UNIVERSIDADE ESTADUAL PAULISTA
JÚLIO DE MESQUITA FILHO”
Instituto de Ciência e Tecnologia
open access scientific journal Volume 26 N 0 01 - 2023 | Special Edition
Campus de São José dos Campos
25th Jubilee
1998 - 20231998 - 2023
Source: macrovector / Freepik
UNIVERSIDADE ESTADUAL PAULISTA
JÚLIO DE MESQUITA FILHO”
Instituto de Ciência e Tecnologia
Campus de São José dos Campos
LITERATURE REVIEW DOI: https://doi.org/10.4322/bds.2023.e3638
1
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Wear of dental ceramics
Desgaste das cerâmicas odontológicas
Nathália de Carvalho RAMOS1,2 , Marina Gullo AUGUSTO3 , Larissa Márcia Martins ALVES4 ,
Cornelis Johannes KLEVERLAAN5 , Amanda Maria de Oliveira DAL PIVA5
1 - São Francisco University, Department of Dentistry, Bragança Paulista, SP, Brazil.
2 - University of Taubaté, Department of Dentistry, Taubaté, SP, Brazil.
3 - Centro Universitário de Cascavel, School of Dentistry, Cascavel, PR, Brazil.
4 - University of São Paulo, Bauru School of Dentistry, Department of Prosthodontics and Periodontology, Bauru, SP, Brazil.
5 - Universiteit van Amsterdam and Vrije Universiteit, Department of Dental Materials Science, Academic Centre for Dentistry
Amsterdam, Amsterdam, The Netherlands.
How to cite: RAMOS NC, AUGUSTO MG, ALVES LMM, KLEVERLAAN CJ, DAL PIVA AMO. Wear of dental ceramics. Braz. Dent. Sci.
2023;26(1):e3638. https://doi.org/10.4322/bds.2023.e3638
ABSTRACT
Based on the development of adhesive dentistry, minimally invasive restorations in ceramics are used as alternatives
to restore a tooth. Dental ceramics are largely applied in the dentistry eld mainly due to their esthetic and
mechanical strength. One of the ceramic properties to be well known before its use is the wear resistance that
should be compatible with the antagonist wear behavior to avoid unwanted performance. Therefore, several
methods have been performed to assess the ceramic materials wear behavior considering different conditions
present in the complex oral medium. This study aimed to compile the methods used to investigate dental ceramics
wear and to describe the wear mechanisms involved on them. Obtaining and analyzing data is also addressed to
discuss the results obtained from different methods, as well as the clinical analysis of wear and future perspectives
on this topic. In conclusion, many methodologies are available to measure the ceramic wear. Therefore, the
methods must be selected based on the clinical signicance of each study and should follow previously reported
parameters for standardization, allowing literature comparison.
KEYWORDS
Dental Materials; Ceramics; Dental wear; Dental restoration wear; Methods.
RESUMO
Com base no desenvolvimento da odontologia adesiva, restaurações minimamente invasivas em cerâmica são
utilizadas como alternativas para restaurar um dente. As cerâmicas odontológicas são amplamente aplicadas na
área odontológica principalmente devido à sua estética e resistência mecânica. Uma das propriedades da cerâmica
a ser bem conhecida antes de seu uso, é a resistência ao desgaste que deve ser compatível com o comportamento
de desgaste do antagonista para evitar desempenhos indesejados. Portanto, vários métodos têm sido realizados
para avaliar o comportamento do desgaste dos materiais cerâmicos considerando diferentes condições presentes no
complexo meio oral. Este estudo teve como objetivo compilar os métodos utilizados para investigar o desgaste das
cerâmicas odontológicas e descrever os mecanismos de desgaste envolvidos nos mesmos. A obtenção e análise de
dados também é abordada para discutir os resultados obtidos a partir de diferentes métodos, bem como a análise
clínica do desgaste e perspectivas futuras sobre esse tema. Em conclusão, muitas metodologias estão disponíveis para
medir o desgaste cerâmico. Portanto, os métodos devem ser selecionados com base na relevência clínica de cada
estudo e devem seguir parâmetros previamente relatados para padronização, permitindo a comparação da literatura.
PALAVRAS-CHAVE
Materiais dentários; Cerâmicas; Desgaste dentário; Desgaste de restauração dentária; Métodos.
2
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
CRITICAL REVIEW
Introduction
Wear is the damage characterized by surface
morphology changes and volume loss during
use, as a consequence of different mechanisms.
To understand each of the wear mechanisms,
it is worth mentioning that the oral medium is
complex; therefore, not only chewing is directly
acting on restorative materials surfaces, but also
food, temperature and pH variation, frequency
and force of chewing, as well as the antagonist.
Some degree of physiological tooth wear is
expected over a lifetime. The rate of progression
varies between individuals and not all tooth wear
needs treatment. Tooth wear can be dened as
pathological if it is beyond the physiological level
relative to the individual’s age and if it interferes
with the self-perception of well-being [1].
Differences in nomenclatures were previously
raised in relation to the terms linked to the wear
process in Engineering and in Dentistry also related
to the wear process of the dental structure as well
as for restorative materials evaluation. Mair (1992)
[2] presented that in engineering, fundamental
wear mechanisms are Abrasion, Adhesive wear,
Fatigue wear, Erosive wear, Corrosive wear and
Fretting wear. The authors deeply described each
mechanism and summarized that in Dentistry,
the terminology to describe wear were Attrition,
Abrasion and Erosion. Later, four wear mechanisms
have been presented as related to dental wear
processes depending on the mechanism of action:
Adhesive wear, Abrasive wear, Fatigue wear and
Corrosive wear [3,4].
Adhesive wear
results from the contact
between two surfaces and transfer from one
material surface to the other. This occurs due
to a cold welding between the material and
the antagonist, which after a certain amount
of movements will result in material loss from
one surface to another [2-4]. In addition, these
transfers can also result in particles liberation in
the medium acting now as a third-body that will
promote abrasion between the surfaces.
Abrasive wear
occurs when a hard antagonist
or particles damages the material surface. It is
divided in two- or three-body wear, according
to the presence of contact between the surfaces
(two-body) or the absence of contact but with
the presence of a third-body that will promote
abrasion [2-4]. Toothbrushing is an example of
abrasion wear in which the toothpaste acts as the
third body. Finally, attrition is the advocated term
for physical loss of mineralized tooth substance
caused by tooth to-tooth contact.
Fatigue wear
is a consequence of repeated
contact on the material that leads to crack
propagation from surface and subsurface damage.
Fatigue will occur with a defect initiation followed
by mechanical degradation until the critical load
is reached under dynamic load [2-4]. Abfraction
has been presented as an example of fatigue wear
in teeth [3].
The term erosion is largely applied in the
dental eld for the effect of a chemical agent
on a surface. However,
Corrosive wear
is the
most appropriate term to refer to the wear of a
surface that has suffered a chemical reaction that
degraded itself. In the last decades, an increase
of erosive tooth wear risk has been observed,
especially in adolescents [5,6]. The etiology of
this condition is related to extrinsic factors such
as the frequent consumption of acidic beverages,
and to intrinsic factors such as gastroesophageal
reflux and eating disorders [7]. Thus, the
erosion simulation has been largely applied
to investigate the inuence of acid agents on
dental substrates [8-13] and on the longevity of
restorative materials [2-4,14-17].
Regardless of the mechanism denition, it
is very difcult to simulate the wear present in
the oral environment. In addition, the association
of two or more mechanisms is very common to
occur, which can even difcult the denition of
the wear process origin [2-4] since attrition and
abrasion can be a consequence from all four wear
mechanisms [3]. In the end, wear is dependent on
the evaluated materials/adjacent structures and
morphology, their interaction and the medium.
Dental ceramics are widespread
materials for oral rehabilitation, with their
well-dened mechanical and wear resistance.
Lambrechts et al. [18] estimated that the
annual enamel wear rate of a molar is 38 µm,
while the average wear of a glass-ceramic is
around 0.34 mm3 per year [19]. However, it
is important to assess the wear of the material
that could lead to a long-term failure, and
to assess the wear caused on the antagonist.
Sripetchdanond et al. [20] showed that enamel
wear is 4 times greater when the antagonist is
lithium disilicate than zirconia. Tougher ceramics
as zirconia promote lower enamel wear than
glass-ceramics as lithia-based or leucite ceramics,
3
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
because ceramics with higher hardness reduce
the formation of hard three-body particles [19].
It is worth mentioning that the wear can
compromise the restorative material and the
antagonist’s morphology and also the vertical
dimension of occlusion [21]. Therefore, to
preserve mainly the natural enamel, the restorative
material is commonly selected not only due to its
properties, but also due to the promotion of less
antagonist wear. Several factors affect enamel
and restorative materials’ wear, such as hardness,
surface conditions, coefficient of friction, as
well as, microstructure factors, as presence of
porosities, voids and crystal characteristics [22].
In dental ceramics, toughness and
roughness [19] plays a role in the wear process,
since in brittle materials wear occurs by fracture.
The ceramic surface condition and ability to resist
the crack propagation predicts both its longevity
and wear potential. In a very simplied way, the
ceramic fracture involves a critical defect, which
propagates through a crack in the material until it
reaches the critical load leading to a catastrophic
fracture. These microdefects can be inherent to
the material, such as a pore or void generated in
the processing, it can be caused by an adjustment
with burs or during polishing procedures, or
even it can be caused by repeated contact on the
ceramic surface (wear facets) during chewing.
Several studies corroborated this information,
including clinical reports [23-27]. The repeated
sliding contact on the ceramic surface will lead
to compressive stresses before movement, shear
stresses at the contact interfaces, and tensile
stresses at the trailing edge of the antagonist [28]:
the fatigue wear. The fracture toughness of a
ceramic determines how much it will resist until
the crack propagates, that is, it is determined by
the size of the microstructural unit of a ceramic
that will determine the SCG (subcritical crack
growth) resistance. The crack propagation will be
favored by humidity (stress corrosion mechanism)
and affected by the material’s microstructure [29]
since the grain structure will determine the crack
growth prole [30,31].
In vitro
studies regarding the wear behavior
of dental materials are inuenced by the applied
load, length of the sliding movement, number
of cycles or time, the surface finishing of the
antagonist, testing environment, etc. There is
no consensus in literature about the testing
parameters, but commonly the “physiological
loads” are in the range of 0.4 to 75 N. However,
it’s common to find studies that used around
100-200 N load, trying to simulate the worst
case scenario, such as that observed in non-
physiological conditions, e. g. bruxism. Regarding
the horizontal sliding, this movement is around
0.5-1 mm [28,32-34] based on the occlusal
guidance amplitude.
Studies use steatite, brittle ceramics,
zirconia, stainless steel or even dental enamel
as antagonists during the tests. Steatite is a
magnesium silicate–based ceramic [34] has a
similar property to porcelain, glass-ceramics, and
dental enamel, its use has been extensively proven
in the literature and is considered the standard
for fatigue wear tests [31,32,35]. Spherical
shapes with a radius of 3 mm are generally used
to approximate the midrange of the molar cusps
radii (2-4 mm) and thus increase the clinical
relevance of the studies. Zirconia balls can also
be used in order to limit the antagonist wear, as
mentioned by Wendler et al. [28]. The contact
of the antagonist with the ceramic surface causes
an initial wear phase in which the wear scar is
properly formed (run-in stage). As these scars
increase, the steady-state wear stage is observed,
generally after 103 cycles. Thereafter, the increase
in width and depth of the wear facets is followed
by the material volume loss.
The restorative material wear is multifactorial
and changes for each oral cavity [36,37].
Therefore, wear tests aim to predict the typical
clinical wear resistance of a material. Different
wear simulation methodologies were developed
to investigate the wear behavior of different
dental materials in the long term, since wear
measurements
in vivo
are complicated and time-
consuming [8,37]. To simulate the processes that
occur in the oral medium, the wear simulation
methods cover different loads, movements,
contact type and duration, medium, presence
of food bolus simulator, toothpaste, saliva or
water [19]. Besides that, different outcomes
are used with limitations to characterize and
predict the behavior of different materials under
standardized conditions [38].
In summary, the
in vitro
tests reported in the
literature to investigate dental ceramics wear are:
Two-body-wear, three-body wear, toothbrushing
and corrosive wear. Therefore, this overview
aimed to present the wear tests methodology,
how data is obtained and how
in vivo
ceramic
wear is being investigated.
4
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
WEAR SIMULATION METHODS FOR
CERAMICS INVESTIGATION
Three-body wear test
Advocated as a sensitive way to measure
differences in the structural integrity, this
method simulates the occlusal wear mimicking
clinical conditions [39]. The main difference
in this method is the third body. Thus, the
abrasive activity of the food bolus is considered
using solid particles that are pressed against
the ceramic surface and the antagonist during
chewing [39-43]. This method has been reported
to evaluate the wear resistance of a ceramic
material or the durability of an extrinsically
characterization [39,43].
According to Lambrechts et al. [44], there
are several equipment to perform the three-body
wear, such as Oregon Health Sciences University
Oral Wear Simulator (OHSU) and University of
Alabama Wear Simulator. This review focused
on the description of the Academic Center for
Dentistry Amsterdam (ACTA) wear machine
as presented in the ISO/TS 14569 [45]. This
device consists of two motor-driven cylindrical
wheels rolling over each other (Figure 1) with
15% difference in the circumferential speed.
To standardize the method, some parameters
are predened and must be pre-checked during a
pilot study. The force between the wheels is kept
at 15 N, the rotational speed of the antagonist
wheel is 129 rpm while the speed of the specimen
wheel is 60 rpm, and both wheels rotate in
opposite directions with resulting slip rate of
15%. All this happens inside a bowl containing
the third body.
The freshly made abrasive medium (the
slurry that simulates food bolus) is used
until 200,000 cycles which is around 55 h
30 min testing [46]. After this period, the water
evaporates and the slurry is no longer properly
abrasive; therefore, it must be replaced by a
new one. The receipt (Figure 2) to standardize
the third body abrasiveness consists of: 120 g of
pandan rice (low fat white rice) grains crushed
in an electric coffee processor for 1 minute; 25 g
of millet seed shells mixed with the rice during
1 min using a blender; 1 g of a water-soluble
bacteriostatic preservative and 270 ml of buffer
solution (pH = 7) of 41.1g KH2PO4 and 9.3g NaOH
in 1L water with 1g NaN3 stabilizer. It is important
to mention that there are several types of reported
third body, limiting the comparison of data
between studies.
Figure 1 - Schematic illustration of specimen wheel, antagonist
wheel and third body movements in a three-body wear test.
Figure 2 - (A-H) Three body slurry manufacturing recipe: 120 g of pandan rice grains (A) crushed in a food processor (B and C) for 1 minute; 25
g of millet seed shells mixed with the rice during 1 min using a blender (D-F); 1 g of a water-soluble bacteriostatic preservative (G) and 270 ml
of buffer solution (pH = 7) (H).
5
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
To prepare the specimens, a wheel is used
to accommodate 10 specimens in rectangular
chambers. Prior to aligning and fixing the
specimens in each chamber with cyanoacrylate
based adhesive, the specimens must be cleaned
in an ultrasonic bath with distilled water for
5 min [47,48]. Next, to standardize all specimens
dimensions before the wear test, the specimens
need to be ground to ensure a wheel standardized
in 48 mm diameter (Figure 3). The grinding
process is performed in two steps using the ACTA
wear machine. Diamond antagonist wheels in
different grits (D 126 and D15) rotate against the
specimen wheel during 200,000 cycles.
After 24 h stored in distilled water, the
specimens are ready to be tested. The number
of cycles depends on the purpose of the study.
Usually for different restorative materials, a test
run consisted of 200,000 cycles [41,42,45]. It is
estimated that 200,000 cycles of three-body wear
cycles corresponds to 6 months of physiological
wear [45]. This number of cycles have also been
reported when investigating ceramics wear
resistance [39,43,49]. However, depending on
the evaluated material and behavior, it can vary
between 140,000 [34] and 1 × 106 cycles [39].
Before the wear test and using a prolometer
(PRK prolometer No. 720702, Perthen GmbH,
Hannover, Germany), the surface profile is
analyzed through 10 tracings in 10 xed positions.
Each reading/tracing considers 1000 measuring
points and a step distance of 100 µm between
them. After the wear run, to measure the amount
of worn ceramic surface, unworn reference
planes on both sides of the specimens should be
recorded (Figure 4). Those references are used
to calculate wear depth from the difference to
the worn surfaces [39,47]. Therefore, after each
wear run, 10 tracings are performed again in the
same predened points to determine the average
vertical loss and the standard deviations in µm of
each group by calculating the arithmetical mean
of 40 tracings [41].
Depending on the evaluated purpose, the
wear rate can also be obtained and informed in
percentage (%) based on the average thickness
for a known layer (100%) that has been removed
during the test, e. g., an external characterization
layer [39] (Figure 5).
Two-body wear tests
The two-body simulates the non-masticatory
tooth movement and through it is possible to
predict the wear behavior of dental materials [44].
The main differences between three- and two-
body wear tests is that the two-body wear test has
contact between specimen and antagonist [48,49].
In the literature, it is possible to observe different
methods for two body wear, traditionally using
a sliding methodology, where the materials are
tested in pairs under nominally non-abrasive
conditions [44]. For example, to evaluate the
ceramic nishing protocol (glaze or polishing)
on the antagonist wear [50,51].
Figure 3 - (A-C) Specimen wheel preparation. Schematic illustration of (A) specimens glued in the specimen wheel, (B) grounded specimens
during preparation to obtain a (C) perfect circular wheel sample.
Figure 4 - Schematic illustration of wear analysis showing the different coordinate axes and the wheel position during the measurement.
6
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
To evaluate the two-body wear behavior of
dental ceramics, previous studies used a pin-on-
disk apparatus [52,53]. For that, polished discs
samples (2 mm thickness) are scanned individually
before the wear test by a scanner (e.g. CEREC
AC Omnicam, Sirona, Brazil) or a prolometer
(e.g. CyberSCAN CT 100; CyberTECHNOLOGIES
GmbH, Eching-Dietersheim, Germany). Then,
the samples must be embedded in acrylic resin,
with the nished surface up, using a plastic mold.
The pin-on-disk is a method widely applied, for
that a pin with a radiused tip as antagonist is
positioned perpendicular to the sample, usually
a at disc (Figure 6). The test machines apply a
movement, either the pin or the disc can slide,
reproducing the wear path [51]. The movement
also can be revolving. The pin is pressed against the
sample at a specied load (the literature reported
a range from 5 N to 30 N) [54,55] by hydraulic
or pneumatic systems and also by weights [56].
The antagonists can vary in composition.
Natural enamel is required to achieve clinical
conditions, however, different morphologies lead
to variations on the results [35]. As an alternative,
steatite is used for
in vitro
wear studies to
allow test standardization and a feasible results
analysis [33,57]. Stainless steel, dental porcelain
and alumina are also materials applied to assess
materials wear resistance [35,44].
The lubrication is necessary in order to
remove residues and also to mimic the oral
environment. For that, the tests can be performed
in distilled water at room temperature [50],
in 33% glycerin lubricant [58] or in those two
combined [59]. The literature also reports the
use of normal saline emulsion [60] or articial
saliva [61,62].
In addition to the load and the lubrication
mentioned above, the two-body wear methodology
involves other parameters such as frequency,
sliding distance and number of cycles. Chewing
activity mainly occurs in the range of 0.94 Hz and
2.17 Hz [63]. However, in order to reduce time-
consuming, studies use a higher cyclic frequency,
being possible to nd in the literature a range
from 1,7 to 30 Hz [55,64]. Regarding the number
of cycles, 1250,000 correspond to one year of
clinical masticatory effort [65] and, as well as
the previous parameters, it is possible to nd a
range in the literature, e.g. 120,000 cycles [66]
and also ten times greater than that [33].
Considering mandibular movements, closing
during chewing promotes occlusal contact
between teeth or restorative material and
teeth [67-69]; in addition, the sliding movement
is also an essential component to wear tests
since it promotes material micro fatigue [70].
Therefore, sliding movement is applied from
0,3 mm to 15 mm [50,60,71].
The wear is measured by volume loss (mm3)
and wear depth (mm). Although the greater
accuracy, volume loss is less reported than wear
depth once clinically a tooth height loss is easier
to visualize than volumetric tooth loss [33]. Linear
measures are obtained as the length of the pin
changes; and, to assess the disc wear, prolometers
or scanners [53,64] can be used [55]. Besides
that, wear can be quantied by weighing both
specimens before and after the test [50,51,72].
Another methodology to study wear-
resistant materials is the ball-on-flat. This
method applies the same parameters as described
above for the pin-on-disk method, however, in
this case, a spherically specimen slides against
the flat disc, resulting in hertzian contact
pressure [73] (Figure 7). The antagonist can
be of different materials, such as zirconia (r =
6,35 mm), steatite (r = 3mm) and alumina (r
= 3,17 mm) [33,74,75]. The sphere slides or
rotates against the polished at surface at a xed
frequency (~ 0,4 2 Hz) delivering a specic
load (10N - 200N) [33,75]. The same parameters
mentioned before are followed for lubricant
medium and number of cycles.
For the ball-on-at specimens the wear can
be assessed by topographical reconstruction.
Figure 5 - Curve fit for the wear rate vs. time according to the
three-body wear test and different ceramic materials (YZHT = high
translucent yttrium-stabilized tetragonal zirconia polycrystal, FDL =
Feldspar ceramic, ZLS1 = reinforced glass ceramic stained in 1 step,
ZLS2 = reinforced glass ceramic stained in 2 steps and HC = Hybrid
ceramic).
7
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
The volume loss (mm3) and wear depth (mm) can
be quantied with 3-dimensional (3D) images
obtained by a micro– computed tomography
scanner (micro-CT 40; Scanco Medical AG) [33]
or a laser scanner (SD Mechatronic Laser Scanner
LAS-20) [76]. The collected data is analyzed
using a 3D reconstruction software (3D-System
Geomatic Wrap). By means of mesh editing and
model superimposition, the maximum wear
depth and volume loss can be obtained by model
comparison before and after wear [76-78].
Another two-body wear method test will
be described using the ACTA wear machine, as
illustrated in Figure 8. For this test, the specimen
wheel is prepared as described for the three-
body wear test above (Figure 3). In addition,
the machine works with the same parameters
presented for the three-body wear test (spring
force of 15 N, 200,000 cycles and rotation speed
of 1 Hz). As well as, the average vertical loss
can be determined according to the difference
between the un-worn lateral references and the
worn surface (Figure 4), where the abrasive
wheel contacted [49]. In the two body wear
test, the antagonist or abrasive wheel is in direct
contact with the specimen wheel and can be
made in different materials, such as: stainless
steel [39,43,49], ceramics [48], enamel [48] and
resin composite [46].
Toothbrushing
Toothbrushing simulation is widely used to
evaluate the abrasion wear of ceramic materials,
as well as, the wear of the glaze or the extrinsically
stain layer [79-84]. However, depending on how
and when toothbrushing is performed, as well as
the type of dentifrice and toothbrush used [85],
toothbrushing frequency and force of brushing,
Figure 6 - Operating diagram of the two-body wear machine testing a pin against a disc specimen.
Figure 7 - Schematic illustration of the two-body wear machine testing a ball against a specimen.
Figure 8. Schematic illustration of specimen wheel and antagonist
wheel in a two-body wear test.
8
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
dental wear may increase [14]. Some studies
reported the negative inuence of toothbrush
on color stability of extrinsic characterized glass
ceramics [81,86] also, with the association with
thermocycling [82] for lithium disilicate and
zirconia.
The number of toothbrushing cycles
needed to simulate the toothbrush wear
varies between different studies, ranging from
7300 strokes [87,88] to 150.000 strokes [84].
In a pilot clinical study, it was found that the
average person brushes between 25-30 cycles
per day on a given surface [89]. This equates to
9125 to 10950 cycles per year, thus, it can be
considered that 10.000 strokes simulates 1 year
of toothbrushing. In addition, 50,000, 100,000,
and 150,000 cycles have been advocated [82] as
proportionally corresponding to approximately
5, 10, and 15 years of brushing in the oral
environment.
Previous studies recommend replacing
the toothbrushes every 10,000 [90] to
50,000 strokes [91]. Another parameter that
can influence the toothbrush test is the load.
In habitual toothbrushing, force peaks reaching
10 N can occur, however, the mean brushing force
was found to be 2.3 ± 0.7 N (max. 4.1 N) [92],
being recommended for in vitro studies to adopt
a representative load of 2 N to 3N.
The toothbrush bristle arrangement also plays
a role in the results, being the ordinary/at-trimmed
toothbrushes more abrasive than the feathered
ones [93]. The ISO/TR 14569-1:2007 [94]
provides guidelines for test methods for the
assessment of resistance to wear by toothbrushing.
This standard establishes that soft toothbrushes
with nylon bristles and rounded tips should be
used. Moreover, it is recommended that the
toothbrush be angled at 15° in relation to the
direction of brushing to minimize the formation of
grooves on the specimen’s surface [95] (Figure 9).
The toothpaste abrasiveness can signicantly
inuence the loss of color and gloss of ceramic
restorations [96]. The abrasiveness of the
toothpastes is usually determined by the REA
(relative enamel abrasiveness) or RDA (relative
dentine abrasiveness). These values compare
the abrasiveness of the tested toothpaste to a
standard one, given a score of 10 to 100. As sound
dentin is considered more susceptible to abrasion
than enamel, the RDA value has become the
main parameter to characterize the abrasiveness
of the toothpastes. Unfortunately, most of the
toothpaste manufacturers do not refer to the RDA
and REA values.
For the toothbrush simulation, a slurry of
the toothpaste and deionized water should be
prepared in a ratio of 1:3 (w/w) [83,84,97]
(Figure 10). The main purpose of the slurry is
to obtain a consistency that better simulates the
conditions in the oral cavity. It is possible to nd
studies using a 1:2 (w/w) [96,98] or 1:1 (w/w)
ratio [81], however a thick slurry can inuence
the results, being necessary to standardize the
ratio to allow comparisons among studies. When
using brushing simulation machines, 1mL of the
suspension should be injected onto the specimen
surfaces every 30 seconds [99].
When the study includes enamel and/or
dentin specimens, the toothpaste slurry should
be prepared by using articial or human saliva
instead of deionized water. This substitution is
recommended due to the remineralizing potential
of these specimens when in contact with saliva
and uorides from the toothpaste [100].
After the toothbrushing simulation, it is
recommended to clean the specimens by an
Figure 9 - (A-C) Schematic illustration of a toothbrushing simulation. (A) Intact surface, (B) Toothbrush head positioning and brush directions
and the metallic device positioned over the specimen. (C) Inspection path of the wear profile after the toothbrushing simulation.
9
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
ultrasonic bath for 10 minutes [101]. This
procedure will remove surface debris, allowing
comparison between specimens. The amount of
wear in micrometers can be obtained by contact
prolometer by using the Rz parameter [83,84,99]
or a dedicated software [97]. In addition,
the wear values can be obtained by 3D laser-
measuring microscopy [102] and vertical loss of
dental cusp [103].
To standardize the method, two parallel
grooves can be marked at the lateral area of the
specimen’s surface as reference, allowing the
superimposition of the initial and nal surface
proles (Figure 9C). A metal strip containing an
orice can be used to protect these grooves, so
only the central area will be brushed [83,84,97]
(Figure 9B). A custom-made device is used to
place the specimens in the same position before
and after brushing. The depth of the abraded
area can be calculated based on the subtraction
of the initial prole from the nal prole by using
a dedicated software program [97]. In addition,
the specimen’s surface prole can be exported in
linear graphs (Figure 11) for illustration [83,84]
Corrosive wear
Considering corrosive wear on ceramics, the
scientic data is scarce. It is worth mentioning
that the evaluation of a ceramic wear corrosion
investigates the material abrasiveness since it
enhances the material roughness leading to the
wear of the antagonist [75,104]. In addition,
corrosion wear could compromise the material
mechanical properties; since different from teeth,
they do not have any self-healing mechanisms [75].
Švančárková et al. [75] evaluated the corrosive
wear of lithium disilicate using the ball-on-at
two-body wear test after two different corrosion
simulations (quasi-dynamic with two corrosive
media or static according to the ISO 6872) [105].
According to the literature, the corrosive wear or
chemical durability can be measured by weight
loss as well as the concentration of leached ions
into the corrosive medium [75,104,106].
Considering the erosion simulation
protocols, different acidic agents have
been reported to affect the ceramic surface
roughness and morphology [107,108], color
stability [106], hardness [105,109], flexural
strength [104] fracture toughness [110] or ion
leaching [75,108]. In addition, ISO 6872 [104]
indicates 4% acetic acid; while different chemical
agents have been used, such as: citrate buffer
solution, juices [108,110]; citric acid and lactic
acid [105,106,111,112]; simulated gastric HCl,
white wine, soda drink [110]; or acidulated
fluoride mouthwash solutions [113], during
different times of exposure; which makes the
results comparison difficult. Therefore, this
overview strongly recommends a standardization
for solutions and the use of profilometry to
Figure 10 - Toothpaste solution being prepared for the toothbrushing simulation.
Figure 11 - Surface profiles of a hybrid ceramic before and after 15
years of toothbrushing simulation.
10
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
investigate the wear rate and material volume
loss, as performed before [7].
RESULTS INTERPRETATION
During wear tests, an extended damage
accumulation zone is formed on the ceramic
surface, with defects that will be initiators and
will lead to crack growth [23]. Therefore, fatigue
degradation is closely related to wear. Repeated
contact causes subsurface cone cracks, which are
initiated by tensile stress, but grow chemically-
assisted by water (stress corrosion) mechanism.
The ceramic surface morphology, density and
also crack-surface angles are closely related to
the friction coefcient at the interface and the
fracture toughness of the material undergoing this
cyclic fatigue. Materials with higher toughness
such as zirconia will exhibit enhanced resistance
to crack propagation and lower wear, while
glassy matrix ceramics (lithium disilicate and
feldspathic, for example) will be more prone
to a propagation of partial cone cracks during
sliding contact fatigue and will present higher
wear rate [31,114,115]. Wendler et al. [28]
analyzed the sliding fatigue wear of ve different
CAD/CAM ceramics (IPS e.max CAD from Ivoclar
Vivadent, Suprinity partially crystallized, Enamic
and Vitablocs Mark II from Vita Zahnfabrik, and
Lava Ultimate from 3M Espe) against zirconia
indenters (antagonist). Their results showed
that glass ceramics (e.max CAD, Suprinity and
Vita Mark II) have a greater zone of subsurface
damage regarding the fatigue wear mechanism,
while composite materials as Lava Ultimate do
not have this subsurface damage area, but show
greater wear facets due to predominant abrasion
effect. Hybrid ceramic materials such as Enamic,
on the other hand, show a combination of the
above, with large surface wear scars associated
with greater subsurface damage.
Comparing similar material using different
wear methods, it is possible to affirm that
different wear methods promote different surface
patterns [49,116]. Surface roughness has been
reported as an important factor on the beginning
of the wear process; however, the ceramic
material microstructure has the most important
effect in the wear rate. This is due to the fact that
surface roughness is immediately changed during
the wear procedure. Therefore, wear resistance
is influenced by the grains arrangement and
materials hardness and composition [49,117].
However, after any of the mentioned methods,
the surface morphology or surface roughness must
be evaluated due to its effect on the materials
properties, such as exural strength [118,119],
wettability [50,120], survival rate [121], as well
as, on the biolm formation [120,122,123].
At least two wear tests are advocated
to characterize the material wear resistance.
However, different sets do not allow direct
comparison or equivalent wear behavior [116].
Two-body wear presents higher wear depth due
to the contact between specimens and antagonist,
while three-body wear presents less abrasion
promoted by the third-body. Normally, the third
body contains soft particles and not all of them
cause wear [3], being also reported as damping
agents [19]. In the two-body wear, the wear is
much higher then the shape of the evaluated
material is modified according to the shape
of the hard antagonist [2,53,74]. Even if the
similarity between highest wear depth provided
by different tests is reported [49], their purpose of
investigation is different. Therefore, implications
based on their results must be done considering
the test’s characteristics.
Minimal variation between the
in vitro
methods [45] are necessary to easily detect
differences between materials. In addition,
clinical studies are essential for the evaluation
of material performance in function. However,
they are complex, involving ethical issues, time
consuming, dependent on patient collaboration,
but crucial to validate
in vitro
methods [19].
SEM as a complementary analysis
After the wear simulations and material loss
analysis, the worn surfaces are submitted to a surface
analysis using a Scanning Electron Microscope
at different magnifications (Figure 12, 13)
[39,43,50,51,72,75,97]. This analysis is used for
a qualitative measure of the caused damage or
to observe the surface morphology after a wear
process. For that, the specimens can be directly
investigated [39,43,49,83,84,86,107,124] or
indirectly through an impression procedure
using polyvinyl siloxane and poured epoxy resin
inside the molde [48]. Therefore, materials
also can be tested in complex geometries, as
tooth restorations [33,77,78]. In this case, to
analyze the progressive wear, impressions can
be obtained to manufacture stone cast models
(Type IV gypsum). The casts should be scanned
11
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
as a baseline and sequentially as the wear circles
intervals. After that, the 3D images are analyzed
in a software program (e.g. GOM Inspect;
GOM) and the volume loss can be measured
by overlapping the images [77], allowing the
analysis in different intervals.
With direct investigation, the specimens
should be cleaned and placed on aluminum
stubs, sputter coated with gold and observed
in a scanning electron microscope [97]. Cross-
sectioned specimens can also be used to analyze
subsurface damage. Generally the specimens
are included in epoxy or acrylic resin after the
wear simulations, a sagittal or transverse cut
can be made in the direction of the wear scar,
so this cross-section is polished with diamond
pastes [125]. Using this type of method, it is
possible to observe the partial cone cracks for the
subsurface damage evaluation [28].
Clinical assessment of ceramic wear
The fast development of different intraoral
scanners has allowed clinicians to make early
diagnosis of dental wear [126-128]. In addition,
scanners allow the impression procedure to be
more comfortable, easier and faster, making
the treatments more practical, comfortable
and agile [129]. The surface wear is analyzed
according to the volume change, maximum and
average prole losses, according to quantitative
monitoring of clinical wear [126,130] (Figure 14).
Epoxy resin models from conventional impressions
have also been reported as an alternative for wear
measurement [131].
The quantitative maximum vertical substance
loss is calculated as a difference between the
rst scanning and the second scanning during
follow-up (Figure 15). Initially, the three-
dimensional models are superimposed using a 3D
analysis software, e.g. GOM Inspect [126,127],
Geomagic, David-Laserscanner [132], or
softwares used in the dental practice, such as
Trios Patient Monitoring tool. After importing the
baseline (initial) model and the second scanning
(or model with wear), it is necessary to perform
the virtual alignment of both. For that, the best-
t alignment can be used when the models are
strongly compatible. Or, a three-point alignment
in which geometric references are pre-determined.
The software choice and the comparison mode
as well as the models alignment depend on the
complexity of each case. The superimposition of
Figure 12 - Surface topography under SEM analysis with 5000 × magnification of the different materials after the three-body wear test (YZHT
= high translucent yttrium-stabilized tetragonal zirconia polycrystal, FDL = Feldspar ceramic, ZLS1 = reinforced glass ceramic stained in 1 step,
ZLS2 = reinforced glass ceramic stained in 2 steps and HC = Hybrid ceramic).
Figure 13 - Scanning electron microscopy images, ×3000, after artificial toothbrushing of (from left to right) a leucite ceramic, a hybrid ceramic
and a felspathic ceramic.
12
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
models is dened as Parametric Inspection and
calculates wear depth values in mm. Besides
the comparison of wear depth, most of the
time the calculation of the material volume loss
is necessary. For that, the volume difference
can be quantified using other software, e.g.,
Materialise [133], Siemens Unigraphics NX
10 [132] or the internal monitoring program in
the scanner [126,134].
Schlenz et al. [127] have investigated the
wear process in natural dentition and observed,
after 12 months follow-up, a mean loss at cusps
ranging between 31 and 43µm in young adults
(18-25 years). This reinforces the need for
monitoring and allowing early diagnosis of wear.
The authors reinforce that the digital impression
accuracy plays an important role in the tissue loss
evaluation in a micrometer scale. However, the
difculty in dening references
in vivo
, concerns
regarding the surface alignment [134] and results
interpretation.
Dental ceramics are not very susceptible
to wear, and because of their wear rates close
to the natural enamel, ceramics may be a good
restorative alternative for oral rehabilitation [49].
An
in vitro
study evaluated the wear of implant-
supported crowns in different materials (lithium
disilicate, zirconia, hybrid ceramic and porcelain
fused to metal) before and after 5-years chewing
simulation. The crowns were evaluated using
laboratorial and intraoral scanners [133]
and there was no difference between both to
detect the ceramic volume loss. Differently,
Aladağ et al. [132] during an
in vivo
study have
found, after 6 months, that hybrid ceramic shows
higher mean wear value (0.38mm3) compared
to lithium disilicate (0.27mm3). In addition, the
authors evaluated crowns in zirconia reinforced
lithium silicate (0.14mm3) and a resin matrix
ceramic material (0.45mm3). The authors have
precisely described their evaluation method,
which were the following steps: Digital impression
of the restorations, antagonist tooth, adjacent
teeth, interocclusal registration and occlusal
contact points using the scanner software (Cerec
4.2, Sirona, Bensheim, Germany), followed
by the generation of .stl files; Next, the files
were exported for superimposition (David-
Laserscanner, V3.10.4, Berlin, Germany). Then,
using the third software (Siemens Unigraphics
NX 10, Siemens PLM Software, Plano, TX, USA),
the images were converted into solids and a
specic area was chosen for analysis (0.005 mm
tolerance). Finally, the difference between them,
as volume loss, was calculated.
Future perspectives
Advances in dental science enable different
ceramics to be used for various indications. And,
mainly due to their esthetic and strength, the use of
ceramic restorations have increased. The assessment
Figure 14 - Schematic illustration of tooth wear evaluation using digital impression and superimposition method. From left to right, initial
scanning, scanning after a period in function and superimposition in forming wear depth in mm, according to the colorimetric scale.
Figure 15 - Schematic illustration of a model presenting the
difference between the first scanning and the second scanning for
monitoring through colorimetric scale.
13
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
of ceramic wear using different simulations provide
information regarding the restorative material
behavior under different conditions. The materials
assessment, especially materials in development or
recently-launched is very important to understand
how the material will behave during function.
However, the simplication in the procedures are
necessary for studies standardization and data
assessment and comparison. The reproducibility
between methodologies must be possible to
decrease the variability in obtained data and to
direct clinical studies.
In vitro
investigations present limitations
while they try to simulate the oral medium.
It is important to mention that only one method
cannot provide universal data since the method
does not simulate all factors present inside
the oral cavity [44,134], e.g., saliva, different
loads, pH and temperature variation, different
brushing devices, food textures, brushing and
chewing frequencies, diets, use of mouth risings,
etc. In addition, the complexity of the wear
process is very difcult to simulate [4,134,135].
Therefore, the results obtained from different
in vitro
methods must be evaluated carefully.
And therefore, to complement the
in vitro
studies,
further clinical reports and studies are advocated
with patients monitoring tools to provide data
regarding ceramic materials wear resistance
during function. This study has the limitation of
considering the most common
in vitro
methods
to investigate different ceramic wear behavior.
Considerations
With the limitation of this study, the
following conclusion can be drawn:
Different methodologies are available to
simulate wear in ceramic restorative materials.
Most of the time to evaluate the ceramic wear
resistance, but also to investigate the glaze and/
or the external characterization wear rate, or even
its effect on the antagonist wear rate. However,
it is important to select carefully the method for
each investigation, based on the main purpose
of the study and focusing on the answer that the
methods can provide. In addition, the appropriate
data collection should be performed to provide
information equivalent to the evaluated topic.
Finally, the tests should follow similar parameters
for standardization and to allow comparison
between different studies.
Author’s Contributions
NCR, MGA, LMMA: Methodology,
Investigation, Writing – Original Draft Preparation,
Writing – Review & Editing, Visualization.
CJK: Investigation, Writing – Original Draft
Preparation, Writing – Review & Editing,
Visualization, AMODP: Conceptualization,
Methodology, Investigation, Writing – Original
Draft Preparation, Writing – Review & Editing,
Visualization, Supervision, Project Administration.
Conict of Interest
The authors have no proprietary, nancial,
or other personal interest of any nature or kind
in any product, service, and/or company that is
presented in this article.
Funding
This research did not receive any specic
grant from funding agencies in the public,
commercial, or not-for-prot sectors.
Regulatory Statement
Not applicable.
REFERENCES
1. Schlueter N, Amaechi BT, Bartlett D, Buzalaf MAR, Carvalho TS,
Ganss C,etal. Terminology of erosive tooth wear: consensus
report of a workshop organized by the ORCA and the cariology
research group of the IADR. Caries Res. 2020;54(1):2-6. http://
dx.doi.org/10.1159/000503308. PMid:31610535.
2. Mair LH. Wear in dentistry--current terminology. J Dent.
1992;20(3):140-4. http://dx.doi.org/10.1016/0300-
5712(92)90125-V. PMid:1624617.
3. Tsujimoto A, Barkmeier WW, Fischer NG, Nojiri K, Nagura Y,
Takamizawa T,etal. Wear of resin composites: current insights
into underlying mechanisms, evaluation methods and influential
factors. Jpn Dent Sci Rev. 2018;54(2):76-87. http://dx.doi.
org/10.1016/j.jdsr.2017.11.002. PMid:29755618.
4. Dionysopoulos D, Gerasimidou O. Wear of contemporary dental
composite resin restorations: a literature review. Restor Dent
Endod. 2021;46(2):e18. http://dx.doi.org/10.5395/rde.2021.46.
e18. PMid:34123754.
5. Schlueter N, Luka B. Erosive tooth wear - a review on global
prevalence and on its prevalence in risk groups. Br Dent J.
2018;224(5):364-70. http://dx.doi.org/10.1038/sj.bdj.2018.167.
PMid:29495027.
6. Agulhari MAS, Giacomini M, Rios D, Bombonatti J, Wang L.
Giomer technology for preventive and restorative clinical
management of erosive tooth wear: a case report. Braz Dent Sci.
2022;25(2):e3162. http://dx.doi.org/10.4322/bds.2022.e3162.
7. Jaeggi T, Lussi A. Prevalence, incidence and distribution
of erosion. Monogr Oral Sci. 2014;25:55-73. http://dx.doi.
org/10.1159/000360973. PMid:24993258.
14
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
8. Bueno TL, da Silva TA, Rizzante FA, Magalhães AC, Rios D,
Honório HM. Evaluation of Proanthocyanidin-based dentifrices
on dentin-wear after erosion and dental abrasion - In situ study.
J Clin Exp Dent. 2022;14(4):e366-70. http://dx.doi.org/10.4317/
jced.59071. PMid:35419173.
9. Frattes FC, Augusto MG, Torres CRG, Pucci CR, Borges AB.
Bond strength to eroded enamel and dentin using a universal
adhesive system. J Adhes Dent. 2017;19(2):121-7. http://dx.doi.
org/10.3290/j.jad.a38099. PMid:28439576.
10. Belmar da Costa M, Delgado AHS, Pinheiro de Melo T, Amorim T,
Mano Azul A. Analysis of laboratory adhesion studies in eroded
enamel and dentin: a scoping review. Biomater Investig Dent.
2021;8(1):24-38. http://dx.doi.org/10.1080/26415275.2021.188
4558. PMid:33629074.
11. Ferretti MA, Theobaldo JD, Pereira R, Vieira-Junior WF,
Ambrosano GMB, Aguiar FHB. Effect of erosive challenge and
cigarette smoke on dentin microhardness, surface morphology
and bond strength. Braz Dent Sci. 2020;23(3):1-8. http://dx.doi.
org/10.14295/bds.2020.v23i3.2041.
12. Augusto MG, da Silva LFO, Scaramucci T, Aoki IV, Torres CRG,
Borges AB. Protective effect of anti-erosive solutions enhanced
by an aminomethacrylate copolymer. J Dent. 2021;105:103540.
http://dx.doi.org/10.1016/j.jdent.2020.103540. PMid:33249109.
13. Mailart MC, Berli PC, Borges AB, Yilmaz B, Baumann T,
Carvalho TS. Pellicle modification with natural bioproducts:
influence on tooth color under erosive conditions. Eur J Oral
Sci. 2022;130(5):e12886. http://dx.doi.org/10.1111/eos.12886.
PMid:35839337.
14. Wiegand A, Schlueter N. The role of oral hygiene: does
toothbrushing harm? Monogr Oral Sci. 2014;25:215-9. http://
dx.doi.org/10.1159/000360379. PMid:24993269.
15. Schlichting LH, Resende TH, Reis KR, Magne P. Simplified
treatment of severe dental erosion with ultrathin CAD-CAM
composite occlusal veneers and anterior bilaminar veneers. J
Prosthet Dent. 2016;116(4):474-82. http://dx.doi.org/10.1016/j.
prosdent.2016.02.013. PMid:27132785.
16. Attin T, Wegehaupt FJ. Impact of erosive conditions on tooth-
colored restorative materials. Dent Mater. 2014;30(1):43-9.
http://dx.doi.org/10.1016/j.dental.2013.07.017. PMid:23962494.
17. Maia KMFV, Rodrigues FV, Damasceno JE, da Costa Ramos RV,
Martins VL, Lima MJP,etal. Water sorption and solubility of a
nanofilled composite resin protected against erosive challenges.
Braz Dent Sci. 2019;22(1):46-54. http://dx.doi.org/10.14295/
bds.2019.v22i1.1660.
18. Lambrechts P, Braem M, Vuylsteke-Wauters M, Vanherle
G. Quantitative in vivo wear of human enamel. J Dent Res.
1989;68(12):1752-4. http://dx.doi.org/10.1177/002203458906
80120601. PMid:2600255.
19. Branco AC, Colaço R, Figueiredo-Pina CG, Serro AP. A state-of-
the-art review on the wear of the occlusal surfaces of natural
teeth and prosthetic crowns. Materials (Basel). 2020;13(16):3525.
http://dx.doi.org/10.3390/ma13163525. PMid:32785120.
20. Sripetchdanond J, Leevailoj C. Wear of human enamel opposing
monolithic zirconia, glass ceramic, and composite resin: an in
vitro study. J Prosthet Dent. 2014;112(5):1141-50. http://dx.doi.
org/10.1016/j.prosdent.2014.05.006. PMid:24980740.
21. Laborie M, Naveau A, Menard A. CAD-CAM resin-ceramic
material wear: a systematic review. J Prosthet Dent. 2022
Apr;S0022-3913(22)00076-2. http://dx.doi.org/10.1016/j.
prosdent.2022.01.027.
22. Oh WS, Delong R, Anusavice KJ. Factors affecting enamel
and ceramic wear: a literature review. J Prosthet Dent.
2002;87(4):451-9. http://dx.doi.org/10.1067/mpr.2002.123851.
PMid:12011863.
23. Kelly JR, Cesar PF, Scherrer SS, Della Bona A, van Noort R, Tholey
M,etal. ADM guidance-ceramics: fatigue principles and testing.
Dent Mater. 2017;33(11):1192-204. http://dx.doi.org/10.1016/j.
dental.2017.09.006. PMid:29017761.
24. Lohbauer U, Krämer N, Petschelt A, Frankenberger R. Correlation
of in vitro fatigue data and in vivo clinical performance of a
glassceramic material. Dent Mater. 2008;24(1):39-44. http://
dx.doi.org/10.1016/j.dental.2007.01.011. PMid:17467049.
25. Scherrer SS, Quinn JB, Quinn GD, Wiskott HW. Fractographic
ceramic failure analysis using the replica technique. Dent
Mater. 2007;23(11):1397-404. http://dx.doi.org/10.1016/j.
dental.2006.12.002. PMid:17270267.
26. Guess PC, Zavanelli RA, Silva NR, Bonfante EA, Coelho PG,
Thompson VP. Monolithic CAD/CAM lithium disilicate versus
veneered Y-TZP crowns: comparison of failure modes and
reliability after fatigue. Int J Prosthodont. 2010;23(5):434-42.
PMid:20859559.
27. Sailer I, Gottnerb J, Kanelb S, Hammerle CH. Randomized
controlled clinical trial of zirconia-ceramic and metal-ceramic
posterior fixed dental prostheses: a 3-year follow-up. Int J
Prosthodont. 2009;22(6):553-60. PMid:19918588.
28. Wendler M, Kaizer MR, Belli R, Lohbauer U, Zhang Y. Sliding
contact wear and subsurface damage of CAD/CAM materials
against zirconia. Dent Mater. 2020;36(3):387-401. http://dx.doi.
org/10.1016/j.dental.2020.01.015. PMid:32007314.
29. Zhang Y, Sailer I, Lawn BR. Fatigue of dental ceramics. J Dent.
2013;41(12):1135-47. https://doi.org/10.1016/j.jdent.2013.10.007.
30. Lawn BR, Borrero-Lopez O, Huang H, Zhang Y. Micromechanics
of machining and wear in hard and brittle materials. J Am Ceram
Soc. 2021;104(1):5-22. http://dx.doi.org/10.1111/jace.17502.
PMid:34565803.
31. Ramos NC, Campos TM, Paz IS, Machado JP, Bottino MA, Cesar
PF, etal. Microstructure characterization and SCG of newly
engineered dental ceramics. Dent Mater. 2016;32(7):870-8.
http://dx.doi.org/10.1016/j.dental.2016.03.018. PMid:27094589.
32. El Zhawi H, Kaizer MR, Chughtai A, Moraes RR, Zhang Y. Polymer
infiltrated ceramic network structures for resistance to fatigue
fracture and wear. Dent Mater. 2016;32(11):1352-61. http://
dx.doi.org/10.1016/j.dental.2016.08.216. PMid:27585486.
33. Kaizer MR, Moraes RR, Cava SS, Zhang Y. The progressive
wear and abrasiveness of novel graded glass/zirconia
materials relative to their dental ceramic counterparts. Dent
Mater. 2019a;35(5):763-71. http://dx.doi.org/10.1016/j.
dental.2019.02.022. PMid:30827797.
34. Kaizer MR, Bano S, Borba M, Garg V, Dos Santos MBF, Zhang
Y. Wear behavior of graded glass/zirconia crowns and their
antagonists. J Dent Res. 2019;98(4):437-42. http://dx.doi.
org/10.1177/0022034518820918. PMid:30744472.
35. Shortall AC, Hu XQ, Marquis PM. Potencial countersample
materials for in vitro simulation wear testing. Dent Mater.
2002;18(3):246-54. http://dx.doi.org/10.1016/S0109-
5641(01)00043-4. PMid:11823017.
36. Heintze SD, Barkmeier WW, Latta MA, Rousson V. Round robin
test: wear of nine dental restorative materials in six different
wear simulators supplement to the round robin test of 2005.
Dent Mater. 2011;27(2):e1-9. http://dx.doi.org/10.1016/j.
dental.2010.09.003. PMid:20888629.
37. Heintze SD. How to qualify and validate wear simulation devices
and methods. Dent Mater. 2006;22(8):712-34. http://dx.doi.
org/10.1016/j.dental.2006.02.002. PMid:16574212.
38. Heintze SD, Faouzi M, Rousson V, Ozcan M. Correlation
of wear in vivo and six laboratory wear methods. Dent
Mater. 2012;28(9):961-73. http://dx.doi.org/10.1016/j.
dental.2012.04.006. PMid:22698644.
15
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
39. Dal Piva AMO, Tribst JPM, Werner A, Anami LC, Bottino
MA, Kleverlaan CJ. Three-body wear effect on different
CAD/CAM ceramics staining durability. J Mech Behav
Biomed Mater. 2020;103:103579. http://dx.doi.org/10.1016/j.
jmbbm.2019.103579. PMid:32090908.
40. de Gee AJ, Pallav P. Occlusal wear simulation with the ACTA
wear machine. J Dent. 1994;22( Suppl 1):S21-7. http://dx.doi.
org/10.1016/0300-5712(94)90167-8. PMid:8201084.
41. van Duinen RN, Kleverlaan CJ, de Gee AJ, Werner A, Feilzer
AJ. Early and long-term wear of ‘fast-set’ conventional glass-
ionomer cements. Dent Mater. 2005;21(8):716-20. http://dx.doi.
org/10.1016/j.dental.2004.09.007. PMid:16026667.
42. Santing HJ, Kleverlaan CJ, Werner A, Feilzer AJ, Raghoebar
GM, Meijer HJ. Occlusal wear of provisional implant-supported
restorations. Clin Implant Dent Relat Res. 2015;17(1):179-85.
http://dx.doi.org/10.1111/cid.12072. PMid:23594356.
43. Tribst JPM, Dal Piva AMO, Werner A, Anami LC, Bottino
MA, Kleverlaan CJ. Durability of staining and glazing on a
hybrid ceramics after the three-body wear. J Mech Behav
Biomed Mater. 2020;109:103856. http://dx.doi.org/10.1016/j.
jmbbm.2020.103856. PMid:32543416.
44. Lambrechts P, Debels E, Van Landuyt K, Peumans M, Van
Meerbeek B. How to simulate wear? Overview of existing
methods. Dent Mater. 2006;22(8):693-701. http://dx.doi.
org/10.1016/j.dental.2006.02.004. PMid:16712913.
45. International Organization for Standardization. ISO/TS 14569-2
- Dental Materials – Guidance on Testing of Wear – Part 2: Wear
by Two-and/or Three Body Contact. Geneva: ISO; 2001.
46. Osiewicz MA, Werner A, Roeters FJM, Kleverlaan CJ. Wear
of direct resin composites and teeth: considerations for oral
rehabilitation. Eur J Oral Sci. 2019;127(2):156-61. http://dx.doi.
org/10.1111/eos.12600. PMid:30609131.
47. Benetti AR, Larsen L, Dowling AH, Fleming GJ. Assessment
of wear facets produced by the ACTA wear machine. J Dent.
2016;45:19-25. http://dx.doi.org/10.1016/j.jdent.2015.12.003.
PMid:26690332.
48. Ludovichetti FS, Trindade FZ, Werner A, Kleverlaan CJ, Fonseca
RG. Wear resistance and abrasiveness of CAD-CAM monolithic
materials. J Prosthet Dent. 2018;120(2):318.e1-8. http://dx.doi.
org/10.1016/j.prosdent.2018.05.011. PMid:30097264.
49. de Carvalho ABG, Dal Piva AMO, Tribst JPM, Werner A, Saavedra
GSFA, Kleverlaan CJ. Effect of microwave crystallization on the
wear resistance of reinforced glass-ceramics. J Mech Behav
Biomed Mater. 2020;111:104009. http://dx.doi.org/10.1016/j.
jmbbm.2020.104009. PMid:32750672.
50. Alves LMM, Contreras LPC, Bueno MG, Campos TMB, Bresciani E,
Valera MC,etal. The wear performance of glazed and polished
full contour zirconia. Braz Dent J. 2019a;30(5):511-8. http://
dx.doi.org/10.1590/0103-6440201902801. PMid:31596336.
51. Abouelenien DK, Nasr HH, Zaghloul H. Wear behavior of
monolithic zirconia against natural teeth in comparison to two
glass ceramics with two surface finishing protocols: an in-vitro
study. Braz Dent Sci. 2010;23(2):12.
52. Alves LMM, Contreras LPC, Campos TMB, Bottino MA, Valandro
LF, Melo RM. In vitro wear of a zirconium-reinforced lithium
silicate ceramic against different restorative materials. J
Mech Behav Biomed Mater. 2019b;100:103403. http://dx.doi.
org/10.1016/j.jmbbm.2019.103403. PMid:31525551.
53. Tribst JPM, Alves LMM, Piva AMOD, Melo RM, Borges ALS,
Paes-Junior TJA,etal. Reinforced glass-ceramics: parametric
inspection of three-dimensional wear and volumetric loss after
chewing simulation. Braz Dent J. 2019;30(5):505-10. http://
dx.doi.org/10.1590/0103-6440201902699. PMid:31596335.
54. Freddo RA, Kapczinski MP, Kinast EJ, de Souza OB Jr, Rivaldo EG,
da Fontoura Frasca LC. Wear potential of dental ceramics and
its relationship with microhardness and coefficient of friction.
J Prosthodont. 2016;25(7):557-62. http://dx.doi.org/10.1111/
jopr.12330. PMid:26288177.
55. Alves LMM, da Silva Rodrigues C, Ramos GF, Campos TMB,
de Melo RM. Wear behavior of silica-infiltrated monolithic
zirconia: effects on the mechanical properties and surface
characterization. Ceram Int. 2022;48(5):6649-56. http://dx.doi.
org/10.1016/j.ceramint.2021.11.214.
56. ASTM International. G99-17 - Standard Test Method for Wear
Testing with a Pin-on-Disk Apparatus.West Conshohocken:
ASTM International; 2010.
57. Preis V, Behr M, Kolbeck C, Hahnel S, Handel G, Rosentritt M.
Wear performance of substructure ceramics and veneering
porcelains. Dent Mater. 2011;27(8):796-804. http://dx.doi.
org/10.1016/j.dental.2011.04.001. PMid:21524788.
58. Lawson NC, Bansal R, Burgess JO. Wear, strength, modulus
and hardness of CAD/CAM restorative materials. Dent
Mater. 2016;32(11):e275-83. http://dx.doi.org/10.1016/j.
dental.2016.08.222. PMid:27639808.
59. Janyavula S, Lawson N, Cakir D, Beck P, Ramp LC, Burgess JO.
The wear of polished and glazed zirconia against enamel. J
Prosthet Dent. 2013;109(1):22-9. http://dx.doi.org/10.1016/
S0022-3913(13)60005-0. PMid:23328193.
60. Kwon MS, Oh SY, Cho SA. Two-body wear comparison of
zirconia crown, gold crown, and enamel against zirconia. J Mech
Behav Biomed Mater. 2015;47:21-8. http://dx.doi.org/10.1016/j.
jmbbm.2014.11.029. PMid:25837341.
61. Bai Y, Zhao J, Si W, Wang X. Two-body wear performance of
dental colored zirconia after different surface treatments. J
Prosthet Dent. 2016;116(4):584-90. http://dx.doi.org/10.1016/j.
prosdent.2016.02.006. PMid:27157606.
62. Alfrisany NM, Shokati B, Tam LE, De Souza GM. Simulated
occlusal adjustments and their effects on zirconia and
antagonist artificial enamel. J Adv Prosthodont. 2019;11(3):162-8.
http://dx.doi.org/10.4047/jap.2019.11.3.162. PMid:31297175.
63. Po JM, Kieser JA, Gallo LM, Tésenyi AJ, Herbison P, Farella
M. Time frequency analysis of chewing activity in the natural
environment. J Dent Res. 2011;90(10):1206-10. http://dx.doi.
org/10.1177/0022034511416669. PMid:21810620.
64. Ozkir SE, Bicer M, Deste G, Karakus E, Yilmaz B. Wear of monolithic
zirconia against different CAD-CAM and indirect restorative
materials. J Prosthet Dent. 2022:128(3):505-11. http://dx.doi.
org/10.1016/j.prosdent.2021.03.023.
65. Sakaguchi RL, Douglas WH, DeLong R, Pintado MR. The wear of
a posterior composite in an artificial mouth: a clinical correlation.
Dent Mater. 1986;2(6):235-40. http://dx.doi.org/10.1016/S0109-
5641(86)80034-3. PMid:3468027.
66. Chong BJ, Thangavel AK, Rolton SB, Guazzato M, Klineberg
IJ. Clinical and laboratory surface finishing procedures for
zirconia on opposing human enamel wear: a laboratory study.
J Mech Behav Biomed Mater. 2015;50:93-103. http://dx.doi.
org/10.1016/j.jmbbm.2015.06.007. PMid:26116957.
67. Ghazal M, Yang B, Ludwig K, Kern M. Two-body wear of resin
and ceramic denture teeth in comparison to human enamel.
Dent Mater. 2008;24(4):502-7. http://dx.doi.org/10.1016/j.
dental.2007.04.012. PMid:17688934.
68. Heintze SD, Cavalleri A, Forjanic M, Zellweger G, Rousson V.
Wear of ceramic and antagonist--a systematic evaluation of
influencing factors in vitro. Dent Mater. 2008;24(4):433-49.
http://dx.doi.org/10.1016/j.dental.2007.06.016. PMid:17720238.
69. Kim MJ, Oh SH, Kim JH, Ju SW, Seo DG, Jun SH,etal. Wear
evaluation of the human enamel opposing different Y-TZP dental
ceramics and other porcelains. J Dent. 2012;40(11):979-88.
http://dx.doi.org/10.1016/j.jdent.2012.08.004. PMid:22892464.
16
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
70. Heintze SD, Reichl FX, Hickel R. Wear of dental materials: clinical
significance and laboratory wear simulation methods -A review.
Dent Mater J. 2019;38(3):343-53. http://dx.doi.org/10.4012/
dmj.2018-140. PMid:30918233.
71. Al-Wahadni AM, Martin DM. An in vitro investigation into the
wear effects of glazed, unglazed and refinished dental porcelain
on an opposing material. J Oral Rehabil. 1999;26(6):538-
46. http://dx.doi.org/10.1046/j.1365-2842.1999.00394.x.
PMid:10397188.
72. Helal MA, Yang B, Saad E, Abas M, Al-Kholy MR, Imam AY,etal.
Effect of SiO2 and Al2O3 nanoparticles on wear resistance
of PMMA acrylic denture teeth. Braz Dent Sci. 2020;23(3):12.
http://dx.doi.org/10.14295/bds.2020.v23i3.1999.
73. ASTM International. G133-05 (Reapproved) - Standard Test
Method for Linearly Reciprocating Ball-on-Flat Sliding Wear.
West Conshohocken: ASTM International; 2016.
74. Borrero-Lopez O, Guiberteau F, Zhang Y, Lawn BR. Wear of
ceramic-based dental materials. J Mech Behav Biomed Mater.
2019;92:144-51. http://dx.doi.org/10.1016/j.jmbbm.2019.01.009.
PMid:30685728.
75. ŠvančárkoA, Galusková D, Nowicka AE, PálkoH, Galusek
D. Effect of corrosive media on the chemical and mechanical
resistance of ips e.max® cad based Li2Si2O5 glass-ceramics.
Materials (Basel). 2022;15(1):365. http://dx.doi.org/10.3390/
ma15010365. PMid:35009514.
76. Vardhaman S, Borba M, Kaizer MR, Kim D, Zhang Y. Wear
behavior and microstructural characterization of translucent
multilayer zirconia. Dent Mater. 2020;36(11):1407-17. http://
dx.doi.org/10.1016/j.dental.2020.08.015. PMid:32958309.
77. Abhay SS, Ganapathy D, Veeraiyan DN, Ariga P, Heboyan
A, Amornvit P,et al. Wear resistance, color stability and
displacement resistance of milled PEEK crowns compared to
zirconia crowns under stimulated chewing and high-performance
aging. Polymers (Basel). 2021;13(21):3761. http://dx.doi.
org/10.3390/polym13213761. PMid:34771318.
78. Bueno MG, Tribst JPM, Borges ALS. Canine guidance
reconstruction with ceramic or composite resin: A 3D
finite element analysis and in vitro wear study. J Prosthet
Dent. 2022;127(5):765.e1-9. http://dx.doi.org/10.1016/j.
prosdent.2022.01.020. PMid:35287971.
79. Aker DA, Aker JR, Sorensen SE. Toothbrush abrasion of color-
corrective porcelain stains applied to porcelain-fused-to-metal
restorations. J Prosthet Dent. 1980;44(2):161-3. http://dx.doi.
org/10.1016/0022-3913(80)90130-4. PMid:6157021.
80. Bativala F, Weiner S, Berendsen P, Vincent GR, Ianzano J, Harris
WT Jr. The microscopic appearance and effect of toothbrushing
on extrinsically stained metal-ceramic restorations. J Prosthet
Dent. 1987;57(1):47-52. http://dx.doi.org/10.1016/0022-
3913(87)90115-6. PMid:3468249.
81. Garza LA, Thompson G, Cho SH, Berzins DW. Effect of
toothbrushing on shade and surface roughness of extrinsically
stained pressable ceramics. J Prosthet Dent. 2016;115(4):489-
94. http://dx.doi.org/10.1016/j.prosdent.2015.09.013.
PMid:26589442.
82. Yuan JC, Barão VAR, Wee AG, Alfaro MF, Afshari FS, Sukotjo
C. Effect of brushing and thermocycling on the shade and
surface roughness of CAD-CAM ceramic restorations. J
Prosthet Dent. 2018;119(6):1000-6. http://dx.doi.org/10.1016/j.
prosdent.2017.06.001. PMid:28965682.
83. Dal Piva AMO, Bottino MA, Anami LC, Werner A, Kleverlaan CJ,
Lo Giudice R,etal. Toothbrushing wear resistance of stained
cad/cam ceramics. Coatings. 2021;11(2):224. http://dx.doi.
org/10.3390/coatings11020224.
84. Tribst JPM, Maria de Oliveira Dal Piva A, Werner A, Sampaio
Silva LT, Anami LC, Bottino MA,etal. Effect of surface treatment
and glaze application on shade characterized resin-modified
ceramic after toothbrushing. J Prosthet Dent. 2021;125(4):691.
e1-7. http://dx.doi.org/10.1016/j.prosdent.2020.12.040.
PMid:33820617.
85. Magalhaes AC, Wiegand A, Buzalaf MA. Use of dentifrices
to prevent erosive tooth wear: harmful or helpful? Braz Oral
Res. 2014;28(Spec No):1-6. http://dx.doi.org/10.1590/S1806-
83242013005000035. PMid:24554098.
86. Anil N, Bolay S. Effect of toothbrushing on the material loss,
roughness, and color of intrinsically and extrinsically stained
porcelain used in metal-ceramic restorations: an in vitro study.
Int J Prosthodont. 2002;15(5):483-7. PMid:12375465.
87. Labban N, Al Amri MD, Alnafaiy SM, Alhijji SM, Alenizy MA,
Iskandar M,etal. Influence of toothbrush abrasion and surface
treatments on roughness and gloss of polymer-infiltrated
ceramics. Polymers (Basel). 2021;13(21):3694. http://dx.doi.
org/10.3390/polym13213694. PMid:34771250.
88. Labban N, Al Amri M, Alhijji S, Alnafaiy S, Alfouzan A, Iskandar
M,etal. Influence of toothbrush abrasion and surface treatments
on the color and translucency of resin infiltrated hybrid ceramics.
J Adv Prosthodont. 2021;13(1):1-11. http://dx.doi.org/10.4047/
jap.2021.13.1.1. PMid:33747390.
89. Garcia-Godoy F, Garcia-Godoy A, Garcia-Godoy C. Effect
of a desensitizing paste containing 8% arginine and calcium
carbonate on the surface roughness of dental materials and
human dental enamel. Am J Dent. 2009 Mar;22(Spec No
A):21A-4A. PMID: 19472558.
90. Miyano Y, Suzuki M, Shinkai K. Toothbrush abrasion of
restorations fabricated with flowable resin composites with
different viscosities in vitro. Materials (Basel). 2021;14(21):6436.
http://dx.doi.org/10.3390/ma14216436. PMid:34771960.
91. Oliveira GU, Mondelli RF, Charantola Rodrigues M, Franco EB,
Ishikiriama SK, Wang L. Impact of filler size and distribution
on roughness and wear of composite resin after simulated
toothbrushing. J Appl Oral Sci. 2012;20(5):510-6. http://dx.doi.
org/10.1590/S1678-77572012000500003. PMid:23138735.
92. Ganss C, Schlueter N, Preiss S, Klimek J. Tooth brushing habits
in uninstructed adults--frequency, technique, duration and force.
Clin Oral Investig. 2009;13(2):203-8. http://dx.doi.org/10.1007/
s00784-008-0230-8. PMid:18853203.
93. Turssi CP, Kelly AB, Hara AT. Toothbrush bristle configuration
and brushing load: effect on the development of simulated non-
carious cervical lesions. J Dent. 2019;86:75-80. http://dx.doi.
org/10.1016/j.jdent.2019.05.026. PMid:31129277.
94. International Organization for Standardization. ISO/TR 14569
Dental materials – Guidance on testing of Wear – Part 1: Wear
by toothbrushing. (2007)
95. Borges AB, Santos LF, Augusto MG, Bonfiette D, Hara AT,
Torres CR. Toothbrushing abrasion susceptibility of enamel
and dentin bleached with calcium-supplemented hydrogen
peroxide gel. J Dent. 2016;49:54-9. http://dx.doi.org/10.1016/j.
jdent.2016.03.009. PMid:27072568.
96. Sulaiman TA, Camino RN, Cook R, Delgado AJ, Roulet JF, Clark
WA. Time-lasting ceramic stains and glaze: a toothbrush
simulation study. J Esthet Restor Dent. 2020;32(6):581-5. http://
dx.doi.org/10.1111/jerd.12590. PMid:32352643.
97. de Andrade GS, Augusto MG, Simões BV, Pagani C, Saavedra
GSFA, Bresciani E. Impact of simulated toothbrushing on surface
properties of chairside CAD-CAM materials: an in vitro study. J
Prosthet Dent. 2021;125(3):469.e1-6. http://dx.doi.org/10.1016/j.
prosdent.2020.08.028. PMid:33279154.
98. Lopes L, Sampaio-Filho H, Albuquerque E, Tardem C, Miranda
M, Barceleiro M. How do the optical properties of the bulk
fill posterior composites change after 2 years of simulated
17
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
toothbrushing? Braz Dent Sci. 2019;22(3):378-86. http://dx.doi.
org/10.14295/bds.2019.v22i3.1752.
99. Alencar-Silva FJ, Barreto JO, Negreiros WA, Silva PGB, Pinto-
Fiamengui LMS, Regis RR. Effect of beverage solutions and
toothbrushing on the surface roughness, microhardness, and
color stainability of a vitreous CAD-CAM lithium disilicate
ceramic. J Prosthet Dent. 2019;121(4):711.e1-6. http://dx.doi.
org/10.1016/j.prosdent.2019.02.001. PMid:30929660.
100. West NX, Davies M, Amaechi BT. In vitro and in situ erosion
models for evaluating tooth substance loss. Caries Res.
2011;45(Suppl 1):43-52. http://dx.doi.org/10.1159/000325945.
PMid:21625132.
101. Augusto MG, Borges AB, Pucci CR, Mailart MC, Torres CRG. Effect
of whitening toothpastes on wear and roughness of ormocer and
methacrylate-based composites. Am J Dent. 2018;31(6):303-8.
PMid:30658376.
102. Nima G, Lugo-Varillas JG, Soto J, Faraoni JJ, Palma-Dibb RG,
Correa-Medina A,etal. Effect of toothbrushing on the surface of
enamel, direct and indirect CAD/CAM restorative materials. Int
J Prosthodont. 2021;34(4):473-81. http://dx.doi.org/10.11607/
ijp.6594. PMid:33651025.
103. Yin R, Jang YS, Lee MH, Bae TS. Comparative evaluation of
mechanical properties and wear ability of five CAD/CAM
dental blocks. Materials (Basel). 2019;12(14):2252. http://dx.doi.
org/10.3390/ma12142252. PMid:31336968.
104. Jakovac M, Zivko-Babic J, Curkovic L, Aurer A. Chemical durability
of dental ceramic material in acid medium. Acta Stomatol Croat.
2006;40(1):65-71.
105. International Organization for Standardization. ISO 6872 -
International Standards for Dental Ceramics. Geneva: ISO; 1995.
106. Mohammed MS, Mohsen CA. Effect of corrosion on some
properties of dental ceramics. Sys Rev Pharm. 2021;12(6):2136-40.
107. Kukiattrakoon B, Hengtrakool C, Kedjarune-Leggat U.
The effect of acidic agents on surface ion leaching and
surface characteristics of dental porcelains. J Prosthet
Dent. 2010;103(3):148-62. http://dx.doi.org/10.1016/S0022-
3913(10)60021-2. PMid:20188237.
108. Kukiattrakoon B, Hengtrakool C, Kedjarune-Leggat U. Effect of
acidic agents on surface roughness of dental ceramics. Dent
Res J (Isfahan). 2011;8(1):6-15. PMid:22132009.
109. Mohsen C. Corrosion effect on the flexural strength & micro-
hardness of ips e-max ceramics. Open J Stomatol. 2011;1(02):29-
35. http://dx.doi.org/10.4236/ojst.2011.12006.
110. Elraggal A, Afifi R, Abdelraheem I. Effect of erosive media on
microhardness and fracture toughness of CAD-CAM dental
materials. BMC Oral Health. 2022;22(1):191. http://dx.doi.
org/10.1186/s12903-022-02230-1. PMid:35590294.
111. Anusavice KJ. Degradability of dental ceramics. Adv Dent Res.
1992;6(1):82-9. http://dx.doi.org/10.1177/0895937492006001
2201. PMid:1292468.
112. Yu P, Xu Z, Arola DD, Min J, Zhao P, Gao S. Effect of acidic agents
on the wear behavior of a polymer infiltrated ceramic network
(PICN) material. J Mech Behav Biomed Mater. 2017;74:154-63.
http://dx.doi.org/10.1016/j.jmbbm.2017.06.001. PMid:28599155.
113. Kermanshah H, Ahmadi E, Rafeie N, Rafizadeh S, Ranjbar
Omrani L. Vickers micro-hardness study of the effect of fluoride
mouthwash on two types of CAD/CAM ceramic materials
erosion. BMC Oral Health. 2022;22(1):101. http://dx.doi.
org/10.1186/s12903-022-02135-z. PMid:35354455.
114. Kruzic JJ, Arsecularatne JA, Tanaka CB, Hoffman MJ, Cesar
PF. Recent advances in understanding the fatigue and wear
behavior of dental composites and ceramics. J Mech Behav
Biomed Mater. 2018;88:504-33. http://dx.doi.org/10.1016/j.
jmbbm.2018.08.008. PMid:30223214.
115. Ramos GF, Ramos NC, Alves LMM, Kaizer MR, Borges
ALS, Campos TMB, et al. Failure probability and stress
distribution of milled porcelain-zirconia crowns with bioinspired/
traditional design and graded interface. J Mech Behav
Biomed Mater. 2021;119:104438. http://dx.doi.org/10.1016/j.
jmbbm.2021.104438. PMid:33798936.
116. Maier E, Grottschreiber C, Knepper I, Opdam N, Petschelt
A, Loomans B,et al. Evaluation of wear behavior of
dental restorative materials against zirconia in vitro. Dent
Mater. 2022;38(5):778-88. http://dx.doi.org/10.1016/j.
dental.2022.04.016. PMid:35459553.
117. Faria ACL, de Oliveira AA, Alves Gomes É, Silveira Rodrigues
RC, Faria Ribeiro R. Wear resistance of a pressable low-fusing
ceramic opposed by dental alloys. J Mech Behav Biomed Mater.
2014;32:46-51. http://dx.doi.org/10.1016/j.jmbbm.2013.12.018.
PMid:24412716.
118. Rossi NR, Sato TP, Marinho CC, Macedo VC, Paes TJA Jr, Kimpara
ET. Influence of surface treatments in flexural strength and
superficial topography of a lithium disilicate ceramic. Braz Dent
Sci. 2019;22(3):305-12. http://dx.doi.org/10.14295/bds.2019.
v22i3.1678.
119. Ozdogan A, Yesil Duymus Z, Ozbayram O, Bilgic R. Effect of
different bleaching agents on the surface roughness and color
stability of feldspathic porcelain. Braz Dent Sci. 2019;22(2):213-9.
http://dx.doi.org/10.14295/bds.2019.v22i2.1695.
120. Contreras L, Dal Piva A, Ribeiro FC, Anami LC, Camargo S, Jorge
A,etal. Effects of manufacturing and finishing techniques of
feldspathic ceramics on surface topography, biofilm formation,
and cell viability for human gingival fibroblasts. Oper Dent.
2018;43(6):593-601. http://dx.doi.org/10.2341/17-126-L.
PMid:29856699.
121. Dal Piva AMO, Tribst JPM, Venturini AB, Anami LC, Bonfante
EA, Bottino MA,etal. Survival probability of zirconia-reinforced
lithium silicate ceramic: effect of surface condition and fatigue
test load profile. Dent Mater. 2020;36(6):808-15. http://dx.doi.
org/10.1016/j.dental.2020.03.029. PMid:32360042.
122. Dal Piva A, Contreras L, Ribeiro FC, Anami LC, Camargo S, Jorge
A,etal. Monolithic ceramics: effect of finishing techniques on
surface properties, bacterial adhesion and cell viability. Oper
Dent. 2018;43(3):315-25. http://dx.doi.org/10.2341/17-011-L.
PMid:29533718.
123. Makkeyah F, Morsi T, Wahsh M, El-Etreby A. An in vitro
evaluation of surface roughness, color stability and bacterial
accumulation of lithium disilicate ceramic after prophylactic
periodontal treatment. Braz Dent Sci. 2021;24(3). http://dx.doi.
org/10.14295/bds.2021.v24i3.2572.
124. Reyes-Sevilla M, Kuijs RH, Werner A, Kleverlaan CJ, Lobbezoo F.
Comparison of wear between occlusal splint materials and resin
composite materials. J Oral Rehabil. 2018;45(7):539-44. http://
dx.doi.org/10.1111/joor.12636. PMid:29663496.
125. Kaizer MR, Gierthmuehlen PC, Dos Santos MB, Cava SS, Zhang Y.
Speed sintering translucent zirconia for chairside one-visit dental
restorations: Optical, mechanical, and wear characteristics.
Ceram Int. 2017;43(14):10999-1005. http://dx.doi.org/10.1016/j.
ceramint.2017.05.141. PMid:29097830.
126. Witecy C, Ganss C, Wöstmann B, Schlenz MB, Schlenz MA.
Monitoring of erosive tooth wear with intraoral scanners
in vitro. Caries Res. 2021;55(3):215-24. http://dx.doi.
org/10.1159/000514666. PMid:33752205.
127. Schlenz MA, Schlenz MB, Wöstmann B, Jungert A, Ganss C.
Intraoral scanner-based monitoring of tooth wear in young adults:
12-month results. Clin Oral Investig. 2022;26(2):1869-78. http://
dx.doi.org/10.1007/s00784-021-04162-6. PMid:34498100.
128. García VD, Freire Y, Fernández SD, Murillo BT, Sánchez MG.
Application of the intraoral scanner in the diagnosis of dental
wear: an in vivo study of tooth wear analysis. Int J Environ Res
18
Braz Dent Sci 2023 Jan/Mar;26 (1): e3638
Ramos NC et al.
Wear of dental ceramics
Ramos NC et al.
Wear of dental ceramics
Date submitted: 2022 Sep 20
Accepted submission: 2022 Oct 10
Amanda Maria de Oliveira Dal Piva
(Corresponding address)
Universiteit van Amsterdam en Vrije Universiteit, Afdeling Materiaalwetenschappen,
Academisch Centrum Tandheelkunde Amsterdam, Amsterdam, Nederlands.
Email: a.m.de.oliveira.dal.piva@acta.nl
Public Health. 2022;19(8):4481. http://dx.doi.org/10.3390/
ijerph19084481. PMid:35457351.
129. Viegas DC, Mourão JT, Roque JC, Riquieri H, Fernandes J, Arrobas
FV,etal. Evaluation of the influence of the impression technique,
scanning direction and type of scanner on the accuracy of
the final model. Braz Dent Sci. 2021;24(1):1-13. http://dx.doi.
org/10.14295/bds.2021.v24i1.2179.
130. Kumar S, Keeling A, Osnes C, Bartlett D, O’Toole S. The
sensitivity of digital intraoral scanners at measuring early
erosive wear. J Dent. 2019;81:39-42. http://dx.doi.org/10.1016/j.
jdent.2018.12.005. PMid:30578831.
131. Ibrahim NM, El-basty R, Katamish H. Clinical evaluation of wear
behavior of human enamel and chipping of veneered zirconia
against monolithic zirconia (randomized controlled clinical trial).
Braz Dent Sci. 2020;23(4):1-11. http://dx.doi.org/10.14295/
bds.2020.v23i4.2081.
132. Aladağ A, Oğuz D, Çömlekoğlu ME, Akan E. In vivo wear
determination of novel CAD/CAM ceramic crowns by using 3D
alignment. J Adv Prosthodont. 2019;11(2):120-7. http://dx.doi.
org/10.4047/jap.2019.11.2.120. PMid:31080573.
133. Marchand L, Sailer I, Lee H, Mojon P, Pitta J. Digital wear
analysis of different CAD/CAM fabricated monolithic ceramic
implant-supported single crowns using two optical scanners. Int
J Prosthodont. 2022;35(3):357-64. http://dx.doi.org/10.11607/
ijp.7430. PMid:33751002.
134. Michou S, Vannahme C, Ekstrand KR, Benetti AR. Detecting
early erosive tooth wear using an intraoral scanner system.
J Dent. 2020;100:103445. http://dx.doi.org/10.1016/j.
jdent.2020.103445. PMid:32750388.
135. Grangeiro MTV, Rodrigues CDS, Rossi NR, da Silva JMD, Ramos
NDC, Tribst JPM, etal. Effect of surface-etching treatment,
glaze, and the antagonist on roughness of a hybrid ceramic
after two-body wear. Materials (Basel). 2022;15(19):6870. http://
dx.doi.org/10.3390/ma15196870. PMid:36234211.