(A) 0 0
6(R)
Debonding 8 (R) 4 (A) 1(A) 3 (A) 1 (R) 3 (R) 0 29 (A) 0 0 1(A) 20(A) 2 (A) 2(R) 0 0
1 (R)
Color match 0 0 0 5 (R) 0 0 0 2 (R) 0 0 0 0 0 0 0 0
Marginal intergrity 2 (R) 0 0 0 0 0 0 10 (R) 0 0 16(R) 0 4(R) 0 0
1 (R)
Marginal 9 (R) 26 0 0 0 0 0 127( R) 0 0 12(R) 0 0 0 0 0
discolouration
Surface roughness 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
Voids 0 0 0 0 0 0 0 0 0 0 0 0 0 0 3(R) 10(R)
Chipping 1 ( R) 0 0 0 0 0 0 0 6(R) 8 0 0 1 (R) 6 0 0
Cracks 0 0 0 0 0 0 0 0 3(R) 7(R) 0 0 2 (R)
Wear of the 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
restoration
Absolute failure 7 4 2 9 2 4 5 42 1 3 5 40 23 5 0 0
Relative failure 20 26 0 5 1 3 0 139 9 15 28 0 5 12 3 10
Survival rate 93% 99% 97% 97% 98.4% 96.9% 94.4% 86.9% 98.4% 97.5% 94.6% 93% 92.7% 98% 100.% 100%
29
(A) = Absolute failure, (R) = Relative failure, NA = Not available

30
9.2 Veneer overall survival rate
For composite veneer restorations, 5 studies provided data on the survival rate of veneers, in
which a retrospective study was conducted by Rapani et al.,(2015) with 90 indirect composite
veneer restorations and observation time of 4 years with an overall survival rate of 95.6%.
Similarly, Coelho de Souza et al.,(2015) conducted a retrospective study where 196 direct
composite veneers were evaluated with 37 failures; the observation time of service for the
veneers was 3.5 years with an overall survival rate of 80.1%. Gresnigt et al.,(2012) conducted
a randomized controlled trial on 96 direct composite veneers, the observation period was 3.5
years with a survival rate of 87.5%. The author also conducted a randomized controlled study
in 2013 on 23 indirect composite veneer with an observational period of 3 years resulted in a
survival rate of 87% (Gresnigt et al., 2013b). The follow up study of Gresnigt et al. (2013b)
was conducted in 2019 on 24 indirect composite veneers over the period of 10 years with a
survival rate of 75% (Gresnigt et al., 2019). Therefore, 5 composite studies provided data of
survival of a total of 429 composite veneers after a mean follow-up time of 4.8 years, translating
a survival rate of 85.4% (Table 9.4).

Table 9.4: Survival rates for studies that included composite veneers

Study Study design Observation # of # of # of Survival


time patients restorations failures rate%

Rapani et al., 2015 Retrospective 4 years 20 90 4 (A) 95.6%


Clinical 11 (A)
Gresnigt et al., 2012 3.5 years 23 96 87.5%
study 5 (R)
Coelho de Souza et 30 (A)
Retrospective 3.5 years 86 196 80.1%
al., 2015 7(R)
Gresnigt et al., Clinical 3(A)
3 years 10 23 87.0%
2013b study 27(R)
Clinical 6(A)
Gresnigt et al., 2019 10 years 11 24 75.0%
study 65(R)

Within the 19 veneers studies on survival of the veneers in which only two studies compared
both composite and ceramic. One study (Gresnigt et al., 2013b) reported a clinical study on
both ceramic and composite veneers where 10 patients received 23 indirect resin composite and
23 ceramic veneer (IPS Empress Esthetic). In total 3 absolute failures were observed in the
group of composite laminate veneers. The overall survival rate observed for composite and
ceramic veneers was 87% and 100 % respectively. Similarly, Gresnigt et al.,(2019) conducted
a follow up study of Gresnigt et al.,(2013b) on 11 patient with a total of 24 indirect composite

31
resin and 24 ceramic veneers (IPS Empress Esthetic) reported 6 absolute failures in indirect
composite veneers with a survival rate of 75% over the period of 10 years whereas the study
reported no failures in ceramic veneers with a survival rate of 100%.

For ceramic veneer studies, a retrospective evaluation was done by Rinke et al.,(2013) where
37 patients had restorations with 130 glass ceramic veneers. Four absolute failures and three
relative failures were observed over a 3 years of observation time with a survival rate of 96.9%.
The author also conducted a retrospective study (Rinke et al., 2018) where a total of 31 patients
were restored with 101 glass ceramic veneers. The observation time was 7 years with seven
absolute and 10 relative failures with overall survival rate of 93%. Chen et al., (2005) conducted
a clinical evaluation of 546 tetracycline stained teeth with ceramic veneers. The veneers were
observed for 2.5 years. The study found a 99% survival rate (Chen et al., 2005). Layton and
Walton (2007) reported a prospective study to analyse 304 feldspathic porcelain veneers on 100
patients. The observation time was 16 years. The study reported 9 failures with a survival rate
of 97% (Layton and Walton 2007). Beier et al.,(2012) conducted a retrospective study on 84
patients with 318 glass ceramic veneers in a long-term analysis up to 20 years. Twenty-three
absolute and 5 relative failures were observed with the overall survival rate of 92.7% (Beier et
al., 2012). Arif et al.,(2019) reported on retrospective evaluation of 26 patients restored with
114 ceramic veneers (IPS Empress Esthetic) over an observation period of 14 years which
resulted in five absolute failure and 12 relative veneer failures with a survival rate of 98.%.
Gurel et al.,(2012) evaluated the clinical performance of 580 ceramic veneers (IPS Empress I,
II and Esthetic) on 66 patients over a 12-year period. Forty veneers failed in the study with a
survival rate of 93% (Gurel et al.,2012). Ozturk and Bolay (2014) conducted a clinical study
on 28 patients with 125 IPS Emax veneers. Subsequently, two absolute failures and one relative
failure was reported with overall survival rate of 98.4% after 2 years of follow up. Fradeani et
al., (2005) retrospectively evaluated the clinical performance of 182 ceramic veneers on 46
patients in which 143 veneers were made of IPS Empress Esthetic (Ivoclar-Vivadent AG,
Schaan Liechtenstein) and 39 restorations of feldspathic veneers. The study was evaluated over
12 years. As a result, five failures were observed with a survival rate of 94.4% at 12 years
(Fradeani et al., 2005). D’arcangelo et al.,(2012) conducted a clinical study on ceramic veneers,
in which 30 patients had restoration with 119 ceramic veneers (the type of ceramic was not
mentioned in the study) with an observation time of 7 years. In effect three absolute and 15
relative failures were observed with a survival rate of 97.5% at 7 years. Gresnigt et al.,(2013a)
conducted a clinical study to evaluate 92 ceramic veneers (Feldspathic porcelain) on 20 patients
over a period of 3.3 years. Five absolute and 28 relative failures were observed with a survival
rate of 94.6% (Gresnigt et al., 2013a). Guess et al.,(2008) conducted a prospective study where
32
25 patients had restoration with 66 IPS Empress Esthetic (Ivoclar-Vivadent AG, Schaan
Liechtenstein) ceramic veneers. After an observation time of 5 years, one absolute failure and
nine relative failures were observed with a survival rate of 98.4%. The authors also conducted
a prospective trial in 2014 comprised of 25 patients with 42 IPS Empress Esthetic (Ivoclar-
Vivadent AG, Schaan Liechtenstein) veneers. The observation period of the veneers was 7
years. Ten veneer restorations were lost during the period with 2 absolute failures reported
giving a survival rate of 97% (Guess et al., 2014).

Overall 16 ceramic veneer studies provided data on survival rates for 3,047 ceramic veneers,
with a mean follow up period of 7.8 years, resulting in a combined survival rate of 95.9%
across all of these studies. The summary of survival of ceramic veneers is illustrated in the
Table 9.4.

Table 9.4: Survival rates for studies that included ceramic veneers
Author/s and year Study design # of # of Observation # of Survival
of publication Pt restorations time failures rate
Rinke et al., 2018 Retrospective 31 101 7 years 7 (A) 93%
10 (R)
Chen et al., 2005 Clinical study 54 546 2.5 year 4(A) 99.0%
26(R)
Guess et al., 2014 Prospective 25 42 7 years 2(A) 97%
Layton and Walton Prospective 100 304 16 years 9(A) 97.0%
5(R)
Ozturk et al., 2014 Clinical Study 28 125 2 years 2 (A) 98.4%
1 (R)
Rinke et al., 2013 Retrospective 37 130 3 years 4 (A) 96.9%
3(R)
Fradeani et al., 2005 Retrospective 46 182 12 years 5 94.4%
Ruiz et al., 2010 Clinical study 70 323 11 years 42(A) 86.9%
139(R)
Guess et al., 2008 Prospective 25 66 5 years 1 (A) 98.4%
study 9 (R)
D’Arcangelo et al., Clinical study 30 119 7 years 3 (A) 97.5%
2012 15 (R)
Gresnigt et al., 2013a Clinical study 20 92 3.3 years 5(A) 94.6%
28(R)
Gurel et al., 2012 Clinical study 66 580 12 years 40(A) 93%
Beier et al., 2012 Retrospective 84 318 20 years 23 (A) 92.7%
5 (R)
Arifet et al., 2019 Retrospective 26 114 4 years 5(A) 98%
12(R)
Gresnigt et al., 2013b Clinical study 10 23 3years 3(R) 100.0%
Gresnigt et al., 2019 Clinical study 11 24 10 years 10 (R) 100.0%

33
Pt = Patients, (A) = Absolute failure, (R) = Relative failure, NA = Not available

9.3 Absolute Veneer Failures


Composite veneer failures
In the five studies reporting on composite veneers, information about the absolute failures was
given. Firstly, Gresnigt et al., (2012) reported that in the group of 96 direct composite veneers,
there were a total of 11 absolute failures in the form of debonding (6) and fracture (5). Out of
the six debonding, two occurred at dentin-veneer interface. Fractured veneers were due to
adhesive failures between the tooth and veneer, postoperative sensitivity was seen in six veneers
which disappeared in a week. The authors also conducted another clinical study on 23 indirect
composite veneers in 2013 and reported three absolute failures in the form of fracture (2) and
debonding (1). The study reported that the debonding was due to the adhesive failure between
the tooth and cement. The two veneer fractures occurred in the incisal area of the veneers.
(Gresnigt et al., 2013b). The follow up study of Gresnigt et al.,(2013b) was reported in 2019
noticed six absolute failures in the form of fracture (3) and debonding (3). The debonding
observed in the study was due to the adhesive breakdown between the tooth and the cement. In
addition, the three fractures were noticed on the incisal area of the indirect veneers. Another
study conducted by Rapani et al.,(2015) on 90 indirect composite veneers reported four absolute
failure in the form of fracture on the incisal surface of the veneers. Lastly, Coelho de Souza et
al.,(2015) reported 30 absolute failures out of 196 direct composite veneers. The absolute
failures were in the form of fracture (30).

Ceramic veneer failures


Out of the 16 studies evaluated only 14 reported absolute failures. Firstly, Chen et al.,(2005)
analysed 546 veneers over 2.5 years. The study noticed four veneer failures in the form of
debonding at 6 months. Fradeani et al.,(2005) evaluated 182 veneers from which five veneers
failed due to fracture. Out of which two longitudinal fractures were replaced and the other three
fractures which were very limited were rebonded, however, the author reported these five
veneers as failures. Layton and Walton (2007) identified the absolute failure of 9 veneers in
their study. The failed veneers were associated with fracture (6) and debonding (3). Guess et
al., (2007) evaluated 66 veneers out of which 42 were overlap veneers and 24 were full veneers
of which one absolute failure by fracture was noted. The author also conducted another
prospective study in 2014 on 66 IPS Empress Esthetic (Ivoclar-Vivadent AG, Schaan
Liechtenstein) in which two absolute failures were observed in the form of debonding (1) and

34
fracture (1). Ruiz et al., (2010) evaluated 323 veneers and reported a total of 42 absolute
failures. 13 of them were attributed to fracture and 29 to debonding. D’arcangelo et al.,(2012)
reported 3 absolute failures in the form of fracture out of the 119 veeners evaluated over a
period of 7 years. Gurel et al.,(2012) evaluated the performance of 580 veneers with aesthetic
pre-evaluation temporary techniques. The study observed a total of 40 absolute failures,
resulting from fractures (20), and debonding (20). Beier et al.,(2012) recorded 23 absolute
failures in which fractures (11) were the main reason for failure, other cause identified was
veneer debonding (2). Marginal discolouration was significantly higher in patients who
smoked. Rinke et al.,(2013) retrospectively evaluated 130 ceramic veneers over a period of 3
years and observed a total of 4 absolute failures all resulting from fractures. The author
followed them up until 2018 and reported seven absolute failures again all from fractures
(Rinke et al., 2018). Gresnigt et al.,(2013a) evaluated 92 veneers of which 26 veneers were on
intact teeth and 66 on existing composite restorations which was in good condition and was not
removed. The study resulted in five absolute failures due to fractures (4) and debonding (1).
The author conducted another randomized controlled study in 2013 on 23 IPS Empress Esthetic
(Ivoclar-Vivadent AG, Schaan Liechtenstein) veneers with no absolute failure (Gresnigt et al.,
2013b). The study was further followed up until 2019 with 24 IPS Empress Esthetic (Ivoclar-
Vivadent AG, Schaan Liechtenstein) veneers over a period 10 years with no absolute failure
(Gresnigt et al., 2019). Ozturk and Bolay (2014) assessed 125 ceramic veneers IPS Emax over
a period of 2 years. Two absolute failures were observed in the form of fractures at the cervical
margin of the veneers. Arif et al., (2019) retrospectively evaluated 114 porcelain laminate
veneers over a period 14 years and reported five absolute veneers failures all attributed to
fractures. Coelho de Souza et al.,(2015) reported 7 relative failures in the form of loss of
marginal integrity.

9.4 Relative Veneer Failure


Composite veneers
Within the five composite veneer studies, three studies reported on relative failures. Gresnigt
et al., (2012) reported five relative failures in the form of loss of marginal integrity (1) surface
roughness (3) and discolouration (1). In addition six teeth showed post-operative sensitivity
which disappeared after a week. Another study by Gresnigt et al., (2013b) reported 27 relative
failures in the form of voids and defects (6), loss of marginal integrity (3) and surface roughness
(18) in indirect composite veneers. Similarly, the follow up of Gresnigt et al., (2013b) in 2019
also reported 65 relative failures in the form of minor voids and defects which was noticed in
14 composite veneers, loss of marginal integrity was reported in 12 composite veneers. Surface

35
roughness was significant in 18 composite veneers and lastly chipping (6) and wear of the
restorations (7) was observed. In addition, postoperative sensitivity was noticed on eight teeth,
which disappeared in 2 weeks.

Ceramic veneers
Within the 16 studies evaluated four studies did not yield any relative failures (Fradeani et al.,
2005; Gurel et al., 2012; Guess et al., 2014; Rinke et al., 2013). A clinical study conducted by
Chen et al., (2005) observed 26 relative failures in the form of discolouration at the cavo surface
margin; no discolouration recurred after the veneers were polished. In addition, postoperative
sensitivity occurred in 20 patients during the first 2 weeks and resolved in all cases. Guess et
al., (2008) recorded 9 relative failures of ceramic veneers with six chipping, which were located
in the palatal area of the veneers; in addition, there were 3 cracks of the veneer. Layton &
Walton (2007) reported five cases of relative failure ,all resulting from colour mismatch. Ruiz
et al., (2010) observed 139 relative failures out of 323 total veneer restorations. Marginal
discoloration was noticed in 127 cases at the tooth restoration interface. In addition 10 cases
with loss of marginal integrity and two cases with colour mismatch was observed. D’arcangelo
et al. (2012) discerned 15 relative failures in which 8 veneers showed chipping (5 chipping in
the palatal area, 2 at the incisal and 1 in the cervical area). Cracks were observed in seven
veneers; since the cracks were very minimal, the veneers were not required to be renewed. Beier
et al., (2012) observed five relative failures in the form of cracks (2), debonding (1) loss of
marginal integrity (1) and chipping (1). Gresnigt et al., (2013a) observed 28 relative failures of
which 16 were due to loss of marginal integrity and 12 resulted from marginal discolouration.
In Gresnigt et al.’s (2013b) study, three relative failures all arising from voids were observed
in ceramic veneers. Similarly, 10 voids were observed in the follow-up study of Gresnigt et al.,
(2019). Ozturk and Bolay (2014) reported one relative failure due to debonding. Rinke et
al.,(2018) noticed 20 relative failures in the form of debonding (8), loss of marginal integrity
(2), chipping (1) and marginal discolouration (9). Arif et al., (2019) reported 12 relative failures
in the form of debonding (2), loss of marginal integrity (4) and crack (6).

9.5 Veneer luting agents

9.5.1 Composite veener luting agents

Among the four composite veneer studies, the direct composite veneer in Coelho de Souza et
al., (2015) were fabricated using a universal composite or a microfilled composite and was light
cured using LED polymerization units. Gresnigt et al., (2012), in their study after etching of
enamel, a 3-methacryloxypropyltrimethoxyl silane coupling agent (MPS) was applied on the

36
existing restoration for 5 minutes. The veneers were bonded using Ena bond (Micerium S.p.A.,
Avegno, Italy ) for enamel plus HFO (Micerium S.p.A., Avegno, Italy ) and clearfil SE bond
(Kuraray Medical Inc, Tokyo, Japan ) for miris 2 (Coltène Whalent GmbH, Alstätten,
Swizterland.) . Furthermore, the indirect composite veneer in the study of Gresnigt et
al.,(2013a) was silanized using monobond S ( Ivoclar Vivadent) before the cementation.
Variolink ( Ivoclar Vivadent) was the adhesive resin used to cement the veneers. Similarly,
another study conducted by Rapani et al.,(2015) for indirect composite veneers used universal
bond on tooth for 10 seconds and light cured for 20 seconds. The veneers were cemented using
light cured composite resin which was applied on to the tooth surface and the inner surface of
the veneers.

9.5.2 Ceramic veneer luting agent

Layton & Walton (2007) used dual cured cement for the cementation of 304 veneers in the
study .In the study of Beier et al., (2012) on 318 ceramic veneers, the veneers were cemented
using dual cured cement. D’arcangelo et al.,(2012) used prime and bond NT as the dentin
adhesive, and a light cured composite cement were used for the cementation of 119 ceramic
veneers. In the study by Ruiz et al.,(2010), the internal surface was treated using 10%
concentration of fluorohydric acid before the cementation of the veneers using dual cured
cement. Gurel et al. (2012) in a study of 580 ceramic veneers, the veneers used light-cured
cement for the cementation. Guess et al. (2008) and Guess et al. (2014) used Heliobond in 66
ceramic veneers to avoid inaccuracies of fit, it was light cured before the cementation. The
veneers were cemented using dual polymerizing composite cements. In the study of Gresnigt
et al.,(2013a) , the 92 veneers were cemented using light cured cement. The veneers in Gresnigt
et al.’s (2013b) and Gresnigt et al.’s (2019) study were cemented using a light cured composite
cement. Similarly, 125 veneers in Ozturk and Bolay’s (2014) study were cemented using light-
cured resins. The 546 ceramic veneers in Chen et al.’s (2005) study were bonded with
tenure(Den-Mat Corporation, USA) and was cemented using a dual cured composite resin. In
the study of Fradeani et al., (2005), the 182 veneers were bonded using syntac (Ivoclar
Vivadent) and was cemented using a dual polymerizing cement. Rinke et al., (2013) and Rinke
et al.,(2018) used dual cured cement for the cementation of 130 and 101 veneers respectively.

37
Discussion

The latest systematic review on the survival rate of composite versus ceramic veneers was done
by Wakiaga et al., (2004). In this review only one study met the inclusion criteria; however, the
study included in the review had a very small sample size to determine the longevity of the
veneers. Therefore, due to the insufficient data, the comparison between indirect and direct
veneers with regards to longevity could not be established.

The current systematic review summarizes the survival rate of composite and ceramic veneer
based on the available data. There are a total of 19 articles, of which two are randomized
controlled trial studies, comparing both ceramic and indirect composite veneers (Gresnigt et
al., 2013b; Gresnigt et al., 2019). The studies were given moderately high ‘quality of reporting’
scores of +3 in accordance with the grade scoring system (Guyatt et al., 2008). Additionally, 3
articles solely focused on composite veneers and 14 on ceramic veneers.

Even though the current systematic review has limited data for composite veneers, these studies
are of a high standard and therefore the survival rate of the composite veneers can be analysed
in this systematic review.

Among the 14 papers addressing ceramic veneers only 6 mentioned the material and the
preparation used, which has a definite impact on the longevity of ceramic veneers (Fradeani et
al., 2005;Gresnigt et a., 2013a; Guess et al., 2014; Layton & Walton 2007 et al., 2007; Rinke
et al., 2013; Ruiz et al., 2010). Among the papers addressing ceramic veneers one study had a
very small number of patients, with several patients not coming for the follow-up appointments
(Guess et al., 2014). Another study reported insufficient data on the preparation designs, veneer
material and operator consistency (Arif et al., 2019). Each of these factors affects the longevity
of the veneers and brings into question the validity of the study (D’arcangelo et al., 2012).

10.1 Survival rate of composite veneers

Within the 3 studies which included 2 direct composite veneer studies and 1 indirect composite
veneer studies, all showed a satisfactory clinical performance. From the overall results obtained
by the individual studies on composite veneers, while there was a limited number of studies,
the quality was good, and we can conclude that indirect composite veneers show a better
survival rate compared to the direct composite veneers at a mean observational period of 4

38
years. The main reason for failure of direct composite veneers was fracture, but detailed
information on the materials and methods is not available.

10.1.1 Direct versus indirect composite veneer

Firstly, Gresnigt et al., (2012) conducted a clinical study on 96 direct composite veneers using
two different micro-hybrid composites. An incisal overlap preparation of 1mm was used in
cases which required translucency. Twelve absolute failures were observed over a period of 3.5
years. The main reasons for failure observed in the study were fracture, debonding and surface
discolouration. Debonding and fracture was due to the consequence of the failure of adhesive
technique as adhesion plays an important role in the longevity of the restorations mainly in
restoration where they is no mechanical retention. In addition to the absolute failures, surface
roughness or marginal discolouration was noticed in 2 out of 45 cases, due to air entrapments
on the outer layer of the composite veneer and was explained by the increased filler size. The
composites in the study had a filler size with a range of 0.02 to 2.5µm content changing the
viscosity and handling properties of the material. In comparison to microfilled composites,
microhybrid composites are hard to polish (Meijering et al., 1995). However, the type of the
material did not affect the survival rate of the composite veneers. Similarly, Coelho de Souza a
et al., (2015) conducted a retrospective study which used micro filled and universal hybrid
composite (microhybrid, nanofilled, nanohybrid) for their direct composite veneers. The
percentage of filler in the composite was 37.5 vl in microfilled composite and was about 63.3vl
in universal hybrid composites. The mean particle size of microfilled was 0.04µm and ranged
between 0.5µm to 75nm for universal hybrid composites. The preparation design was not
mentioned in the study. The main reasons for failure were fracture and caries. The authors
reported that vital teeth have better resistance than non-vital teeth as bonding veneers to
endodontically treated teeth decreases compared to adhesive techniques on vital teeth (Bitter et
al., 2009). There was no difference in the survival rate based on the type of material used for
composite veneers as observed in Gresnigt et al.’s (2012) study. However, the microfilled
composite veneers showed a better lustre , less surface staining, colour match and better
marginal adaptation compared to hybrid composites. Therefore, both the direct composite
veneers studies did not show a significant change in the survival rate with the use of different
composite resin material but identified different types of failures (Coelho de Souza et al., 2015;
Gresnigt et al., 2012).

In comparison, Rapani et al.’s (2015) study of 90 indirect composite veneer on the maxillary
arch over the period of 4 years reported the main reason for failure to be a fracture that was
confined to the tooth cement interface and the veneer restoration. The study reported a survival

39
rate of 95.6% which was higher than the two direct composite veneers studies (Coelho de Souza
et al., 2015; Gresnigt et al., 2012). Rapani et al., (2015), therefore, reported in the indirect
composite veneer study that adhesive technique procedures strengthen the tooth structure which
improves the biological and mechanical function and leads to aesthetic results with reduced
cement shrinkage.

When direct composite veneers are compared to indirect veneers, the surface texture of the
indirect composite veneers are influenced by the finishing and polishing techniques which can
also increase the clinical longevity (Gresnigt et al., 2012). The degree of conversion also affects
the colour stability and the opacity of the veneers which then in turn influences the clinical
longevity. On the other hand, the risk factors of direct composite veneers are composite fracture
and marginal discoloration.

To summarise, the composite veneer studies had a mean observation time of 4.8 years, with an
overall survival rate of 85.4%. The main reasons for failure in all the studies were fracture and
marginal discolouration. From the composite veneer studies reviewed, failure of veneers was
caused mainly by fracture irrespective of the type of material used, but adhesive system played
an important role in the longevity of composite veneers.

10.2 Survival rate of ceramic veneers

Within the 14 ceramic veneer studies, 3 studies used feldspathic porcelain, 7 studies used leucite
reinforced glass ceramic and four studies did not mention the type of ceramic material used.
All the studies in the review showed an overall survival rate of 95.9% over a mean observation
time of 7.8 years. The main reasons for failure in the ceramic veneers were parafunctional
activity and the type of substrate used for the bonding of the veneers.

10.2.1 Feldspathic veneers


In the 14 ceramic veneers studies, 3 studies used feldspathic porcelain for the fabrication of
ceramic veneers. According to the grade scoring system used for reviews (Guyatt et al., 2008),
all the three studies were graded ‘moderate’ score of 3. In these studies, butt joint or palatal
joint were used for the incisal preparation. Layton and Walton (2007) conducted a prospective
study on 304 feldspathic porcelain restorations which used palatal overlap for the incisal
preparation for an observational period of 16 years. In their study, the preparation was mostly
limited to the enamel and then cemented using dual cure unfilled resin composite cement, they
depict a good long-term survival rate of 97% at five to six years, with a low failure rate when

40
the veneers were bonded to enamel substrates. The main reasons for failure in the study were
fractures.

Two other studies that used feldspathic porcelain reported that failure was due to degradation
of the adhesive system leading to veneer fracture. Both of the studies used butt joint and palatal
overlap. Firstly, the study by Fradeani et al.,(2005) reported on 182 veneers fabricated with IPS
Empress and feldspathic porcelain, out of which 143 were made using IPS Empress and 39
veneers were fabricated using feldspathic porcelain. The veneers were luted using light cure
cements. The study had excluded patients with poor oral conditions, gingivitis, periodontitis
and patients with parafunctional activity which can affect the survival rate of the veneers.
Fracture was the main reason for failure and was noticed commonly in IPS Empress veneers.
In this research dentin was exposed in 50% of the cases and therefore, the amount of tooth
reduction is most likely responsible for the failures and not the ceramic material used. The other
prospective study was done on 92 feldspathic veneers by Gresnigt et al.,(2013a).They used butt
joint preparation and cemented with dual cure adhesive cement. Failures were also due to
fracture and chipping, with three of the fractures due to the parafunctional habits. However,
there was chipping observed in the cervical area of the teeth which was due to exposed dentin.

Therefore, from the three studies reviewed feldspathic porcelain, which used incisal overlap
preparation and similar bonding techniques,the observed fracture were either due to preparation
technique or the degradation of the adhesive. Therefore, the failures were not related to the
material but more to which substrates the veneers were bonded and ageing of the adhesive
cement.

10.2.2 Leucite reinforced glass ceramic veneers

Leucite based glass ceramic was used in 7 studies in this review. IPS Empress I, II, Esthetics
were used in the studies. Two thirds of the failure in these studies were credited to fracture and
debonding. Guess et al.,(2008) also indicated that the fracturing which was seen in 66 IPS
Empress Esthetic (Ivoclar-Vivadent AG, Schaan Liechtenstein) veneers out of which 42 used
butt joint preparation and 24 were add a palatal chamfer. Both the preparation designs were
similar in proximal and buccal surfaces, whereas the full veneers had a distinct palatal
extension. The veneers were luted using dual polymerizing composite. The study reported that
the fracture seems to be more frequent when strong laterotrusion contact on the restoration was
present. In addition, chipping was also observed with parafunctional activity. Another reason
was the failure of the adhesive system, leading to debonding of the veneers. This study
concluded that ageing of the adhesive system could be the main reason for failure of the veneers.

41
The author also conducted a follow up study of Guess et al.,(2007) on 66 IPS Empress Esthetics
in 2014 with 42 butt joint preparation and 24 palatal chamfer preparation. This study did not
have enough samples as many patient were lost during the 7-year follow up. The study reports
failure due to the ageing of the adhesive system, but this study based its conclusions on a very
small sample size (Guess et al.,2014). Another study by Gurel et al. (2012) also reported 20
fractures in 580 veneers due to improper bonding techniques. The study suggested that the
bonding technique may influence the marginal discolouration of the veneers. The study also
reported a low failure rate with a survival rate of 93% over 12 years. This seems to corroborate
the findings of other research regarding the type of substraste on which the veneers are bonded.

Therefore, leucite-reinforced glass ceramic studies observed a high survival rate in the studies.
However, the failure of the veneers in the studies were due to different reasons. The major
reason for failure was ageing of the adhesive system and the preparation technique. In addition,
parafunctional activity was also a reason for fracture of veneers.

On the other hand, four studies did not mention the type of the ceramic used. Firstly, Chen et
al.,(2005) conducted a study in 546 tetracycline-stained teeth with ceramic veneers which was
bonded using dual curing composite cement. No incisal preparation design was used in the
study as the teeth were already worn incisally. This study uses an extensive labial reduction to
remove the stains, leading to dentin exposure. Nevertheless the overall results showed a high
survival rate of 99% of the veneers using an improved adhesive technique. The study reapplied
silane coupling agents on to the veneer surface before luting the veneers to tooth surface. This
is a contradiction with the existing literature showing more debonding when the substrate is
dentin. This study with a grade of ‘low’ due to the sparse data in accordance with the grading
scoring system used for reviews (Guyatt et al., 2008) had a main limitation of a very short
observation time (only 2.5 years), which could also explain the good survival rate observed.
Another study done by D’Arcangelo et al.,(2012) of 119 veneers over an observational period
of 7 years showed promising longevity of the veneers. All the veneers were bonded on to the
enamel but the veneer with a minimum dentin exposure during preparation was treated with
immediate dentin sealing prior to the bonding. The main reason for the low failure with a
survival rate of 97.5% seen in the study was due to the importance given to the preparation
margins which was limited to enamel. The authors reported that the favourable marginal
adaptation was correlated to the cementation and preparation designs. The study reported no
debonding of veneers.

42
10.3 Relative failure of composite and ceramic veneers

“A relative failure was defined as fracture, chipping, dislodgement or shade correction of the
veneer restoration allowing repair or correction with a direct resin composite material or
rebonding without the need for new veneer restorations or other restoration” (Meijering et al.,
1998).

According to the recent clinical study which compared the survival rate of composite and
ceramic veneers over the period of 10 years findings, observed the common variables that
influenced the survival rate of veneers were discolouration, surface roughness, chipping and
wear of the restoration (Gresnigt et al., 2019).

10.3.1 Surface roughness


According to Gresnigt et al.’s (2012) study, surface roughness was observed in 3 out of 45 cases
caused by the air entrapments in the outer layer of the composite veneers. The study also
mentioned increased filler content leads to harder material, hence making it difficult to polish.
In addition, the study also added that degradation of the filler due to the hydrolysis and
shrinkage will result in the separation of the resin matrix which causes porosities between the
interfaces of the fillers. In Gresnigt et al.’s (2013b) study, 18 cases out of 23 indirect composite
veneers presented surface roughess which was attributed to the aging of the resin material which
leads to the leakage of the resin matrix which eventually leads to rough surface. The follow-
up study (Gresnigt et al., 2019) reported that the surface roughness was observed more
frequently in indirect composite veneers compared to the ceramic veneers, and was due to the
ageing of the adhesive cement. Therefore, the study reports that surface roughness was one of
the factors for replacing resin composite restoration.

10.3.2 Colour match


For composite veneer, a study conducted by Gresnigt et al.,(2012) reported one case of colour
mismatch out of 96 direct composite veneers evaluated. This was due to air entrapment in the
superficial layer of veneers. Similarly the study conducted by Gresnigt in 2019 also reported
eight indirect composite veneer cases with colour mismatch whereas the ceramic veneers in the
study balanced with the surrounded teeth. Another study conducted by Guess et al.,(2008)
showed a good colour match of IPS Empress Esthetic (Ivoclar-Vivadent AG, Schaan
Liechtenstein) due to the high translucency of the material. Therefore, the variable colour match
was less favourable for composite veneers compared to ceramic veneers.

43
10.3.3 Marginal discolouration and wear of the restoration
For ceramic veneers marginal discolouration was associated with decreased marginal intergrity
and the degradation of the adhesive cement especially in dual polymerized cement. Exogenous
deposits, unfavourable enamel morphology, and marginal dentin exposure are the additional
factors associated with the marginal discolouration (Guess et al., 2008). A study conducted by
Fradeani et al.,(2005) observed reduced marginal discolouration as the study had applied a light
polymerizing luting composite for the cementation of veneers. In Chen et al.’s (2005) study,
out of 546 tetracycline stained treated teeth, 26 veneers reported slight discolouration at the
margin due to the opaque porcelain which produced whitish discolouration. Another study by
Gresnigt et al.,(2013a) observed 12 out of 87 laminate veneers due to the ageing and
degradation of the adhesive resin or due to the polymerization shrinkage. For composite
veneers, Coelho de Souza et al. (2015) reported lower marginal discolouration in microfilled
composite compared to universal composites due to the lower filler size. Moreover a recent
systematic review done in 2014 comparing nanofilled and microfilled composites for anterior
aesthetic reported that microfilled composites had a better performance for anterior aesthetics
(Kaizer et al., 2014).

10.4 Effect of bonding agent on ceramic veneer failure

Within 19 veneer studies, 10 studies used light cured cement (Chen et al., 2005; Coelho de
Souza et al., 2015; D’arcangelo et al., 2012; Gresnigt et al., 2012; Gresnigt et al., 2013a;
Gresnigt et al., 2013b; Gurel et al., 2012; Öztürk & Bolay, 2014; Rapani et al., 2015) and 9
studies used dual cured cement (Beier et al., 2012; Fradeani et al., 2005; Guess et al., 2014;
Guess & Stappert, 2008; Layton & Walton, 2007; Rinke et al., 2013; Rinke et al., 2018; Ruiz
et al., 2010). According to the study of D’arcangelo et al.,(2012), dual cured cement are
benefited by their self-curing system which helps in the conversion even in absence of radiant
source, however the disadvantage of dual cured cement is that it is extensively fluid and mixture
of two cements may lead to porosities. On the other hand, light cured cements are easily
manageable and provides an accurate margin, but the light polymerization of the cement
through the veneers can be a problem in a thick and high chroma veneer restoration resulting
in blocking partially the light resulting in a partial curing. Therefore, D’arcangelo et al.,(2012)
conducted a study on 119 veneers over an observational period of 7 years showed a promising
longevity of the veneers with a survival rate of 97.5%. The major limitation of the study was
that the type of ceramic material was not mentioned. All the veneers were cemented using light
cured cement. The main reason for the low failure seen in the study was due to the importance
given to the minimal preparation margins which were kept on enamel. The authors reported

44
that the favorable marginal adaptation and colour match was correlated to the cementation and
preparation designs. The study reported no debonding of veneers. In Gurel et al., (2012) study
conducted on 580 IPS Empress Esthetic (Ivoclar-Vivadent AG, Schaan Liechtenstein) veneers
reported that fracture, microleakage and marginal discoloration was associated with incomplete
polymerization of light cured adhesive cement which increases the hydrolysis and leads to
weariness of the bond within the ceramic veneers either from the functional load or due to the
degradation of the resin matrix.

10.4.1 Effect of parafunctional activity on ceramic veneer failure

Beier et al.,(2012) conducted a study on 318 glass-ceramic veneers with the longest
observational period (20 years). The type of glass ceramic and the preparation type was not
mentioned which was the major limitation in the study. Therefore, the study was given
moderate ‘quality of reporting’ scores of 3 in accordance with the grade scoring system (Guyatt
et al., 2008). This study observed that the main reason for fracture was parafunctional activity.
According to this study the risk for veneer failure was 8 times higher in parafunctional activity.
However, the study observed a survival rate of 92.7% after 5 years which corresponds with
Layton and Walton (2007) study. In addition, this study also observed marginal discoloration
in smoker patients. A recent study conducted by Arif et al.,(2019) on 114 IPS Esthetics veneers
also observed chipping and fracture which was due to parafunctional activity. The survival rate
of the study was consistent with the survival rate in Fradeani et al.’s (2005) study. Another
study conducted by Ruiz et al.,(2010) on 323 IPS Empress Esthetic (Ivoclar-Vivadent AG,
Schaan Liechtenstein) veneers out of which 124 had a feather edge preparation design and 199
veneers had a palatal chamfert. The study reported that 13 fractures and 29 decementations
were observed in their study attributed to parafunctional activity.

10.4.2 Effect of substrate on ceramic veneer failure


Rinke et al.,(2013) conducted a study on 130 invasive ceramic veneers preparation, which
resulted mainly in a dentine substrate available for bonding. The type of material used was not
mentioned in the study. The veneers used palatal chamfer and were bonded using dual curing
cements. However, this study had a relatively short observation period (36 months) and many
patients were lost during the follow up. The study observed fracture and debonding as the main
reason for failure which was associated with the bonding to the dentin substrate. This study
showed a reduced survival rate as the veneer preparation exposed over 50% of dentin. The
same authors conducted a seven year follow up study in 2018. The samples was reduced to 101
veneers as four patients were lost during the follow up. This study observed seven ceramic
fractures and eight debonding failure with a survival rate of 93.6% after 7 years. According to
45
the Guess et al.,(2014) study, which used invasive veneer preparation design, they showed a
similar survival rate of 97.5% after seven years. The results of this study confirm a recent
systematic review conducted by Morimoto et al.,(2016) in which the main reason for the
ceramic fracture was due to higher flexion of the veneers when bonded to a dentin substrate
(Rinke et al., 2018).

10.5 Composite veneers versus ceramic veneers

Of the 19 veneer studies reviewed, only two studies compared composites and ceramic veneers.
Firstly, Gresnigt et al.,(2013b) conducted a randomized controlled trial study on 23 indirect
composite veneers and 23 IPS Empress Esthetic (Ivoclar-Vivadent AG, Schaan Liechtenstein)
veneers in the same mouth over the period of three years. The study was given moderately high
‘quality of reporting’ scores of +3 in accordance with the grade scoring system (Guyatt et al.,
2008). This study used palatal chamfer for the incisal preparation. A total of three absolute
failures were noticed for indirect composite veneers (debonding and fracture).The overall
survival rate of composite veneers and ceramic laminate veneers observed in the study was 87%
and 100% respectively.

A follow-up study of Gresnigt et al.,(2013b) was conducted in 2019 on 24 IPS Empress


Esthetics and 24 composite veneers. The study was given moderately high ‘quality of reporting’
scores of +3 in accordance with the grade scoring system (Guyatt et al., 2008). The study
reported 6 absolute failures in indirect composite resin in the form of three debondings and
three fractures, similar to the 2013 study none of the ceramic veneers failed. The fracture
observed in the composite veneers were at the incisal edge when the substrate was
predominately dentin and bruxing habit was present. Overall in this high-grade study ceramic
veneers performed better than the composite veneers.

46
Conclusions

• Within its limitations, this systematic review identified that, ceramic veneers show a
higher survival rate than composite veneers with a mean observational period of 7.8
years and 4.8 years respectively.
• From the overall type of failure , ceramic veneers have a tendency to demonstrate more
absolute failures than relative failure when compared to composite veneers.
• The variables colour match, marginal discolouration, surface roughness was less
favourable for composite veneers.
• The failure of the composite veneers was mainly dependent on ageing of the adhesive
system used and not the type of composite material used .
• The failure of the ceramic veneers was mainly dependent on parafunctional activity,
ageing of the adhesive system and dentin exposure.

47
Recommendations
The author recommends that more randomised controlled trials involving both types of veneers
are required to better assess the survival rate and quality of survival of composite veneers and
ceramic veneers.

Clinical Significance
It seems that when indicated, anterior ceramic veneers may be preferred over composite
veneers. Nevertheless, patients in need of veneer restoration should be made acquainted with
the options available to them along with the relative lack of research evidence for each type of
veneer.

48
References

Aboushelib, M., de Jager, N., Kleverlaan, C. J., & Feilzer, A. J. (2007). Effect of loading
method on the fracture mechanics of two layered all-ceramic restorative systems. .
Dental Materials, 23(8), 952-959.
Aboushelib, M. N., Kleverlaan, C. J., & Feilzer, A. J. (2006). Microtensile bond strength of
different components of core veneered all-ceramic restorations: Part II: Zirconia
veneering ceramics. Dental Materials, 22(9), 857-863.
Ali, Z., Eliyas, S., & Vere, J. W. (2015). Choosing the right dental material and making sense
of the options: evidence and clinical recommendations. 23:150-162. Eur J Prosthodont
Restor Dent, 23, 150-162.
Anusavice, K. J. (1996). Phillips Science of Dental Materials. Saunders, 707, 3.
Arif, R., Dennison, J. B., Garcia, D., & Yaman, P. (2019). Retrospective evaluation of the
clinical performance and longevity of porcelain laminate veneers 7 to 14 years after
cementation. The J Prosthet Dent, 122(1), 31-37.
Aristidis, G., & Dimitra, B. (2002). Five-year clinical performance of porcelain laminate
veneers. Quintessence International, 33(3), 185-189.
Aschenbrenner, C., Lang, R., Handel, G., & Behr, M. (2012). Analysis of marginal adaptation
and sealing to enamel and dentin of four self-adhesive resin cements. Clinical Oral
Investigations, 16(1), 191-200.
Attia, A., & Kern, M. (2004). Fracture strength of all-ceramic crowns luted using two bonding
methods. 91(3):247-252. The J Prosthet Dent, 91(3), 247-252.
Aykor, A., & Ozel, E. (2009). Five-year clinical evaluation of 300 teeth restored with porcelain
laminate veneers using total-etch and a modified self-etch adhesive system. Operative
dentistry, 34(5), 516-523.
Bagheri, R., Burrow, M. F., & Tyas, M. (2005). Influence of food-simulating solutions and
surface finish on susceptibility to staining of aesthetic restorative materials. . Journal
of Dentistry, 33(5), 389-398.
Bagis, B., Aydoğan, E., & Bagis, Y. H. (2008). Direct restorative treatment of missing
maxillary laterals with composite laminate veneer: a case report. The open dentistry
journal, 2(93-95).
Baldissara, P., Llukacej, A., Ciocca, L., Valandro, F. L., & Scotti, R. (2010). Translucency of
zirconia copings made with different CAD/CAM systems. The J Prosthet Dent, 104(1),
6-12.

49
Beier, U. S., Kapferer, I., Burtscher, D., & Dumfahrt, H. (2012). Clinical performance of
porcelain laminate veneers for up to 20 years. International J Prosthet Dent,, 25(1), 79-
85.
Ben-Amar, A. (1989). Porcelain laminate veneers: for improved aesthetics of anterior teeth.
Refuat Hashinayim, 7(1), 17-23.
Bergoli, C. D., Meira, J. B. C., Valandro, L. F., & Bottino, M. A. (2014). Survival rate, load to
fracture, and finite element analysis of incisors and canines restored with ceramic
veneers having varied preparation design. Operative dentistry, 39(5), 530-540.
Beun, S., Glorieux, T., Devaux, J., Vreven, J., & Leloup, G. (2007). Characterization of
nanofilled compared to universal and microfilled composites. Dental Materials, 23(1),
51-59.
Bhandari, S., Aras, M., & Chitre, V. (2012). An in vitro evaluation of the microleakage under
complete metal crowns using three adhesive luting cements. The Journal of Indian
Prosthodontic Society, 12(2), 65-71.
Bitter, K., Noetzel, J., Stamm, O., Vaudt, J., Meyer-Lueckel, H., Neumann, K., & Kielbassa,
A. M. (2009). (2009). Randomized clinical trial comparing the effects of post placement
on failure rate of postendodontic restorations: preliminary results of a mean period of
32 months. Journal of endodontics, 35(11), 1477-1482.
Bryant, R. W., & Mahler, D. B. (1986). Modulus of elasticity in bending of composites and
amalgams. The J Prosthet Dent, 56(2), 243-248.
Burgoyne, A. R., Nicholls, J. I., & Brudvik, J. S. (1991). In vitro two-body wear of inlay-onlay
composite resin restoratives. The J Prosthet Dent, 65(2), 206-214.
Burke, F. T. (2012). Survival rates for porcelain laminate veneers with special reference to the
effect of preparation in dentin: a literature review. Journal of esthetic and restorative
dentistry, 24(4), 257-265.
Calamia, J. R. (1989). Clinical evaluation of etched porcelain veneers. Am J Dent, 2(1), 9-15.
Casselli, D. S. M., & Martins, L. R. M. (2006). Postoperative sensitivity in Class I composite
resin restorations in vivo. Journal of Adhesive Dentistry, 8(1), 53-58.
Castelnuovo, J., Tjan, A. H., Phillips, K., Nicholls, J. I., & Kois, J. C. (2000). Fracture load and
mode of failure of ceramic veneers with different preparations. J Prosthet Dent,, 83(2),
171-180.
Cauwels, R., Lassila, L. V., Martens, L. C., Vallittu, P. K., & Verbeeck, R. M. (2014). Fracture
resistance of endodontically restored, weakened incisors. Dental Traumatology, 30(5),
348-355.

50
Chai, S., Bennani, V., Aarts, J. M., & Lyons, K. (2018). Incisal preparation design for ceramic
veneers: A critical review. The Journal of the American Dental Association, 149(1), 25-
37.
Chen, J., Shi, C. X., Wang, M., Zhao, S. J., & Wang, H. (2005). Clinical evaluation of 546
tetracycline-stained teeth treated with porcelain laminate veneers. Journal of Dentistry,
33(1), 3-8.
Christensen, G. (2008). Thick or thin veneers? The Journal of the American Dental Association,
139(11), 1541-1543.
Clyde, J. S. (1988). Porcelain veneers: a preliminary review. Br Dent J, 9, 9-14.
Coelho de Souza, F., Gonçalves, D. S., Sales, M. P., Erhardt, M. C. G., Corrêa, M. B., Opdam,
N. J., & Demarco, F. F. (2015). Direct anterior composite veneers in vital and non-vital
teeth: A retrospective clinical evaluation. Journal of Dentistry, 43(11), 1330-1336.
Conrad, H. J., Seong, W. J., & Pesun, I. J. (2007). Current ceramic materials and systems with
clinical recommendations: a systematic review. The J Prosthet Dent, 98(5), 389-404.
Çöterta, H. S., Dündarb, M., & Öztürka, B. (2009). The effect of various preparation designs
on the survival of porcelain laminate veneers. Margin, 11(5), 405-411.
D’arcangelo, C., De Angelis, F., Vadini, M. D., & D’Amario, M. (2012). Clinical evaluation
on porcelain laminate veneers bonded with light-cured composite: results up to 7 years.
Clinical Oral Investigations, 16(4), 1071-1079.
da Costa, D. C., Coutinho, M., de Sousa, A. S., & Ennes, J. P. (2013). A meta-analysis of the
most indicated preparation design for porcelain laminate veneers. J Adhes Dent, 15(3),
215-220.
De Munck, J. D., 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. Journal of dental research, dental clinics, dental prospects, 84(2),
118-132.
Demarco, F., Corrêa, M. B., Cenci, M. S., Moraes, R. R., & Opdam, N. J. (2012). Longevity of
posterior composite restorations: not only a matter of materials. Dental Materials, 28(1),
87-101.
Dietschi D. and Spreafico R. (1997) Adhesive Metal-Free Restorations: Current concepts for
the aesthetic treatment of posterior teeth. Chicago: Quintessenc Publishing Co.

Donovan, T. E. (2008). Factors essential for successful all-ceramic restorations. The Journal of
the American Dental Association, 139(4), 14-18.

51
Dukic, W., Dukic, O. L., Milardovic, S., & Delija, B. (2010). Clinical evaluation of indirect
composite restorations at baseline and 36 months after placement. Operative dentistry,
35(2), 156-164.
Dumfahrt, H., & Schäffer, H. (2000). Porcelain laminate veneers. A retrospective evaluation
after 1 to 10 years of service: Part II--Clinical results. International J Prosthet Dent,
13(1) , 9-13
Dunne, S. M., & Millar, B. J. (1993). A longitudinal study of the clinical per- formance of
porcelain. Br Dent J, 175, 317-321.
Erkut, S., Caglar, A., Yilmaz, B., Küçükeşmen, H. C., & Ozdemir, E. (2013). Microleakage of
different provisionalization techniques for class I inlays. Journal of Dental Sciences,
8(1), 1-7.
Ferracane, J. L., & Mitchem, J. C. (2003). Relationship between composite contraction stress
and leakage in Class V cavities. American journal of dentistry, 16(4), 239-243.
Ferracane, J. L., Moser, J. B., & Greener, E. H. (1985). Ultraviolet light-induced yellowing of
dental restorative resins. J Prosthet Dent, 54(4), 483-487.
Fleming, G. j. P., Maguire, F. R., Bhamra, G., Burke, F. M., & Marquis, P. M. (2006). The
strengthening mechanism of resin cements on porcelain surfaces. Journal of dental
research, dental clinics, dental prospects, 85(3), 272-276.
Font, A., Ruiz, F. S., Ruíz, M. G., Rueda, C. L., & González, A. M. (2006). Choice of ceramic
for use in treatments with porcelain laminate veneers. Med Oral Patol Oral Cir Bucal,
11, 297-302.
Fradeani, M. (1998). Six-year follow-up with Empress veneers. International Journal of
Periodontics & Restorative Dentistry, 18(3), 216-225.
Fradeani, M., Redemagni, M., & Corrado, M. (2005). Porcelain laminate veneers: 6-to 12-year
clinical evaluation-a retrospective study. International Journal of Periodontics &
Restorative Dentistry, 25(1), 9-17.
Friedman, M. J. (1998). A 15-year review of porcelain veneer failure--a clinician's
observations. Compendium of continuing education in dentistry, 19(6), 625-628.
Garber, D. (1993). Porcelain laminate veneers: ten years later, part I—Tooth preparation. J
Esthet Dent, 5(2), 56-62.
Gerdolle, D. A., Mortier, E., Loos-Ayav, C., Jacquot, B., & Panighi, M. M. (2005). In vitro
evaluation of microleakage of indirect composite inlays cemented with four luting
agents. The J Prosthet Dent, 93(6), 563-570.

52
Geurtsen, W., & Leyhausen, G. (2001). Chemical-biological interactions of the resin monomer
triethyleneglycol-dimethacrylate (TEGDMA). Journal of dental research, dental
clinics, dental prospects, 80(12), 2046-2050.
Gilmour, A. S., & Stone, D. C. (1993). Porcelain laminate veneers: a clinical success? Dent
Update, 20(4), 167-169.
Giordano II, R. A., Pelletier, L., Campbell, S., & Pober, R. (1995). Flexural strength of an
infused ceramic, glass ceramic, and feldspathic porcelain. The J Prosthet Dent, 73(5),
411-418.
Giordano, R., & McLaren, E. A. (2010). Ceramics overview: classification by microstructure
and processing methods. Compendium of continuing education in dentistry, 31(9), 682-
684.
Gonzaga, C., Cesar, P. F., Miranda Jr, W. G., & Yoshimura, H. N. (2011). Slow crack growth
and reliability of dental ceramics. Dental Materials, 27(4), 394-406.
Gresnigt, M. M., Kalk, W., & Özcan, M. (2012). Randomized controlled split-mouth clinical
trial of direct laminate veneers with two micro-hybrid resin composites. Journal of
Dentistry, 40(9), 766-775.
Gresnigt, M., Kalk, W., & Özcan, M. (2013a). Clinical longevity of ceramic laminate veneers
bonded to teeth with and without existing composite restorations up to 40 months.
Clinical Oral Investigations, 17(3), 823-832.
Gresnigt, M., Kalk, W., & Özcan, M. (2013b). Randomized clinical trial of indirect resin
composite and ceramic veneers: up to 3-year follow-up. J Adhes Dent, 1(2), 181-190.
Gresnigt, M. M. M., Cune, M. S., Jansen, K., Van der Made, S. A. M., & Özcan, M. (2019).
Randomized clinical trial on indirect resin composite and ceramic laminate veneers: up
to 10-year findings. Journal of Dentistry, 86, 102-109.
Guess, P., Schultheis, S., Bonfante, E. A., Coelho, P. G., Ferencz, J. L., & Silva, N. R. (2011).
All-ceramic systems: laboratory and clinical performance. Dental Clinics, 55(2), 333-
352.
Guess, P. C., Selz, C. F., Voulgarakis, A., Stampf, S., & Stappert, C. F. (2014). Prospective
clinical study of press-ceramic overlap and full veneer restorations: 7-year results.
International Journal of Prosthodontics, 27(4), 355-358.
Guess, P. C., & Stappert, C. F. (2008). Midterm results of a 5-year prospective clinical
investigation of extended ceramic veneers. Dental Materials, 24(6), 804-813.
Gurel, G., Morimoto, S., Calamita, M. A., Coachman, C., & Sesma, N. (2012). Clinical
performance of porcelain laminate veneers: outcomes of the aesthetic pre-evaluative

53
temporary (APT) technique. International Journal of Periodontics & Restorative
Dentistry, 32(6), 625-635.
Guyatt, G. H., Oxman, A. D., Vist, G. E., Kunz, R., Falck-Ytter, Y., Alonso-Coello, P., &
Schünemann, H. J. (2008). GRADE: an emerging consensus on rating quality of
evidence and strength of recommendations. BMJ, 336(7650), 924-926.
Haralur, S. B. (2018). Microleakage of porcelain laminate veneers cemented with different
bonding techniques. Journal of clinical and experimental dentistry, 10(2), 166-171.
Heffernan, M. J., Aquilino, S. A., Diaz-Arnold, A. M., Haselton, D. R., Stanford, C. M., &
Vargas, M. A. (2002). Relative translucency of six all-ceramic systems. Part II: core
and veneer materials. The J Prosthet Dent, 88(1), 10-15.
Hekimoǧlu, C., Anil, N., & Yalçin, E. (2004). A microleakage study of ceramic laminate
veneers by autoradiography: effect of incisal edge preparation. Journal of oral
rehabilitation, 31(3), 265-269.
Heymamn, H. O., Sturdevant, J. R., Bayne, S., Wilder, A. D., Sluder, T. B., & Brunson, W. D.
(1991). Examining tooth flexure effects on cervical restorations: a two-year clinical
study. The Journal of the American Dental Association, 122(5), 41-47.
Highton, R., & Caputo, A. A. (1987). A photoelastic study of stresses on porcelain laminate
preparations. J Prosthet Dent, 58(2), 157-161.
Hui, K. K., Williams, B., Davis, E. H., & Holt, R. D. (1991). A comparative assessment of the
strengths of porcelain veneers for incisor teeth dependent on their design
characteristics. British dental journal, 171(2), 51-55.

Ikejima, I., Nomoto, R., & McCabe, J. F. (2003). Shear punch strength and flexural strength of
model composites with varying filler volume fraction, particle size and silanation.
Dental Materials, 19(3), 206-211.
Janda, R., Roulet, J. F., Latta, M., Steffin, G., & Rüttermann, S. (2005). Colour stability of
resin-based filling materials after aging when cured with plasma or halogen light.
European journal of oral sciences, 113(3), 251-257.
Kaizer, M. R., de Oliveira-Ogliari, A., Cenci, M. S., Opdam, N. J., & Moraes, R. R. (2014). Do
nanofill or submicron composites show improved smoothness and gloss? A systematic
review of in vitro studies. Dental Materials, 30(4), 41-78.
Karaagaclioglu, L., & Yilmaz, B. (2008). Influence of cement shade and water storage on the
final colour of leucite-reinforced ceramics. Operative Dentistry, 33(4), 386-391.
Kelly, J., & Benetti, P. (2011). Ceramic materials in dentistry: historical evolution and current
practice. Australian dental journal, 56(1), 84-96.

54
Kelly, J. R. (2008). Dental ceramics: what is this stuff anyway? The Journal of the American
Dental Association, 139(1), 4-7.
Korkut, B., Yanıkoğlu, F., & Günday, M. (2013). Direct composite laminate veneers: three case
reports. Journal of dental research, dental clinics, dental prospects, 7(2), 105-111.
Lawn, B. R., Pajares, A., Zhang, Y., Deng, Y., Polack, M. A., Lloyd, I. K., & Thompson, V. P.
(2004). Materials design in the performance of all-ceramic crowns. Biomaterials,
25(14), 2885-2892.
Layton, D., & Walton, T. (2007). An up to 16-year prospective study of 304 porcelain veneers.
International J Prosthet Dent, 20(4), 389-396.
LeSage, B. (2013). Establishing a classification system and criteria for veneer preparations
Compend Contin Educ Dent,, 34(2), 104-112 and 114-115.
Lin, B. A., Jaffer, F., Duff, M. D., Tang, Y. W., & Santerre, J. P. (2005). Identifying enzyme
activities within human saliva which are relevant to dental resin composite
biodegradation. Biomaterials, 26(20), 4259-4264.
Luthardt, R., Holzhüter, M., Sandkuhl, O., Herold, V., Schnapp, J. D., Kuhlisch, E., & Walter,
M. (2002). Reliability and properties of ground Y-TZP-zirconia ceramics. Journal of
dental research, dental clinics, dental prospects, 81(7), 487-491.
Lutz, F., Setcos, J. C., Phillips, R. W., & Roulet, J. F. (1983). Dental restorative resins. Types
and characteristics. Dental Clinics of North America, 27(4), 697-712.
Magne, P., & Belser, U. C. (2004). Novel porcelain laminate preparation approach driven by a
diagnostic mock-up. Journal of esthetic and restorative dentistry, 16(1), 7-16.
Magne, P., Kwon, K. R., Belser, U. C., Hodges, J. S., & Douglas, W. H. (1999). Crack
propensity of porcelain laminate veneers: a simulated operatory evaluation. The J
Prosthet Dent, 81(3), 327-334.
Magne, P., Perroud, R., Hodge, J. S., & Belser, U. C. (2000). Clinical performance of novel-
design porcelain veneers for the recovery of coronal volume and length. International
Journal of Periodontics & Restorative Dentistry, 20(5), 440-457.
Matsumura, H., Nakamura, M., Tanoue, N., & Atsuta, M. (2000). Clinical evaluation of an
urethane tetramethacrylate-based composite material as a prosthetic veneering agent.
Journal of oral rehabilitation, 27(10), 846-852.
McKinney, J. E., & Wu, W. (1985). Chemical softening and wear of dental composites. Journal
of dental research, dental clinics, dental prospects, 64(11), 1326-1331.
McLaren, E., & LeSage, B. (2011). Feldspathic veneers: what are their indications?
Compendium of continuing education in dentistry (Jamesburg, NJ: 1995), 32(3), 44-49.

55
Meijering, A. C., Creugers, N. H., Mulder, J., & Roeters, F. J. (1995). Treatment times for three
different types of veneer restorations. Journal of Dentistry, 23(1), 21-ì26.
Meijering, A. C., Creugers, N. H. J., Roeters, F. J. M., & Mulder, J. (1998). Survival of three
types of veneer restorations in a clinical trial: a 2.5-year interim evaluation. Journal of
Dentistry, 26(7), 563-568.
Mitra, S. B., Wu, D., & Holmes, B. N. (2003). An application of nanotechnology in advanced
dental materials. The Journal of the American Dental Association, 134(10), 1382-1390.
Moher, D., Shamseer, L., Clarke, M., Ghersi, D., Liberati, A., Petticrew, M., & Stewart, L. A.
(2015). Preferred reporting items for systematic review and meta-analysis protocols
(PRISMA-P) 2015 statement. Systematic reviews, 4(1), 1-9.
Montes-G, G. M., & Draughn, R. A. (1986). In vitro surface degradation of composites by water
and thermal cycling. Dental Materials, 2(5), 193-197.
Morimoto, S., Albanesi, R. B., Sesma, N., Agra, C. M., & Braga, M. M. (2016). Main Clinical
Outcomes of Feldspathic Porcelain and Glass-Ceramic Laminate Veneers: A Systematic
Review and Meta-Analysis of Survival and Complication Rates. International J Prosthet
Dent, 29(1), 39-49.
Moszner, N., & Klapdohr, S. (2004). Nanotechnology for dental composites. International
Journal of Nanotechnology, 1(1-2), 130-156.
Moszner, N., & Salz, U. (2001). New developments of polymeric dental composites. Progress
in polymer science, 26(4), 535-576.
Musanje, L., & Darvell, B. W. (2003). Polymerization of resin composite restorative materials:
exposure reciprocity. Dental Materials, 19(6), 531-541.
Musanje, L., & Darvell, B. W. (2006). Curing-light attenuation in filled-resin restorative
materials. Dental Materials, 22(9), 804-817.
Myers, M. L., Caughman, W. F., & Rueggeberg, F. A. (1994). Effect of restoration
composition, shade, and thickness on the cure of a photoactivated resin cement. J
Prosthet Dent, 3(3), 149-157.
Nandini, S. (2010). Indirect resin composites. Journal of conservative dentistry. JCD, 13(4),
184-194.
Nomoto, R., Uchida, K., & Hirasawa, T. (1994). Effect of light intensity on polymerization of
light-cured composite resins. Dental materials journal, 13(2), 198-205.
Okuda, M., Pereira, P. N., Nakajima, M., & Tagami, J. (2001). Relationship between
nanoleakage and long-term durability of dentin bonds. Operative dentistry, 26(5), 482-
490.

56
Örtengren, U., Andersson, F., Elgh, U., Terselius, B., & Karlsson, S. (2001). Influence of pH
and storage time on the sorption and solubility behaviour of three composite resin
materials. Journal of Dentistry, 29(1), 35-41.
Öztürk, E., & Bolay, Ş. (2014). Survival of porcelain laminate veneers with different degrees
of dentin exposure: 2-year clinical results. Journal of Adhesive Dentistry, 16(5), 481-
489.
Papadopoulos, T., Sarafianou, A., & Hatzikyriakos, A. ( 2010). Colour stability of veneering
composites after accelerated aging. European journal of dentistry, 4(2), 137-142.
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(1), 65-76.
Peumans, M., Van Meerbeek, B., Yoshida, Y., Lambrechts, P., & Vanherle, G. (1999).
Porcelain veneers bonded to tooth structure: an ultra-morphological FE-SEM
examination of the adhesive interface. Dental Materials, 15(2), 105-119.
Peutzfeldt, A. (1997). Resin composites in dentistry: the monomer systems. European journal
of oral sciences, 105(2), 97-116.
Piemjai, M., & Arksornnukit, M. (2007). Compressive fracture resistance of porcelain
laminates bonded to enamel or dentin with four adhesive systems. J Prosthet Dent,
16(6), 457-464.
Pincus, C. L. (1937). "Building mouth personality" A paper presented at: California State
Dental Association. San Jose, California.
Pini, N. P., Aguiar, F. H. B., Lima Danl Lovadino, J. R., Terada, R. S. S., & Pascotto, R. C.
(2012). Advances in dental veneers: materials, applications, and techniques. Clinical,
Cosmetic and Investigational Dentistry, 4, 9-16.
Pires de Souza, F. D. C. P., Drubi Filho, B., Casemiro, L. A., Garcia, L. D. F. R., & Consani,
S. (2009). Polymerization shrinkage stress of composites photoactivated by different
light sources. Brazilian dental journal, 20(4), 319-324.
Priyalakshmi, S., & Ranjan, M. (2014). A review on marginal deterioration of composite
restoration Journal of Dental and Medical Sciences, 13(1), 6-9.
Pröbster, L., Geis‐Gerstorfer, J., Kirchner, E., & Kanjantra, P. (1997). In vitro evaluation of a
glass–ceramic restorative material. Journal of oral rehabilitation, 24(9), 636-645.

Radz, G. M. (2011). Minimum thickness anterior porcelain restorations. Dental Clinics of North
America, 55(2), 353-370.

57
Rapani, M., Ricci, L., Rapani, C., Diomede, F., & Cardelli, P. (2015). Prefabricated Composite
Anterior Veneers: Retrospective Clinical Evaluation with a Four-Year Follow-up.
BAOJ Dentistry, 1(1), 1-4.
Reich, S. M., Wichmann, M., Rinne, H., & Shortall, A. (2004). Clinical performance of large,
all-ceramic CAD/CAM-generated restorations after three years: a pilot study. The
Journal of the American Dental Association, 135(5), 605-612.
Rekow, D., & Thompson, V. P. (2005). Near-surface damage-a persistent problem in crowns
obtained by computer-aided design and manufacturing. Proceedings of the Institution
of Mechanical Engineers, Part H. Journal of Engineering in Medicine, 219(4), 233-243.
Rinke, S., Lange, K., & Ziebolz, D. (2013). Retrospective study of extensive heat‐pressed
ceramic veneers after 36 months. Journal of esthetic and restorative dentistry, 25(1),
42-52.
Rinke, S., Pabel, A. K., Schulz, X., Rödiger, M., Schmalz, G., & Ziebolz, D. (2018).
Retrospective evaluation of extended heat-pressed ceramic veneers after a mean
observational period of 7 years. Journal of esthetic and restorative dentistry, 30(4), 329-
337.
Rosentritt, M., Esch, J., Behr, M., Leibrock, A., & Handel, G. (1998). In vivo colour stability
of resin composite veneers and acrylic resin teeth in removable partial dentures.
Quintessence International, 29(8), 517-522.
Ruiz, M. G., Font, A. F., Rueda, C. L., González, A. M., Rodríguez, J. L. R., & Ruiz, M. F. S.
(2010). A clinical longitudinal study 323 porcelain laminate veneers. Period of study
from 3 to 11 years. Medicina oral, patología oral y cirugía bucal. Ed. inglesa, 15(3),
531-537.
Sadaqah, N. R. (2014). Ceramic laminate veneers: materials advances and selection. Open
Journal of Stomatology, 4(05), 268-279.
Schmidt, K. K., Chiayabutr, Y., Phillips, K. M., & Kois, J. C. (2011). Influence of preparation
design and existing condition of tooth structure on load to failure of ceramic laminate
veneers. The J Prosthet Dent, 105(6), 374-382.
Schmitter, M. (2012). Minimally invasive lithium disilicate ceramic veneers fabricated using
chairside CAD/CAM: a clinical report. The J Prosthet Dent, 107(2), 71-74.
Seymour, K. G., Cherukara, G. P., & Samarawickrama, D. Y. (2001). Stresses within porcelain
veneers and the composite lute using different preparation designs. J Prosthodont,
10(1), 16-21.
Sheets, C. G., & Taniguchi, T. (1990). Advantages and limitations in the use of porcelain veneer
restorations. J Prosthet Dent,, 64(4), 406-411.

58
Shih, W. Y. (2016). Microleakage in different primary tooth restorations. Journal of the
Chinese Medical Association, 79(4), 228-234.
Simonsen, R. & Calamia, J (1983). Tensile bond strength of etched porcelain. Journal of Dental
Research, 62(291 (Abstr. 1 154)).
Skupien, J., Opdam, N., Winnen, R., Bronkhorst, E., Kreulen, C., Pereira-Cenci, T., &
Huysmans, M. C. (2013). A practice-based study on the survival of restored
endodontically treated teeth. Journal of endodontics, 39(11), 1335-1340.
Smales, R. J., & Etemadi, S. (2004). Long-term survival of porcelain laminate veneers using
two preparation designs: a retrospective study. International Journal of Prosthodontics,
17(3), 323-326.
Spazzin, A., Guarda, G. B., Oliveira-Ogliari, A., Leal, F. B., Correr-Sobrinho, L., & Moraes,
R. R. (2016). Strengthening of porcelain provided by resin cements and flowable
composites. Operative dentistry, 41(2), 179-188.
Sundh, A., Molin, M., & Sjögren, G. (2005). Fracture resistance of yttrium oxide partially-
stabilized zirconia all-ceramic bridges after veneering and mechanical fatigue testing.
Dental Materials, 21(5), 476-482.
Taskonak, B., Anusavice, K. J., & Mecholsky Jr, J. J. (2004). Role of investment interaction
layer on strength and toughness of ceramic laminates. Dental Materials, 20(8), 701-708.
Terry, D. A. (2004). Direct applications of a nanocomposite resin system: Part 1--The evolution
of contemporary composite materials. Practical procedures & aesthetic dentistry:
PPAD, 16(6), 417-422.
Thompson, J., Stoner, B. R., Piascik, J. R., & Smith, R. (2011). Adhesion/cementation to
zirconia and other non-silicate ceramics: where are we now? Dental Materials, 27(1),
71-82.
Turgut, S., & Bagis, B. (2011). Colour stability of laminate veneers: an in vitro study. Journal
of dentistry, 39, 57-64.
Vaidyanathan, T., & Vaidyanathan, J. (2009). Recent advances in the theory and mechanism of
adhesive resin bonding to dentin: a critical review. Journal of Biomedical Materials
Research Part B: Applied Biomaterials: An Official Journal of The Society for
Biomaterials, The Japanese Society for Biomaterials, and The Australian Society for
Biomaterials and the Korean Society for Biomaterials, 88(2), 558-578.
Wakiaga, J. M., Brunton, P., Silikas, N., & Glenny, A. M. (2004). Direct versus indirect veneer
restorations for intrinsic dental stains. Cochrane database of systematic reviews, 1.
Walls, A. W., Steele, J. G., & W, W. R. (2002). Crowns and other extra-coronal restorations:
porcelain laminate veneers. Br Dent J, 193(2), 73-76.

59
Walls, A. W. G., Murray, J. J., & McCabe, J. F. (1988). Composite laminate veneers: a clinical
study. Journal of oral rehabilitation, 15(5), 439-454.
Welbury, R. R. (1991). A clinical study of a microfilled composite resin for labial veneers.
International journal of paediatric dentistry, 1(1), 9-15.
Wendt Jr, S. L., & Leinfelder, K. F. (1990). The clinical evaluation of heat-treated composite
resin inlays. The Journal of the American Dental Association, 120(2), 177-181.
Woody, T. L., & Davis, R. D. (1992). The effect of eugenol-containing and eugenol-free
temporary cements on microleakage in resin bonded restorations. Operative dentistry,
17(5), 175-180.
Zarone, F., Epifania, E., Leone, G., Sorrentino, R., & Ferrari, M. ( 2006). Dynamometric
assessment of the mechanical resistance of porcelain veneers related to tooth
preparation: a comparison between two techniques. J Prosthet Dent, 95(5), 354-363.

60

You might also like