You are on page 1of 66

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/356552813

Dental resin composites: A review on materials to product realizations

Article in Composites Part B Engineering · November 2021


DOI: 10.1016/j.compositesb.2021.109495

CITATIONS READS

60 5,671

5 authors, including:

Kiho Cho Ginu Rajan


The University of Hong Kong University of Wollongong
28 PUBLICATIONS 233 CITATIONS 161 PUBLICATIONS 2,575 CITATIONS

SEE PROFILE SEE PROFILE

Paul Farrar Leon Prentice


SDI SDI Ltd
21 PUBLICATIONS 248 CITATIONS 57 PUBLICATIONS 484 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Kiho Cho on 25 January 2022.

The user has requested enhancement of the downloaded file.


Dental Resin Composites: A Review on Materials to Product
Realizations
Kiho Cho a,b,*, Ginu Rajan c,d, Paul Farrar e, Leon Prentice e and B. Gangadhara Prusty b,c,**
a
Dental Materials Science, Division of Applied Oral Sciences and Community Dental Care,
Faculty of Dentistry, The University of Hong Kong, Hong Kong SAR, China
b
School of Mechanical and Manufacturing Engineering, University of New South Wales,
Sydney, NSW 2052, Australia
c
ARC Training Centre for Automated Manufacture of Advanced Composites, University of
New South Wales, Sydney, NSW 2052, Australia
d
School of Technologies, Cardiff Metropolitan University, Cardiff, CF52YB, The United
Kingdom
e
SDI Ltd, Bayswater, VIC 3153, Australia

* Corresponding author - K. Cho (dkcho@hku.hk)


** Corresponding author – B. G. Prusty (g.prusty@unsw.edu.au)

Abstract: Dental resin composites have revolutionized dental care and enabled minimally

invasive dentistry to preserve healthy tooth structure and provide natural-appearing esthetic

results; these materials are now good alternatives to metal and amalgam restorations.

Nevertheless, dental composites are being further developed to enhance their long-term clinical

performance and longevity. With a complete understanding of the various material

characteristics and design strategies of dental composite systems, frontiers of dental restoration

research can aim towards creating novel materials that can produce very similar properties,

functionalities, and internal structures as hard dental tissues. In this review article, the authors

present an overview of the synthesis of the advanced dental composite systems with the recent

research and development fields over the last 5 years. This review also explores how to control

and optimize the required properties of the composites, ideally to increase the

longevity/durability of restorations preventing recurrent caries. Research studies and

commercial products are introduced to forecast the demands and trends of resin-based dental

composites, in order to assist clinicians and researchers in optimal selection of materials to

fulfill mechanical, physical, biological, and functional requirements.

Keywords: Dental resin composites, Dental restorations, Polymer, Characterization, Caries.

1
1. Introduction

A dental restoration is a clinical treatment achieved by excavating the damaged or decayed

teeth and replacing natural structures with a durable material such as metal, ceramic, cement,

and composite. Of these, the research and development of dental resin composites (DRCs) have

been gradually extended for direct restoration on the anterior and posterior tooth because of

their good mechanical and physical characteristics along with significant potential to generate

additional functionality such as antibacterial and therapeutic activities [1-3]. Dental composite

research is currently undergoing a reformation phase due to the increased demands for low-

cost, non-toxic, higher-longevity, biological functionality, and superior aesthetics on the DRCs

fulfilling minimally invasive (MI) restorative treatments. Also, they have been designed to

replace amalgam restorations which are still prevalent in most low-income countries due to the

low-cost and simplified clinical treatment [4, 5]. Except developing countries, direct

restorations using DRCs for posterior teeth have exceeded the use of amalgam since 1998. It

is particularly of note that amalgam restorations in children are far less used than in adults [6].

The replacement of amalgam fillings, which may be associated with public health concern due

to the mercury component, may reduce dental health complications such as oral lesions or

allergic reactions [7].

DRCs are typically a mixture of dental resins and diverse inorganic fillers. Resins are composed

of a mixture of two or more monomers to achieve balanced functionalities in workable

rheology and the desired mechanical properties before and after curing [8]. Accordingly, an

understanding of the chemical characteristics and their polymerization kinetics is necessary to

design the optimized dental material systems by evaluating the effects of the trade-offs between

resin viscosity, degree of conversion, and mechanical strength. A shorter photo-curing time

and an increased depth of cure and degree of conversion by the visible light source (blue light,

2
λ ≈ 450 ~ 495 nm), which is much less harmful to the human body, particularly to skin and

eyes, than ultraviolet (UV, λ < 400 nm) irradiation, have attributed to the fast-growing

popularity of DRCs in dentistry. Also, the rapid advances of solid-state light-emitting diode

technology has provided high-quality performance and ease of implementation of DRC in

dental treatments [9, 10].

With the continuous advent of micro-/nano-science and technology, the addition of various

filler contents, such as silica glass, quartz, ceramic, metal, pre-polymerized particles, and

natural minerals in diverse shapes and sizes, can enhance the mechanical properties [11-14] of

the dental composites while obtaining the practical functionalities such as low shrinkage

volume and stress [15, 16], desired flowability (or viscosity) [17, 18] and shade [19, 20], and

good biocompatibility [21-23] for various clinical applications. Several factors, including

weight fraction, shape and size, orientation, and dispersion of the fillers in the resin matrix are

critical in controlling the characteristics of the designed composite systems. To maximize the

physical, mechanical, and biological performance of the resultant composites, strong interfacial

reactions between fillers and resin matrix are essential. The silane monomers, which contain

organic-inorganic functional groups and can chemically bridge the inorganic fillers to organic

resins to enhance the interfacial bonding, are most widely used in polymer composite industries

as well as dentistry to modify the surface of the filler materials [24-27].

The mechanical, physical, and biological characteristics of the interaction between fillers and

resin matrix is primarily affected by the physical and chemical characteristics of the filler

surface, so it is necessary to accurately investigate the effect of surface modification on the

interfacial properties. For instance, in fiber reinforced dental composite systems [28, 29], they

can be examined through the experimental studies and measurements such as fiber pull-out

[30] and microdroplet pull-out tests [31]. However, experimental tests in micro-/nano-scale

3
require highly sensitive techniques to obtain reliable data, and are often correspondingly

expensive. Computer simulations, especially molecular mechanics and dynamic simulations,

can offer a new method for predicting mechanical properties of the composite systems and

dynamic interfacial behaviors around the filler materials [32, 33]. Providing a deeper

understanding of interfacial behaviors at the molecular level is important in the determination

of the expected mechanical properties of the newly designed composite systems in macroscale

[13, 34].

Despite efforts to improve the mechanical performance of the DRCs, the combined mechanical

and functional properties of these materials are less than ideal as compared to healthy human

teeth, particularly for restorations in posterior teeth. In addition, due to the inherent nature of

the resins used in the composite materials, polymerization shrinkage increases the residual

stress and can lead to interfacial debonding between the restored composite and tooth wall. The

fractured surface and micro-cracks could supply a favorable environment to oral bacteria and

promote the accumulation of biofilm, which is associated with recurrent decay and

postoperative hypersensitivity [35, 36]. Therefore, the incorporation of antibacterial and

remineralization functionalities in dental composites is beneficial for the prevention of

recurring caries and microbial destruction and it would increase the longevity of the restorative

composites.

To improve the performance of DRCs or to reduce their production costs, a large variety of

new research on dental resins, fillers, and their interfacial interactions have been conducted

over the last decade. The complexities of regulation, validation, and product assurance in

biomedical materials raises the barrier to entry for new technologies, and this new research

may take time to reach production. Furthermore, the nature of biological interfaces, patient-

specific considerations, and long-term influences on restoration longevity mean that

4
demonstration of efficacy may also take time. Biomedical materials manufacturers must have

robust methods to demonstrate the safety, efficacy, and applicability of the DRCs they develop,

supported by fundamental understanding of the chemistries and structures that underpin the

performance of the materials. An examination of the general trends of recent research and

commercial DRC products would help researchers and practitioners to elucidate future research

directions and assist in optimal selection of dental materials.

2. Dental Resin Composites

2.1 Dental composite system

The term “composite” or “composite material” refers to a structural material made from two or

more constituent materials normally to provide better characteristics or to add further

functionalities to the original components. Dental restoration materials have also been

developed in various types of resin-based composite system, which mainly consists of three

components, an organic resin matrix, inorganic fillers, and a coupling agent to chemically

anchor the resin to the filler surface [37]. They contain other important ingredients, such as

initiators for triggering polymerization, inhibitors for preventing side reactions of unwanted

polymerization or formation of by‐products, and pigments (or dyes) for matching the

appropriate shade with the natural tooth. Each is critical for achieving the successful composite

restoration that will benefit patients, provide an alternative to amalgam restorations, exhibit

easy clinical handling for filling and repair, allow less tooth removal, show a higher fracture

toughness compared to the dental ceramics, and be a good color match to the tooth. With these

properties, DRCs are now extensively used in restorative dentistry.

The early dental composites were introduced in the middle of 1950s as a mixture of two

monomers and macro-sized silicate particles with a focus on increasing mechanical properties

[38]. The emergence of micro-/nano-manufacturing technologies enabled the synthesis of

5
Figure 1. Evolution of classification of dental resin composites based on filler materials.

smaller and better-defined fillers with various structures such as particles, tubes, rods, hollow

spheres, and fibers; multi-functionalized dental composites could be developed by integrating

a variety of fillers (Figure 1). Surface modification on the filler surface using physical and

chemical processes enhances the interfacial bond and dispersion of fillers in the resin matrix

simultaneously improving the targeted mechanical characteristics [39, 40]. Over the last decade,

there have been increasing interests in bioactive DRCs capable of performing antimicrobial

activities [41-46] and remineralization of initial carious lesions [42,47]. Thus, basic studies on

the characteristics of each component are necessary to determine the design and process

strategy that can result in enhanced properties of the resulting dental composite materials.

2.2 Dental resin

6
The dental resin is an active organic monomer system in which the base monomer is mixed

with the diluent monomer to attain an optimized viscosity offering acceptable clinical handling

properties. Also, photo-initiators as the trigger of the free radical polymerization reaction, and

inhibitors (or stabilizers) for preventing oxidation of monomers under aerobic conditions are

contained in the dental resin systems. Mechanical and physical properties of the composites

before and after polymerization are directly dependent on the type of monomers and a ratio of

the monomer fraction. Among many types of dental monomers, the base monomers, bisphenol

A glycidyl methacrylate (Bis-GMA), ethoxylated bisphenol A dimethacrylate (Bis-EMA), and

urethane dimethacrylate (UDMA), and as cross-linking diluents to adjust the viscosity of

mixtures, triethylene glycol dimethacrylate (TEGDMA), decanediol dimethacrylate (D3MA),

and 2-hydroxyethyl methacrylate (HEMA) are most commonly used to compose DRCs (Figure

2) [8, 48-50]. These methacrylate monomers are characterized by alkene groups that contain

carbon-carbon double bond (C=C) functional groups, which allows facilitating a fast chain

polymerization and cross-linking of polymers. The reactive methacrylate groups in monomers

easily polymerize through light-initiated and/or chemically initiated curing, inducing the first

step in the chain-reaction polymerization process and the second step of cross-linking reaction,

which is highly associated with mechanical and physical properties such as fracture strength,

modulus, shrinkage, and dimensional stability of the composites.

Bis-GMA was introduced in 1962 by Bowen for dental composites with a mixture of silica

particles [51] and has been widely used in the dental resin matrix due to its high molecular

weight (MW = 512.6 g/mol) and high viscosity (η = 1369 Pa·s) of the Bis-GMA molecule

which typically results in relatively higher mechanical performance [52]. Its high viscosity and

low flexibility are conferred by the aromatic rings in the backbone with hydroxyl groups that

form strong hydrogen bonds interrupting the rotational motion of the molecule. It is usually

blended with a low viscosity monomer such as TEGDMA (MW = 286.3 g/mol, η = 0.05 Pa·s)

7
Figure 2. Molecular structures of the dental monomers and photopolymerization kinetics.

not only to increase the clinical handling properties but also to improve the degree of

conversion (DC) and cross-link density by increasing the mobility of monomers and radicals

in the matrix during the photopolymerization process [52]. New monomers such as glycerol

8
dimethacrylate (GDMA), acetylated glycerol dimethacrylate (AGDMA), and glycerol

trimethacrylate (GTMA), which can substitute TEGDMA, were synthesized showing

acceptable mechanical and cytotoxicity properties [53]. Later, a combination with Bis-EMA

(MW = 540.0 g/mol, η = 0.9 Pa·s) has come to attention due to its low viscosity, high DC, and

low polymerization shrinkage stress compared to Bis-GMA. However, a release of bisphenol

A (BPA) from Bis-GMA and Bis-EMA has been reported, and it is often considered as a

harmful material in the human body at high concentration. Thus, UDMA (MW = 470.0 g/mol,

η = 05~10 Pa·s) and BPA-free monomers have been introduced such as 9,9-Bis[4-((2-(2-

methacryloyloxy)ethyl-carbamate)ethoxy)phenyl]fluorene (Bis-EFMA), trimethacrylate

tris(4-hydroxyphenyl)methane triglycidyl methacrylate (TTM), 2-hydroxy-1-ethyl

methacrylate (2EMATE-BDI), and ORMOCER-based resin [54-57]. Although UDMA-based

composites exhibit higher polymerization shrinkage and more brittle mechanical

characteristics than Bis-GMA due to the molecular characteristics of the shorter chain-length,

UDMA is more flexible (less viscous) and allows adding more inorganic fillers without mixing

additional diluent monomers. It can also be modified by integrating a hydrophobic substituent

within the monomer backbone, which results in a reduction of water sorption and solubility of

UDMA composite systems [58]. In a recent study, urethane acrylic monomer-based resins such

as TMXDI-HEA, TMXDI-HPA, and NBDI-HEA, exhibited higher mechanical properties (up

to 40% in elastic modulus and up to 21% in flexural strength) compared to the cured UDMA-

based resin [59]. They can be good candidates to substitute Bis-GMA and UDMA as base

monomers in dental restoration composites, however acrylate monomers may exhibit a higher

level of toxicity than methacrylate monomers and should be carefully considered for dental

materials.

Even though it is necessary to incorporate diluent monomers into DRCs, most often TEGDMA

to achieve the optimal viscosity and handling properties, their high polymerization shrinkage

9
and stress lead to marginal leakage and recurrent caries. Several research studies introduced in

recent studies new diluents with low viscosity as a substitute for TEGDMA monomer in dental

resin systems. The developed monomers showed many interesting mechanical and physical

characteristics as summarized in Table 1 and the related chemical structure of the emerging

diluent monomers are presented in Figure 2. These studies showed the enhanced properties of

low volume shrinkage and shrinkage stress without compromising mechanical and biological

properties to act as a viable candidate to replace TEGDMA, including aromatic methacrylates

[60], triethylene glycol divinylbenzyl ether (TEG-DVBE) [61], diallyl(5-(hydroxymethyl)-1,3-

phenylene) decarbonate (HMFBA) and 5-(hydroxymethyl)-1,3-phenylene bis(2-

methylacrylate (HMFBM) [62], diallyl (propane-2,2-diylbis (1,4-phenylene)) biscarbonate

(BPhADAC) [63], allyl(2-(2-(((allyloxy)carbonyl)oxy)benzoyl)-5-methoxyphenyl) carbonate

(BZ-AL) [64], trifunctional monomers (5A1603DA, 5A13DMA) [65], and n-methyl-bis(ethyl-

carbamate-isoproply-α-methylstyryl)amine (Phene) [66]. Especially, a new hydrolytically

stable monomer, TEG-DVBE, was suggested as a good alternative to design a longer-lasting

dental resin composite in water-rich oral environments. A combination with UDMA showed

superior water solubility and modulus recovery upon water sorption compared to Bis-

GMA/TEGDMA system [61].

Table 1. Emerging dental monomers and their characteristics.

Functional
Type Monomer Measured characteristics Ref.
group
Bis-EFMA Methacrylate Compared to Bis-GMA based resin,
• BPA-free
• Increased degree of conversion (DC)
• Reduced water sorption (WSP) and
Base solubility (WSB) [54]
• Higher flexural strength (FS) and modulus
(FM)
• Similar biocompatibility (no cytotoxicity)

10
TTM Methacrylate Compared to Bis-GMA based resin,
• BPA-free
• Comparable FS, FM, DC, and volumetric
shrinkage (VS) [55]
• Similar biocompatibility (no cytotoxicity)
• Increased WSP and WSB

2EMATE- Methacrylate Compared to Bis-GMA based resin,


BDI • BPA-free
• Reduced WSP and WSB [56]
• Lower DC

TMXDI- Acryl Compared to Bis-GMA and UDMA based


HEA, resin,
TMXDI- • Higher FS and FM
HPA, • Similar viscosity and refractive index (RI) [59]
XDI-HPA,
NBDI-HEA

Aromatic Methacrylate Compared to TEGDMA,


methacrylates • Low viscosity and a high RI (1.553–1.573)
• Higher DC
• High depth of cure and excellent aesthetic
[60]
properties (low transparency)
• Higher FM
• Lower shrinkage stress (SS) and WSP

TEG-DVBE Vinyl Compared to Bis-GMA/TEGDMA,


• Superior in FS, fracture toughness, and FM
recovery upon water sorption
[61]
• Lower WSB
• Higher DC

Diluent / HMFBA Allyl Compared to TEGDMA,


Comonomer HMFBM • Similar FS
• Lower VS
[62]
• Higher DC
• Superior cytotoxicity

BPhADAC Allyl Compared to TEGDMA,


• Higher DC
• Lower VS and WSP [63]
• Comparable FS and cytotoxicity

BZ-AL Allyl Compared to TEGDMA,


• Higher FS, but lower FM
• Higher DC
[64]
• Lower SS, but no difference in the VS
• Similar biocompatibility (no cytotoxicity)

11
5A13DA Methacrylate Compared to TEGDMA,
5A13DMA + Allyl • Higher DC
• Lower WSB [65]
• Lower VS
• Low cytotoxicity

Phene Isopropenyl Compared to Bis-GMA/TEGDMA,


• Lower VS and SS
• Comparable FS and FM up to 20 wt% use [66]
of Phene
Bis-EFMA: 9,9-Bis[4-((2-(2-methacryloyloxy)ethyl-carbamate)ethoxy)phenyl]fluorene, TTM:
trimethacrylate tris(4-hydroxyphenyl)methane triglycidyl methacrylate, 2EMATE-BDI: 2-hydroxy-1-ethyl
methacrylate, TMXDI: 1,3-bis(1-isocyanato-1-methylethyl)benzene, HEA: 2-hydroxyethyl acrylate, HPA:
2-hydroxy propyl acrylate, XDI: 1,3-bis(isocyanatomethyl)benzene, NBDI: norbornane diisocyanate,
HMFBA: diallyl(5-(hydroxymethyl)-1,3-phenylene) dicarbonate, HMFBM: 5-(hydroxymethyl)-1,3-
phenylene bis(2-methylacrylate), TEG-DVBE: triethylene glycol divinylbenzyl ether, Phene: N-methyl-
bis(ethyl-carbamate-isoproply-α-methylstyryl)amine, BPhADAC: diallyl (propane-2,2-diylbis (1,4-
phenylene)) biscarbonate, BZ-AL: allyl(2-(2-(((allyloxy)carbonyl)oxy)benzoyl)-5-methoxyphenyl)
carbonate.

The combination of these base and diluent monomers has been extensively investigated to

improve the mechanical and physical properties of the composites and contribute to extending

the dental material industry. A systematic investigation of material property prediction models

has been developed to optimize input parameters such as weight fraction and concentration of

monomers [67-69] and fillers [70]. To date, there are two major types of DRCs classified

according to handling characteristics: flowable and packable materials. The handling can be

controlled by adjusting intrinsic factors such as type and fraction of monomers and fillers. The

low viscosity of flowable composites allows them to be easily injected into the prepared cavity,

ensuring improved marginal adaptation compared to packable composites. They can also be

used in bulk-filling posterior restoration [71, 72] as a single increment up to 4- or 5-mm depth

as well as a base layer under the packable composites. Packable composites can be synthesized

by loading high viscosity resins or increasing filler contents (up to 60~80% by weight). They

will present better stability in shaping tooth contours and an enhanced quality of surface

hardness after curing, so they are commonly applied for constructing onlay and overlay

restorations. Unlike flowable composite, the high viscosity of the packable composites may

risk generating internal voids along the cavity walls or sharp corners, which lessen the

12
longevity of the restorations [73]. A low level of DC and depth of cure, which is mainly caused

by the reduced mobility of monomer radicals and light penetration, limits their usage to thin

layer structures (normally less than 2 mm in thickness).

2.3 Filler materials

A composite material is reinforced with filler materials that have properties different from the

resin matrix. The main purpose of blending fillers is to provide the enhanced mechanical and

physical properties to the resultant dental composites. The resin matrix contributes to the

success of direct restoration in a prepared tooth cavity, and filler materials add additional

functionalities such as strength and modulus [74, 75], surface hardness [76, 77], toughness

[78], aesthetics [79], polymerization shrinkage and stress [80, 81], hydrothermal aging

resistance [82], and antibacterial activities [83-86]. The literature studies on resin based dental

composites reinforced with various types of micro-/nano-filler materials show a gradual

increase every year (Figure 3, Table 2). The effect of filler on these properties depends on the

type, size, shape, fraction, dispersion and orientation, and surface modification of the

embedded fillers [87-89]. Silica-based glasses [27, 90-95], ceramic [25, 79, 96, 97], metals [98-

101], natural minerals [102-108], pre-polymerized particles [109, 110], and cellulose crystal

particles [111, 112] have been incorporated into dental resins with various shapes such as

irregular particle, sphere, plate, fiber, and tube.

Silica-based glass particles are the most popular fillers for dental composites because they are

strong, hard, and highly chemical durable. The formulated glasses can be manufactured with

diverse properties even using low-cost techniques. Their good optical characteristics assist

13
Figure 3. Trend of publications over research on dental resin composites (a) and percentage
breakdown of publications by filler types (b). For publication tracking, keywords "resin (or
polymer) dental restoration composite(s)" and the name of each filler material were matched
only in the ‘Title’ of the article to find the influential articles that closely associate with the
research topic. Google Scholar was used as a tracking tool.

photocuring kinetics, allowing deeper penetration of light into the resin matrix and offer a wide

range of flexibility to adjust the color of the composites through the addition of pigments. Due

to this reason, carbon fibers and carbon nanotubes are not suitable for dental composite

applications even though they provide excellent mechanical strength. Zirconia- and alumina-

based ceramic fillers [113] are used to increase the elastic modulus, hardness, and wear

resistance. However, a high fraction of ceramic particles can lead to excessive wear of the

opposing natural teeth by abrasion [114]. In recent years, a metallic powder such as Ag, Au,

TiO2 (metal oxide) has been studied to augment the antibacterial properties, which would help

impede the development of recurrent lesions near the interface between composite and tooth

wall [115-117]. Metal ions released from the surface would inhibit biofilm formation by killing

or damaging the bacteria [45]. Their antibacterial capabilities have been verified in many in-

vitro experimental studies, but clinical studies are required to support their effectiveness in

complex oral environments.

14
Table 2. Various types of filler materials used in dental resin composites.

Filler
Filler materials Shape Size Functions Ref.
categories
Mechanical properties (compressive / [82, 84,
Silica nanoparticle Spherical nanoparticle 195-920 nm (D)
flexural strength, modulus) 85]
Calcinated Spherical nanoparticle Mechanical properties (flexural/compressive
Silica 1-3 µm (D) [12]
colloidal silica cluster strength, F. modulus, hardness)
based
Silicate bioactive
particles Micro/Nanoparticle 5.6 µm / 70 nm (D) Antimicrobial, remineralizing effect [87]
glass
250-500 µm (L), 5 µm (D) Mechanical properties (toughness, flexural [68,
Glass fiber Micro-sized short fiber
200-400 µm (L), 17 µm (D) strength, modulus) 86]
Glass- Al2O3-0.5B2O3-
8-16.9 µm (D) [89,
ceramic / SiO2, Micro/Nanoparticle Mechanical properties, antimicrobial effect
50 nm (D) 90]
Ceramic ZrO2
Ag: 10-15 nm (D) [91,
Metals Ag, TiO2, Au Nanoparticle/Nanotube TiO2: 10 nm (D), 200 nm (L) Antimicrobial 92,
Au: 5 nm (D) 94]
Halloysite
Nanotube 30-70 nm (D), 1-3 μm (L) Mechanical properties, antimicrobial effect [95]
nanotube
Nanoparticle [71,
Nanocrystal, Mechanical properties, antimicrobial and
Hydroxyapatite 5-6 μm (D) 24,
Micro-sized cluster remineralizing effect
Mineral 1-4 μm (D) 96]
particles 116 nm (D) [97, 98
Calcium phosphate Micro/Nanoparticle Remineralizing effect
15-23 μm (D) 99]
Zeolite
(Calcium-rich Nanoparticle < 4 μm (D) Remineralizing effect [100, 101]
zeolite)
Polymer
Pre-polymerized
based Micro/Nanoparticle - Mechanical properties, antimicrobial effect [102]
particles
particles

15
Micro/nanofabrication technologies for small-sized structures such as particle, fiber, wire, and

tube are rapidly growing [47, 118]. Such advances as shown in Figure 4 have eventually led to

the successful development of novel dental composites. The early stages of dental composites

can be classified by macro-sized glass fillers that were ground into particles of various sizes

(10 ~ 100 µm in diameter) in a ball mill. These large particles with non-uniform sizes would

be easily exposed to the outer surface of the composite as the resin matrix is scraped off over

time. Its irregular shape with sharp edges protruding from the composite surface may cause

wear of the opposing tooth. Irregular surfaces may irritate the inner cheek or tongue. Also, the

roughened surface above the threshold of 0.2 μm (Ra) increases bacterial attachment and

biofilm formation [119]. To overcome those disadvantages, nano-sized silica fillers (10 ~ 100

nm) have been employed. These smaller particles have a significantly higher specific surface

area (m2·g-1) compared with conventional macro-fillers. The large surface area requires more

resin to cover the whole area of the small particles compared to the large ones. For a given

particle fraction, therefore, smaller particles result in higher viscosity and reduce the mobility

of radicals during the polymerization, leading to a lower value of DC. These features limit the

high fraction loading of nanofillers. For this reason, nanofillers are normally blended with

micro-fillers (0.1~5 µm) to compensate for their drawbacks; these are often called hybrid dental

composites [102, 120]. Based on studies that have demonstrated that strength and modulus of

the composites are proportional to the filler loading [73], diverse combinations of fillers with

different weight fraction have been incorporated to develop the optimized flowable (< 60 wt%)

and packable (60 ~ 80 wt%) dental composites.

Hybrid composites adopted the incorporation of micro-sized fibers, whiskers, and nanotubes

to attain the increased requisites in mechanical properties such as strength, modulus, and

16
Figure 4. (a) Schematic illustration of the role of the functionalized fillers in dental restoration
composites. SEM images of filler materials showing the variation in size and geometry: (b)
silica colloidal nanoparticle clusters, (c) short S-glass fibers, (d) SiO2 nanofibers, (e) TiO2
nanotubes and nanoparticles, (f) dipentaerythritol penta-acrylate phosphate conditioned
zirconia (ZrO2) nanoparticle, (g) ZrO2 nanoparticles coated with mesoporous SiO2, (h)
calcium-rich zeolite, (i) surface modified urchin-like hydroxyapatite, (j) micro-sized clusters
of hydroxyapatite nanorods, (k) chitosan grafted halloysite nanotubes. Part (a) is adapted from
reference (Ref.) [102], (b) from Ref. [12], (d) from Ref. [93], (e) from Ref. [31], (f) from Ref.
[97], (i) from Ref. [24], and (k) from Ref. [102] with permission from Elsevier, part (c) from
Ref. [31], (g) from Ref. [25], and (h) from Ref. [107] with permission from American Chemical
Society (ACS), and (i) from Ref. [103] with permission from Springer Nature.

fracture toughness. Even though fiber reinforced composites are ubiquitous in many industries,

17
introduction into dental composite systems has been limited due to the relatively fine/short

fibers required. Recent advances in short fiber cutting techniques enable glass fibers to be

introduced in dental resin composites [31, 93, 121, 122]. Short glass fiber reinforced resin

composites exhibited an increase in flexural strength and fracture toughness [123]. Although

they showed potential for further development of short glass fiber reinforced dental composites,

exploration of the reinforcement mechanisms at the atomic level [13] and macroscale is

required to maximize the mechanical performance.

Hydroxyapatite whiskers, a naturally occurring mineral form of calcium apatite, have been

investigated as a bioactive and biocompatible filler, taking the additional advantages in

toughness and water absorption properties of the resulted composites [24, 78, 103, 124-126].

Halloysite nanotubes (HNTs) that are naturally occurring white color minerals have also been

employed in recent years to synthesize dental materials as shown in Figure 4. HNTs are the

new type of attractive additive especially in dental materials research because of their unique

physical and chemical characteristics, low cost in manufacturing, and good biocompatibility

compared to other types of synthesized nanotubes [102, 127-129].

2.4 Filler surface treatment

The prevalent failure mode of the composites under the impact and static loading is debonding

of the interface between the filler and resin. In reinforced materials, many filler materials such

as silica sphere, carbon fiber and glass fiber when used without or with improper surface

treatment result in composites with less yield strength than expected. This has been attributed

to poor bonding and delamination of the interface between the filler and matrix [130].

Mechanical performances of the dental composites are also strongly dependent on the strength

of the interfacial bond between the filler surface and the dental resins, and therefore it is

essential to form a high enough interfacial bond to guarantee desired properties of the dental

18
composites. The interfacial stress transfer is effectively performed when the interface is bridged

through the strong chemical interaction or mechanical binding, and consequently, it can

increase the mechanical and physical properties such as strength, toughness, wear, and water

absorption [131, 132]. On the contrary, if they fail to attain enough interfacial strength or

maintain a poor interfacial phase, it will increase the risk of failure of the composites because

a crack easily initiates first at the weakest interface and propagates along the unstable interface.

Mechanical interlocking, chemical bond via coupling agents, or a combination of the two can

provide a robust interfacial bond between the filler and the dental resin matrix.

In the mechanical interlocking [133], creating critical surface roughness of the filler materials

is key to improve the bonding strength between the filler and resin matrix by increasing contact

surface area in micro/nanoscale. Surface morphology can be basically modified by two

approaches, a top-down method using etching techniques and a bottom-up using coating and

grafting techniques as shown in Figure 5. The roughened filler surface increases the shear

resistance. Silica-based fillers, which are most widely used in dental material applications, can

be etched to increase the surface roughness in nanoscale by using strong acid or base solutions

such as hydrochloric acid (HCl), sulfuric acid (H2SO4), nitric acid (HNO3), phosphoric acid

(H3PO4), and sodium hydroxide (NaOH) [27, 134-137]. Surface topography can be measured

by using atomic force microscopy, and chemical composition change before and after the

etching process can be estimated under the X-ray photoelectron spectroscopy measurement.

Chemical etching leads to an increase in the density of hydroxyl groups (-OH) which

chemically react with coupling agents to anchor the polymer matrix (Figure 5(a)). In contrast,

a bottom-up approach is based on grafting and coating techniques, for instance, a different

inorganic nanomaterial or organic polymers can be chemically bonded on the filler materials.

Nano-hydroxyapatite grafted E-glass fibers increased the surface roughness of the fibers and

19
Figure 5. Schematic representations of surface functionalization methods on the various types
of dental filler materials, including (a) an atomic-level selective metal etching and SCA
grafting on short S-glass fibers, (b) nano hydroxyapatite grafting on E-glass fibers, (c)
methacrylate-polyhedral oligomeric silsesquioxane grafted SiO2 hybrid particle, (d) Bis-
GMA/TEGDMA grafting on polyacrylonitrile nanofibers, (e) 10-methacryloyloxydecyl
dihydrogen phosphate and dipentaerythritol penta-acrylate phosphate grafting on the nano-
zirconia filler, (f) functionalization on barium glass filler with thiourethane oligomers. Part (a)
is adapted from Ref. [27] and (c) from Ref. [40] with permission from ACS, (b) from Ref. [93]
and (e) from Ref. [97] with permission from Elsevier, (d) from Ref. [139] and (f) from Ref.
[143] with permission from Springer Nature.

enhanced the glass transition temperature and abrasion resistance of the resulted composites

[93]. Recent studies include polymer-grafted hybrid nanoparticles, such as methacrylate-

polyhedral oligomeric silsesquioxane on SiO2 [40], Bis-GMA/TEGDMA on polyacrylonitrile

nanofibers [138], and UDMA on ZnO [139], 10-methacryloyloxydecyl dihydrogen phosphate

and dipentaerythritol penta-acrylate phosphate on the nano-zirconia particles [97], have been

20
synthesized to enhance the interfacial bond between resins and fillers that is a key factor

influencing the overall physical and mechanical properties of DRCs.

To create a chemical (covalent) bond between organic and inorganic phases, silane coupling

agents (SCAs), which are synthetic hybrid organic-inorganic compounds, are widely used in

composite systems [140]. One end (typically an alkoxy substituent) of SCAs reacts with -OH

groups of the filler surface achieving optimal covalent bonds through the hydrolysis and

condensation kinetics. Surface silanized fillers are hydrophobic, enhancing wettability and

chemical compatibility with the monomer resin matrix. Full coverage of the fillers by

monomers without micro-/nano-voids on the filler surface is advantageous. Surface voids on

the fillers are undesirable in the composites since they later work as intrinsic defects degrading

mechanical properties of the composites. In a different manner, the other end of SCAs

(methacrylate, acrylate, epoxy, and amine reactive groups) would conjugate with functional

groups (C=C double bond) of polymer matrix during the photopolymerization process, thus

chemically bridging the interface. Accordingly, enhancement in the strength of the composites

would be achieved via the bifunctional SCA molecules such as (3-methacryloxypropyl)

trimethoxy silane (MPTS) and (3-aminopropyl) triethoxysilane (APTES) which create the

desired sound interface by firmly anchoring the micro-/nano-fillers with the resin matrix [141,

142]. In the same way, thiourethane-silanized fillers (Figure 5(f)) to DRC resulted in reduction

in polymerization stress and an increase of fracture toughness without compromising the

degree of conversion [143, 144].

The surface treatment can enhance the dispersion capability of nano-fillers. Nanostructures

have a large specific surface area and high surface energy. Due to this, they strive to reduce

their surface energy and tend to attract each other (agglomeration) composite industry. To

increase the dispersion of nano-fillers in the resin matrix, their surface properties can be

21
modified by using various mechanical and chemical treatments such as high-powered

sonication [145], micro-nozzle spray [146], plasma treatment [147], surfactant grafting [148],

water and organic solvent evaporation [11, 102], and pH level control [149]. Well-dispersed

nano-fillers enhanced the mechanical properties by impeding the propagation of micro-/nano-

crack in the composite, where the fillers serve as an excellent crack stopper for dispersion-

strengthening mechanisms, such as crack deflection, pinning, branching, and bridging.

3. Requirements and Characteristics of Dental Resin Composites

A longevity or survival rate, as a measure of the success of restorative dental treatments,

significantly depends on the true properties of the dental composites in physical, mechanical,

Figure 6. The required performance of dental resin composites in dentistry.

22
biological, and practical aspects (Figure 6) although patient- and dentist-related factors such as

patient’s occlusion and hygiene habits, characteristics of the restoration, and the dentist’s

clinical skills and treatment decisions are also key [150, 151]. Thus, when developing new

dental materials, their combined properties must be optimized to prevent physical failures and

recurrent caries in a harsh oral environment. The assessments of materials must be conducted

in the multiple aspects of inspection and characterization, and it can provide information to the

dentist as to which restoration material is most appropriate for a given situation.

3.1 Mechanical and Physical Characteristics

Elastic Modulus and Strength: The elastic modulus and strength of dental composites are

fundamental mechanical properties that are required for long-term restoration survival. The

elastic modulus represents an intrinsic mechanical property of the material and mathematically

quantifies a resistance against the applied loads (compression or tension) in the elastic

deformation region. The dental composites therefore can be expected to have sufficient

modulus values to resist the masticatory forces, and especially they are desired to possess

higher values in posterior restorations because the occlusal forces on the tooth are higher in the

posterior as compared to the anterior [152]. Also, the difference between the elastic moduli of

the layered restorations or from dental tissues will increase the possibility of interfacial failure

or marginal fracture. Ideally, dental restorations require attributes exhibiting a similar elastic

modulus to hard dental tissue so that the restoration will have equivalent stress and deformation

to the healthy tooth by minimizing the damage on the surrounding or opposite tooth under the

various occlusal loads. However, due to the intrinsic characteristics of polymer resins in DRCs,

typical modulus values are low when comparable to enamel. When compared to the moduli of

enamel and dentin, which are about 40-90 GPa and 11-20 GPa [153-155], respectively,

depending upon age, sex, and other factors, DRCs have much lower values (e.g., packable: 4-

23
20 GPa, flowable: 2-9 GPa) [74, 156-160]. The lower modulus of the restoration can potentially

protect the healthy tooth by sacrificing itself, however an ongoing area of research is to increase

modulus values of resin composites towards that of dentin or enamel, especially for use in

posterior high stress-bearing occlusal surface area.

The strength of a material is defined as its ability to withstand the applied load without failure

or plastic deformation. Unlike modulus, strength is not an inherent property of the material.

Internal and external factors such as material’s geometry, existing defects, manufacturing

process, and temperature influence the strength. Its value can be measured through various

methods, like tensile, compression, bend, and shear tests, but each measured value of the

material under the particular stresses applied can differ. For example, brittle materials such as

ceramics, glass, and enamel typically show an order of magnitude higher strength in

compression than in tension because the initiated cracks from the internal defects remain closed

under a compressive load, as opposed to tension in which defects may accelerate the crack

initiation and growth. It will not cause further crack propagation continuously, thereby

increasing the compressive strength [161]. The loading forces that are generated during routine

mastication can create diverse strain and stress (compressive, tensile, shear, bending, torsional)

at the level of the restorations. Restorative composites are synthesized with a variable

composition of resin and additives and show different performance against the applied loads.

For this reason, various mechanical tests on the restorative materials should be considered to

provide critical information to dentists to help them identify suitable materials in restorative

treatments. From the obtained strain-stress curve, the maximum stress that the material can

withstand without being fractured is defined as ultimate strength (or sometimes as breaking

strength). On the other hand, the yield strength represents the maximum stress beyond that the

material undergoes permanent deformation. In brittle materials such as ceramic and glass, the

ultimate strength is almost the same as the yield strength without showing plastic deformation.

24
Whereas in ductile systems such as polymers and metal, the ultimate strength can be higher

than their yield strength on account of strain hardening; that is, capacity to absorb mechanical

stress during the plastic deformation. The strength of dental restorations is commonly measured

with a flexural test (e.g., 3-point or 4-point bend tests) according to ISO 4049 [162], and the

typical flexural strength values of flowable and packable dental composites range from 67-181

MPa and 76-132 MPa, respectively [31, 75, 123, 157-159, 163]. The strength of DRCs with an

Table 3. Summary of available testing methods to determine fracture, deformation, and wear
resistance of dental resin composites.

Mechanical
Properties Testing methods Standards
requirements
3-Point bending ISO 4049
Biaxial flexural ASTM F394-78
Compression ASTM D695
Strength / Tensile ASTM D638
Modulus Diametral tensile ANSI / ADA Spec 27
Impact ISO 179/1961
Transverse impact DIN 53 453
Shear punch ASTM D732
Linear elastic fracture
Fracture and ISO 13586
mechanics
deformation
Fracture Plane-strain (Chevron-
resistance ASTM E1304-97
toughness notch)
Single-edge notched beam ASTM D5045
J-R Curves ASTM 6068
Calculated from strength
Toughness -
tests
Fatigue strength - staircase -
Strain controlled fatigue
Fatigue ISO 15850
testing: tension-tension
Strain controlled fatigue
ASTM E606 / E606M
testing: tension-compression
Indentation ISO 14577-1
Surface Vickers ISO 6507-1 / ASTM E384
hardness
Surface / Modulus Knoop ISO 6507-1 / ASTM E384
abrasion Rockwell ISO 2039-2
resistance
Wear by toothbrushing ISO TS No. 14569-1
Wear Wear by two- and/or three
ISO TS No. 14569-2
body contact

25
increased filler content up to a certain limit showed more rigid and strong, but with brittle

characteristics. Above a critical amount of filler, it will lead to strength degradation with a

formation of poor interfacial bond between resin and fillers or aggregation/agglomeration of

fillers that work as new defects in the composite system [164].

The elastic modulus and strength of the dental composite systems can be controlled through

resin type, shape and size of the filler, ratio of resin and filler, filler dispersion and orientation,

and interfacial adhesion between filler and resin. It should be noted that, in the harsh

environment of the mouth, several different stresses may be combined and applied on

restorations or restoration-tooth interface. Hence, many different types of testing methods

(Table 3) such as bending, compression, shear, impact, and diametral tensile test are necessary

to optimize material properties and to provide essential information to dentists for the best

material and technique selection during dental restoration [165, 166].

Fracture toughness: Many prospective clinical studies demonstrated that fracture of

restoration is the main cause of restoration failure, followed by recurrent caries near the

restoration margins, over a 17-year follow up period [167-169]. Fracture of the restoration or

the tooth occurred earlier, more than 60% during the first 3-year period, and caries occurred

later, more than 75% after 3 years in service [170]. Thus, one of the essential mechanical

characteristics used to assess reliability of restorative materials is fracture toughness, which

describes the ability of the material to resist fracture, typically calculated in terms of stress

intensity near the preexisting crack tip of known size expressing a relationship between the

applied stress and the length of crack. The value of the critical stress-intensity or plane-strain

fracture toughness, KIC, represents the level of fracture strength of material in units of MPa·m1/2

or MN·m-3/2 [171], and unlike strength it is an inherent property of the material. Various

fracture toughness test methods, such as single-edge-notched beam (SENB), chevron-notched

beam, compact tension, disk-shaped compact tension, double torsion method, and the

26
indentation fracture method, have been applied to evaluate the relative fracture toughness of

resin composites [172, 173]. Each method requires different specimen geometries associated

with sizes and shapes of preexisting crack and loading configurations. Because fracture failure

occurs where the composite restoration is subjected to various loading conditions (e.g., bending,

compression, tension, shear, torsion), it is difficult to design an optimal testing method for

dental restorative composites. Showing a good reliability for most dental restorative material

applications, the SENB flexural test has been widely used because of its good reproducibility,

less material use, and simple sample fabrication method. Compared to metal or brittle ceramic

materials, it is relatively easy to produce an accurate shape of the pre-crack in the composite

samples, especially for photocurable dental composites, where the crack can be shaped by

positioning the sharp-edged blade in the middle of the custom-made mold [174, 175].

In general, fracture toughness of dental tissues ranges 0.7-1.27 MN·m-3/2 and 3.1 MN·m-3/2 for

enamel and dentin, respectively [176]. Dentin has a greater toughness value than enamel

because it possesses a much higher organic content. Most of the commercial DRCs exhibit

somewhat low fracture toughness value than dentin in the range of 0.37-2.4 MN·m-3/2 [177].

To prevent bulk fracture of the restoration, fracture toughness (KIC) can be improved by

optimizing the composition of filler materials, employing fiber shape fillers, and also

increasing interfacial bonding between filler and resin. Among them, the most used approach

to achieve improvements in fracture toughness is by addition of filler particles up to the critical

fraction. In a study of the commercial DRCs, the composites possessed highest KIC value when

they are composed with 57-65% volume fraction of filler materials [177]. Above this value,

KIC decreased slightly which can be resulted by the increased internal defects, such as air voids,

interfacial flaws, broken fillers, and filler agglomerates, being integrated in the composites and

by the reduced polymerization of the monomers in the highly packed composite systems. An

increase in toughness by employing the short glass fibers into the dental composites has been

27
observed, in which fibers worked as stress absorbance according to fiber-bridge toughening

mechanism, also by providing additional reinforcement, crack-pinning and deflection [27, 95,

171, 178-183].

Polymerization shrinkage: The polymerization process in dental composites typically results

in polymerization shrinkage (PS), induced by chemical, thermal or post contraction effects.

The effects of PS on dental restorations are vital given the growing demands for longevity and

durability of polymer dental materials, where the mechanical stresses that develop within the

material during curing and lifespan are a key determinant. Molecular densification and related

macroscopic effects of PS strain and/or stress influences the material’s post-cure characteristics,

and thus higher shrinkage leads to higher shrinkage stresses and lower bonding efficiency,

reducing the mechanical and physical properties of the restorative material leading to

microleakage and failure of the restoration [184-188].

As PS is a vector quantity and PS strain/stress distribution across a material depends on the

excitation/initiation mode, the shrinkage pattern is often anisotropic. Different approaches were

developed to manage the shrinkage of biomaterials, such as low modulus intermediate

materials, different polymerization initiation mechanisms and alternative excitation methods

to delay curing kinetics [189-192].

Depending on the type of measurement approach, PS measurement methods can be classified

into three categories, average (or total) volumetric PS, post-gel PS and PS profile. Using these

techniques, typically the magnitude and kinetics of PS are measured. However, it is important

to note that most of PS measurement techniques are indirect methods and can vary when the

polymerization conditions change. Therefore, the polymerization conditions must be known to

compare the PS strain/stress of dental composites measured using these techniques.

28
Table 4. Major polymerization shrinkage measurement methods.

Of the indirect polymerization and polymerization kinetics measurements, volumetric PS

measurement is the most common method, where it provides the total/average volumetric PS

of the dental composite. Various methods exist to measure volumetric PS of dental resins as

shown in Table 4. Of these, bonded disk, linometer, dilatometer and Archimedes methods are

well established and commonly used in industry and academia [193].

Total PS in a dental composite could be divided into pre-gel (viscous to gel) and post-gel (gel

to solidification) phases. During pre-gel phase the material still possesses fluidity and

flowability, i.e., any shrinkage/contraction would result in deformation of the material and with

negligible stress. However, when the crosslinking reaches its apex in post-gel phase, the gel is

converted to solid (higher elastic properties), and any contraction would result in minimal

deformation and maximum shrinkage stress. Post-gel PS measurement methods neglect any

deformation of viscous flow in pre-gel phase and measure the clinically significant post-gel

deformation. Multiple studies have confirmed the usage of strain-gauges for the efficient

measurement of linear PS [194]. Another emerging method to measure the post-gel PS is using

optical fiber methods, where an optical fiber Bragg grating (FBG) is embedded inside the dental

composite, where it measures linear PS strain arising from the post-gel shrinkage [195].

29
Figure 7. Evolution of PS strain within a dental composite measured using a CFBG sensor.
Figure is adapted from Ref. [198] with permission from Springer Nature.

The third category of PS measurement method is the PS profile measurement approach, which

is a relatively new method. Traditional PS measurement systems mentioned above only

provides an overall PS measurement, but the true PS may vary within the material. It has long

been known that the PS and DC are correlated to each other for photo-cured dental composites

and is also non-uniform, but only recent developments in measurement techniques have

directly revealed such non-uniformity [195, 196]. This has led to the need to measure the PS

profile to understand the true shrinkage behavior of the material. One method for PS profile

measurement is the digital image correlation technique where it provides full field shrinkage

strain in real time but faces limitations in time and strain resolution due to the nature of the

technique relying on visually tracking physical movements on the sample surface [197].

Another recently demonstrated method for post-gel PS profile measurement is using an optical

chirped fiber Bragg grating (CFBG) embedded inside the dental composite, where it provides

the PS strain across the material in the direction of the embedded CFBG [198]. Figure 7 shows

the evolution of PS strain within the material measured by a CFBG sensor, where the shrinkage

30
is higher at the center region compared to the edges, which correlates to the curing lamp

intensity distribution and degree of conversion.

3.2 Biological characteristics

Recurrent caries usually develops at the restoration margin, which is the boundary between the

restoration material and adjacent dental tissue. Thus, it is important to understand the physical

and biological interactions between restorations and tooth tissue to enhance the lifetime of the

restoration. The interface is a physically weak area, and micro-leakage of bacteria and

chemicals into the propagated cracks along the interface can result in colonization of salivary

bacteria and carious lesions [199, 200]. Also, constant exposure to caustic acids can accelerate

demineralization of tooth tissue increasing the risk of developing tooth decay [201, 202]. One

of the effective ways for the prevention of recurrent caries is to add functionalized fillers/agents

in the restorations to constantly apply antibacterial and remineralization actions around the

restoration/tooth interface. Much research has been done to improve the biological properties

of DRCs, mainly focusing on improving antibacterial and remineralizing capabilities.

Antimicrobial ability: Antimicrobial activity of the dental restorative composites is attributed

to two major mechanisms, (1) interacting with the outer membrane of the oral bacteria and

causing cell membrane lysis or repelling bacterial attachment, and (2) interacting with bacterial

proteins and DNA and disrupting protein/DNA synthesis (Figure 8) [8, 45, 203, 204].

Considering these biological kinetics, restorative dental composites have been synthesized by

incorporating various types of antibacterial agents and filler materials. They can be categorized

into two main classes: an ion-releasing bioactive filler and non-releasing filler. The composites

integrating the former type of fillers, such as metal nanoparticles (Ag, Cu, TiO2, ZnO) [44, 99,

205-209] and surface pre-treated particles with fluoride [210], chlorhexidine [83, 211, 212],

nitrogen [213], and chitosan [102, 214] that can be ionized in the presence of water, practically

31
Figure 8. Schematic illustration of antimicrobial activities of the functionalized dental
composite [102].

exhibit high antibacterial activity. The water solubility of the released ions can affect the space

of antibacterial zone. The release of antimicrobial agents, however, will degrade the

mechanical properties of the composite surface by generating nanopores or interfacial

delamination and sometimes change the shade of restorations [215]. Finally, the antibacterial

ability will gradually become inactive. Thus, time- and space-controlled release of antibacterial

agents from dental composites has been studied extensively. Accordingly, to overcome these

drawbacks of ion releasing composites, the second type of agent, a non-releasing (or

immobilized) agent, is emerging because it is a more bio-sustainable and somewhat

environmentally friendly approach [41, 216-218]. Once the positively charged fillers attract the

32
negatively charged oral bacteria, they disrupt the structures and functions of bacterial cell

membrane inducing cell lysis, which is described as ‘contact-killing mechanism’ [208, 218].

Quaternary ammonium compounds (QACs) are the best-known polymeric biocides as they

become cationic (positive) surface-active agents in water solution [219-221]. One of the

suggested kinetics of antimicrobial activity for QACs is that the substituted nitrogen part of the

molecule attracts the bacteria, and a long, lipophilic alkyl chain penetrates bacterial cell

membranes leading to a leakage of the cytoplasmic material and a loss of viability or cell death.

Moreover, by using a natural betulin biocompound, monomethacrylate- and dimethacrylate-

functionalized betulin derivatives (M1Bet and M2Bet) were synthesized as antibacterial

comonomers. The partial replacement (10 wt %) of Bis-GMA by M1Bet and M2Bet enhanced

flowability, polymerization rate, cell viability, and antibacterial activity against S. mutans

without compromising mechanical properties [222]. In addition, several studies on the

modification of DRCs have been conducted by developing new non-degradable antimicrobial

agents using 2-methacryloyloxyethyl phosphorylcholine polymers [41, 223-225]. They

established novel methodologies to inhibit biofilm formation on the composite surface

embedding protein-repellent functions.

The human mouth is a complex environment and contains a wide range of microorganisms, but

only a few bacterial species are associated with dental caries (Table 5). It is believed that

Streptococcus mutans (S. mutans), S. sanguinis, and Actinomyces are most prevalent bacteria

that form biofilms with an increase of pH-lowering and cause caries [226, 227]. In addition,

Lactobacillus species are actively involved in the microbial colonization and plaque formation.

They are characterized by their ability to survive in an acidic environment and to produce lactic

acid that can cause dental caries with the metabolic cooperation with S. mutans [228].

Porphyromonas gingivalis plays a critical role in initiation of dental plaque and it is highly

associated with the periodontal disease [229, 230]. Therefore, understanding of the processes

33
that lead to bacterial biofilm formation and controlling of their activities to suppress microbial

life are most important to prevent caries and periodontal disease, and for this reason, these

bacteria have been most frequently used for performing antibacterial testing.

Remineralization: Many studies have been done recently on the specific bioactive responses

of dental restoration composites to the process of tooth tissue development. Among the many

effective bone remineralizing agents, calcium phosphates have been extensively used as ion-

Table 5. List of oral bacteria causing dental disease.

Gram stain Bacteria Shape Growth requirements


Actinomyces naeslundii Bacillus Facultatively anaerobic
Actinomyces odontolyticus Bacillus Facultatively anaerobic
Actinomyces viscosus Bacillus Facultatively anaerobic
Bacillus subtilis Bacillus Aerobic
Candida albicans Fungus Facultatively anaerobic
Enterococcus faecalis Coccus Facultatively anaerobic
Enterococcus faecium Coccus Facultatively anaerobic
Lactobacillus acidophilus Bacillus Facultatively anaerobic
Lactobacillus casei Bacillus Facultatively anaerobic
Parvimonas micra Coccus Obligately anaerobic
Gram-positive
Peptostreptococcus Coccus Obligately anaerobic
Staphylococcus aureus Coccus Facultatively anaerobic
Streptococcus anginosus Coccus Facultatively anaerobic
Streptococcus constellatus Coccus Facultatively anaerobic
Streptococcus gordonii Coccus Facultatively anaerobic
Streptococcus intermedius Coccus Facultatively anaerobic
Streptococcus mutans Coccus Facultatively anaerobic
Streptococcus oralis Coccus Facultatively anaerobic
Streptococcus sanguinis Coccus Facultatively anaerobic
Streptococcus sobrinus Coccus Facultatively anaerobic
Escherichia coli Bacillus Facultatively anaerobic
Fusobacterium Bacillus Obligately anaerobic
Gram-negative Porphyromonas gingivalis Bacillus Obligately anaerobic
Prevotella melaninogenica Bacillus Obligately anaerobic
Pseudomonas aeruginosa Bacillus Facultatively anaerobic

34
releasing fillers in various orthopaedic and dental applications, such as bone graft substitutes,

orthodontic cements, adhesives, sealants, dentifrices, and restorative composites, due to their

excellent bioactive and osteoconductive properties. They can enhance bone repair and

mineralization capability increasing the durability of the clinical treatments, e.g., epitaxial bone

growth on the surface [231-233], tooth repair [105, 124, 234-237], and biological bond

formation between the composite and tissue [238-240]. Due to the chemical similarity of the

mineral phase in bones, especially of enamel tissue, the artificially manufactured calcium

phosphates provided excellent biocompatibility and bone regeneration ability with the

biochemically mediated bonding osteogenesis [241]. The various types of synthetic calcium

phosphate materials [102, 103, 242-245], such as monocalcium phosphate (Ca(H2PO4)2),

dicalcium phosphate (CaHPO4), tricalcium phosphate (α-, β-TCP; Ca3(PO4)2), octacalcium

phosphate (Ca8H2(PO4)6·5H2O), amorphous calcium phosphates (CaxHy(PO4)z·nH2O),

hydroxyapatite (Ca10(PO4)6(OH)2), whitlockite (Ca9Mg(HPO4) (PO4)6), fluorapatite

(Ca10(PO4)6F2), and biphasic calcium phosphate (a mixture of hydroxyapatite and β-tricalcium

phosphate), showed a remarkable effect on the remineralization of enamel and dentin. Calcium

(Ca+2) and phosphate (PO43-) ion-releasing composites can promote the nucleation and growth

of minerals crystals with a protein matrix, which mimics the natural regeneration of the tooth.

However, as different calcium/phosphate compositions exhibit different mechanical properties

and ion releasing rates, hybrid variants of calcium phosphates can moderately enhance the

performance of the restorative composites in terms of mechanical and biological characteristics

in a variety of applications [246].

Bioactive glass [247], which is mainly composed of SiO2-Na2O-CaO- P2O5 system, is another

group of popular bio-reactive filler material for bone tissue regeneration applications. By

adjusting the ratios of constituent components, these compositional features can modulate their

specific surface degradation rate to induce a highly reactive surface to an aqueous medium and

35
tissue. Even though bioactive glasses exhibit lower mechanical properties in strength, modulus,

and toughness compared to the typical range of glass fillers, they can be incorporated in DRCs

to promote remineralization of enamel and dentin showing good biocompatibility [248-252].

Also, at the tissue/restoration interface, the released ions of calcium and phosphate from the

bioactive glass stimulate the tooth surface to form a strong bond with restoration by

regenerating healthy tissue that will increase the longevity of dental restorations preventing

microbial ingress as well.

Fluoride has been suggested extensively as an effective anticariogenic agent in dental research

and dentistry [253]. It is associated with effects both on antibacterial activity by reducing the

acid tolerance of the bacteria [254] and on remineralization of the calcified tissues due to its

high affinity for calcium [255]. Some studies devote their efforts to develop fluoride-containing

bioactive particles. CaF2/SiO2 core-shell nanoparticles were synthesized with a sol-gel method

based on the CaF2 nanoparticles and the DRCs were prepared with Bis-GMA/TEGDMA and

50 wt% of CaF2/SiO2 nanoparticles which showed superior mechanical properties in flexural

strength, flexural modulus, compressive strength, and hardness and sustained fluoride ion

release [256]. In another study, LiAl-F layered double hydroxide (LDH) powder was

synthesized and their performance was examined by mixing with commercial DRCs [257]. The

mechanical properties, color change, ion leaching, cytotoxicity, and fluoride release of the

composites proved the LiAl-F LDH as a potential filler for the multifunction DRCs. Also,

MgAl and CaAl LDH composites were synthesized [258] and their capability to absorb and

release a sustained low-level fluoride demonstrated the potential to prevent carious lesions and

promote tooth remineralization. Several attempts to develop new DRCs with fluoride released

fillers such as CaF2 [210, 259, 260], NaF [261], and SrF2 [262] have also been made in recent

studies.

36
4. Products and Applications

Market development: Since their widespread acceptance during the 1970s, dental composites

have evolved and branched-off into the vast array of products now available to the modern

clinician. The driving force for the development of dental composites is through a combination

of improved materials technology, solving clinical challenges and listening to clinicians’

personal preferences. A list of a variety of DRCs currently available in the market is

summarized and the major resins and filler materials used in the flowable and packable DRCs,

which are grouped to five different composite categories by the filler type and fraction, are

demonstrated on Table 6. However recently, flowable and bulk filled composites are emerging

as the preferred option for many clinicians. First launched in the 1990s, flowable composites

were initially considered inferior to packable composites due to their high polymerization

shrinkage, higher wear, and lower mechanical properties. However, flowable products are now

used extensively in the dental setting due to their ease of use and broad clinical application.

Mechanical properties of flowable composites have also improved significantly, such that

many commercial products now rival the properties of high-strength universal composites.

With the technological maturity of current dental composites, many manufacturers are now

designing materials to improve the user experience. The maturity of the product development

pathway also demands significantly greater control, validation, and product demonstration. A

simplified Product Development Pathway is illustrated in Figure 9.

For an example of enhanced user experience, universal composites have been commercialized

that claim to match the shade of all patients with one product. Another example is that of bulk-

fill composites which are designed to restore large cavities in one placement step instead of the

traditionally used incremental layering technique. These advanced products simplify dental

procedures and provide a range of restorative options available to the dentist.

37
Natural tooth structure is a very durable material, and the ultimate goal is to retain the natural

dentition for as long as possible. This has led to a focus on MI dentistry that is expected to

further develop materials in the future, including dental composite products, that are better

suited to restoring whilst retaining as much natural tooth as possible.

38
Table 6. List of commercial dental resin composites (Flowable).

Type Classification
Product name Manufacturer Resins Fillers
(physical) (intrinsic)
TMPTMA, UDMA, Bis-GMA, Zinc oxide, dental glass, amorphous
Fill-Up Coltene
TEGDMA silica
Bis-GMA, Bis-EMA,
Tetric Ivoclar- Ytterbium trifluoride (YbF3)
tricyclodecane dimethanol
PowerFlow Vivadent particle, barium glass
Hybrid dimethacrylate
composite Bis-GMA, UDMA,
IPS Empress Ivoclar- YbF3 particle, barium glass, SiO2-
tricyclodecane dimethanol
Direct Vivadent ZrO2 mixed particle
dimethacrylate
UDMA, TEGDMA, Bis-MEPP, Short glass fiber (E-glass), barium
EverX Flow GC
DMA monomer glass and SiO2 particle
Constic DMG DMA monomer, Bis-GMA Barium glass
Microcomposite Heliomolar Ivoclar-
Flowable Bis-GMA, UDMA, TEGDMA YbF3 particle, SiO2 particle
Flow Vivadent
Filtek Bulk Fill UDMA, Bis-GMA, Bis-EMA, Zirconia and silica Filler, YbF3
3M
Microhybrid Flowable TEGDMA particle
composite
EcuSphere-Flow DMG DMA resin, Bis-GMA Dental glass
Filtek Supreme BisGMA, TEGDMA, Procrylat YbF3 particle, nano silica particle,
3M
Ultra Flowable resins nanozirconia/silica cluster
PALFIQUE Bis-GMA, Bis-MPEPP,
Tokuyama SiO2-ZrO2 nano spherical filler
Nanocomposite Universal Flow TEGDMA, UDMA
PALFIQUE Bis-GMA, Bis-MPEPP,
Tokuyama SiO2-ZrO2 nano spherical filler
Bulk Flow TEGDMA
ESTELITE
Tokuyama Bis-MPEPP, TEGDMA, UDMA SiO2-ZrO2 nano spherical filler
Flow Quick

39
Synergy D6 / Silica, barium glass,
Coltene Bis-GMA, UDMA, TEDGMA
Flow pre-polymerized filler
Barium-alumino-fluoro-borosilicate
Dentsply glass, strontium alumino-fluoro-
SDR Flow+ TEGDMA, UDMA
Sirona silicate glass; fume silica, YbF3
particle
Ivoclar- YbF3 particle, barium glass,
Tetric EvoFlow Bis-GMA, UDMA, D3MA
Vivadent SiO2-ZrO2 mixed particle
Herculite Ultra
Kerr BPA-EDA, TEGDMA, Bis-GMA YbF3 particle, SiO2 particle
Flow
HEMA, UEDMA, GDMA, BPA-
Vertise Flow Kerr YbF3 particle, SiO2 particle
EDA
Beautifil Flow Bis-GMA, Bis-MPEPP,
Shofu Aluminofluoro-borosilicate glass
Plus X TEGDMA
Nanohybrid Aluminofluoro-borosilicate glass,
FIT SA Shofu UDMA, HEMA
composite zirconium silicate, glass powder
Beautifil-Bulk Bis-GMA, UDMA, Bi-MPEPP,
Shofu Aluminofluoro-borosilicate glass
Flowable TEGDMA
Barium aluminium borosilicate
VisCalor VOCO Bis-GMA, TCDDMA
glass, silica particle
Admira Fusion HEDMA, Bis-GMA, UDMA,
VOCO Ormocer
Flow TEGDMA
Wave SDI Ltd UDMA, TEGDMA, Bis-EMA Strontium glass, silica
Barium glass, ytterbium trifluoride,
Luna Flow SDI Ltd UDMA, TEGDMA, Bis-EMA
silica
Barium glass, ytterbium trifluoride,
Aura Easyflow SDI Ltd UDMA, TEGDMA, Bis-EMA
silica
Venus Diamond Barium aluminium fluoride glass,
Kulzer UDMA, EBADMA
Flow ytterbium-fluoride, silicium oxide

40
Table 6. (continued) List of commercial dental resin composites (Packable).

Type Classification
Product name Manufacturer Resins Fillers
(physical) (intrinsic)
Brillant
Coltene Bis-GMA, TEGDMA Zinc oxide, Ytterbium trifluoride
EverGlow
Miris 2 Coltene Bis-GMA, UDMA, TEDGMA Zinc oxide
Bis-MEPP, TEGDMA, Sphere TEC spherical pre-
Dentsply
CeramX Spectra Urethane modified Bis-GMA polymerised filler,
Sirona
Hybrid dimethacrylate YbF3 particle
composite Dentsply
Surefil one MOPOS, BADEP, Acrylic acid Glass filler
Sirona
Bis-GMA, UDMA, Bis-PMA,
Ivoclar- YbF3 particle, barium glass,
Tetric PowerFill Bis-EMA, tricyclodecane
Vivadent SiO2-ZrO2 mixed particle
dimethanol dimethacrylate
Ivoclar- UDMA, Bis-EMA, Bis-GMA, YbF3 particle, barium glass, SiO2-
Packable Tetric Prime
Vivadent D3MA ZrO2
EcuSphere-Carat DMG DMA resin, Bis-GMA Dental glass
Microcomposite Ivoclar-
Heliomolar / HB Bis-GMA, UDMA, D3MA YbF3 particle, SiO2 particle
Vivadent
Filtek Z250 UDMA, Bis-GMA, Bis-EMA, Silane treated ceramic, aluminium
3M
Microhybrid Universal TEGDMA oxide
composite Filtek P60 UDMA, Bis-GMA, Bis-EMA,
3M Zirconia and silica filler
Posterior TEGDMA
Nano silica particles, Nano zirconia
Filtek Supreme Bis-GMA, UDMA, Bis-EMA,
3M particle,
Ultra Universal PEGDMA, TEGDMA
Nanocomposite zirconia/silica cluster
Filtek One Bulk AUDMA, AFM, DDDMA, YbF3 particle, silica filler,
3M
Fill UDMA zirconia, silica/zirconia particle

41
ESTELITE
Tokuyama Bis-GMA, TEGDMA SiO2-ZrO2 nano spherical filler
Sigma Quick
Paradigm Nano Bis-GMA, UDMA, Bis-EMA, Nano silica particle, nano
3M
Hybrid Universal PEGDMA, TEGDMA zirconia/silica cluster
Ecosite Bulk Fill DMG DMA monomer, Bis-GMA Barium glass
EBPADMA, Bis-GMA,
Ecosite Elements DMG Barium glass
TEGDMA
Venus Diamond Kulzer TCD-DI-HEA, UDMA Barium aluminium Fluoride glass
Barium aluminium Fluoride glass,
Venus Pearl Kulzer TCD-DI-HEA, UDMA
pre-polymerized filler
Ivoclar- YbF3 particle, barium glass, SiO2-
Tetric EvoCeram Bis-GMA, UDMA
Vivadent ZrO2
Harmonize Kerr BPA-EDA, TEGDMA Spherical silica, zirconia particle
Nanohybrid
BPA-EDA, UEDMA,
composite Barium glass, silica nanoparticle,
Premise Kerr 1,6-hexanediyl bismethacrylate,
pre-polymerized filler
HDDA
Aluminofluoro-borosilicate glass
Beautifil II Shofu Bis-GMA, TEGDMA
(pre-reacted glass-ionomer), Al2O3
Beautifil-Bulk Bis-GMA, UDMA, Bis-MPEPP,
Shofu Aluminofluoro-borosilicate glass
Restorative TEGDMA
HEDMA, Bis-GMA, UDMA,
Admira Fusion VOCO Ormocer
TEGDMA
Barium glass, pre-polymerised
Aura SDI Ltd UDMA, TEGDMA, Bis-EMA
filler, silica
Luna SDI Ltd UDMA, TEGDMA, Bis-EMA Strontium glass, silica
UDMA, tricyclodecane Strontium glass, ytterbium
Luna 2 SDI Ltd
dimethanol dimethacrylate trifluoride silica

42
UDMA, Bis-EMA, DX-511 Fluoroaluminosilicate glass, pre-
Kalore GC
monomer, DMA monomer polymerized filler, silicon dioxide
UDMA: urethane dimethacrylate / 1,6-bis(methacrylyloxy-2-ethoxycabonylamino)-2,4,4-trimethylhexane; AUDMA: aromatic urethane
dimethacrylate; DUDMA: diurethane dimethacrylate; TEGDMA: triethyleneglycol dimethacrylate / 2,2´-ethylenedioxydiethyl
dimetharcylate; Bis-EMA: bis-phenol A ethoxylated dimethacrylate; Bis-GMA: bisphenol A diglycydyl ether dimethacrylate / 2,2-bis[4-2-
hydroxy-3-methacryloxypropoxy)phenyl]propane; Bis-MEPP: 2,2[prime]-bis[4-methacryloxy-ethoxy)-phenyl-propane; HEMA: 2-
hidroxyethyl methacrylate; HEDMA: hydroethyl dimethacrylate; DMA: dimethacrylate; TCD-DI-HEA: bis-
(acryloyloxymethyl)tricyclo[5.2.1.02,6]decan; D3MA: 1,10-decanediol dimethacrylate; TMPTMA: trimethylolpropane trimethacrylate;
BPAEDA: bisphenol A ethoxylate diacrylate; HDDA: hexanediol diacrylate.

Figure 9. Simplified New Product Development Pathway, mapped against Manufacturing Readiness Levels (MRLs).

43
Manufacturing - 3D printing: The fast evolution of digital technologies in dentistry, such as

3-dimensional (3D) scanning and imaging, computer aided design and simulation, and 3D

printing, has assisted dentist to make correct diagnostic-based decisions and enabled patients

to receive accurate dental treatments in a more informative and speedy manner [263, 264].

Among the various 3D digital dental system, 3D printing technologies using light-curable

resins (e.g., stereolithography apparatus (SLA), digital light processing (DLP), and

lithography-based ceramic manufacturing (LCM)) are expanding into the dental industries

because of their high accuracy and speed, cost-effectiveness, and good productivity even in

dental office and laboratory. Small and complex structures, including indirect restorations,

bridges, dentures, aligner, implant abutments, and surgical guides can be fabricated based on

the patient intra-oral and tooth data [265-269]. The selection of appropriate photocurable

monomer is most important for successful 3D printing, and then functionalized filler materials

can be mixed with resin to increase their mechanical and biological properties.

DLP-3D printable dental composites were synthesized by adding Ag-HNT nanofiller (1, 2, 3

wt%) with a mixture of UDMA and TEGDMA (6:4) as depicted in Figure 10 (a) [270]. The

composite could release Ag+ and improve the antibacterial properties while also maintaining

non-toxicity. UDMA-based resins were also studied to develop the occlusal splint and crown

applications (Figure 10(b)) [271, 272]. Further attention to developing new manufacturing

method was paid, and a polymer-infiltrated ceramic network (Figure 10(c)) was successfully

developed through combining 3-D printing and dental resin infiltration technologies [273]. In

this study, 3D-printable precursor slurry was prepared by mixing a high concentration of silica

(SiO2) nanoparticles with HEMA/TEGDMA monomers. Then, the SLA 3D-printed composite

was sintered at high temperature (1,150 °C) to create a nano-porous structure and subsequently

infiltrated and polymerized with dental resin monomers, UDMA and TEGDMA, which

resulted in the increased surface hardness and flexural modulus. The feasibility and

44
Figure 10. SLA and DLP 3D-printed dental structures using dental resins and composites. (a)
Schematic diagram of antibacterial Ag-HNT composite preparation and 3D printed dental
crown and bridge, (b) fabrication of a crown using UDMA-based monomer, (c) polymer
(UDMA/TEGDMA)-infiltrated ceramic network (PICN) composite crown, (d) an accuracy
measurement of the 3D printed crown composed of Bis-EMA/UDMA/TEGDMA resins. Part
(a) is adapted from Ref. [270] with permission from Springer Nature, (b) from Ref. [271] with
permission from MDPI, (c) from Ref. [273] with permission from SAGE Publications, (d) from
Ref. [274] with permission from Elsevier.

applicability of dental monomers such as Bis-EMA, UDMA, and TEGDMA as 3D printing

resins using ultraviolet-DLP 3D printers have been explored, and the studies in mechanical and
45
biological properties and accuracy of printed objects demonstrated a great potential for digital

dentistry [274, 275]. A biocompatibility of various dental resins has been investigated by

assessing the post curing monomer release of 3D-printed structures in methanol and water

[276].

Moreover, in other recent studies, the fit of interim crowns fabricated using 3D printing with

acrylic-based biocompatible polymer (VeroGlaze MED620; Stratasys) has been studied, and

3D-printed crowns showed an excellent fit in the proximal, marginal, and internal regions

compared to those made by milling and compression molding methods [241]. A new approach

for direct restoration of maxillary central incisors applying 3D-printed template was introduced

showing the enhanced aesthetic appearance of the tooth [277]. In future, the scope of 3D

printings/bioprinting for the regenerative dentistry and craniofacial bone tissue engineering

will be gradually extended to achieve superior performance and better esthetics with advanced

3D digital technologies and newly emerging biocompatible dental resins [278-282].

5. Conclusion

This review has summarized recent advances in DRCs and approaches to enhance their

functional performance in mechanical, physical, and biological aspects. The recent research

trend of using fillers and resins in dental restorative composites was also discussed in this

review, and silica glass materials including bioactive and fibrous types showed the most

frequent usage (37.6% among the 157 studied cases) as a filler material. With the evolution of

nanomaterial technologies, nanocomposite system synthesized with various nanofiller

materials, such as nanoparticles (crystals and clusters), nanotube, nanofibers (rods and wires),

and nanosheets, can provide a huge impact on the clinical performance of dental restorations

and quality of manufacturing of dental products. Also, surface treatment technologies on

nanomaterials to increase interfacial bonding between filler and resin and to add biological

46
functionalization such as antibacterial activities and remineralization have been explored to

develop multi-functionalized restorative materials. Although there are still many scientific

progresses must be done in the development of new dental restorative materials, a combined

quantitative analyzing method in material properties and integration of fast-growing digital

technologies and dentistry information will inspire many researchers in new product

development and dentist in the dental care sector.

CRediT authorship contribution statement

K. Cho: Conceptualization, Writing - original draft preparation, review & editing,

Visualization. G. Rajan: Writing - original draft preparation, review & editing. P. Farrar:

Writing - original draft preparation, review & editing. L. Prentice: Writing - review & editing.

B. G. Prusty: Conceptualization, Writing - review & editing.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

This research was supported by AusIndustry funded CRC-Project (CRCPSEVEN000013) with

SDI Ltd. The authors also acknowledge the support received from the ARC Training Centre

for Automated Manufacture of Advanced Composites (IC160100040), supported by the

Commonwealth of Australia under the Australian Research Council’s Industrial

Transformation Research Program.

47
References

[1] Pratap B, Gupta RK, Bhardwaj B, Nag M. Resin based restorative dental materials:
characteristics and future perspectives. Jpn Dent Sci Rev. 2019;55(1):126-38.
[2] Zhang K, Zhang N, Weir MD, Reynolds MA, Bai Y, Xu HHK. Bioactive dental composites
and bonding agents having remineralizing and antibacterial characteristics. Dent Clin North
Am. 2017;61(4):669-87.
[3] Demarco FF, Collares K, Correa MB, Cenci MS, Moraes RR, Opdam NJ. Should my
composite restorations last forever? Why are they failing? Braz Oral Res. 2017;31(suppl 1):e56.
[4] Eltahlah D, Lynch CD, Chadwick BL, Blum IR, Wilson NHF. An update on the reasons
for placement and replacement of direct restorations. J Dent. 2018;72:1-7.
[5] Broadbent JM, Murray CM, Schwass DR, Brosnan M, Brunton PA, Lyons KS, et al. The
dental amalgam phasedown in New Zealand: A 20-year rrend. Oper Dent. 2020;45(3):255-64.
[6] Lynch CD, Farnell DJJ, Stanton H, Chestnutt IG, Brunton PA, Wilson NHF. No more
amalgams: Use of amalgam and amalgam alternative materials in primary dental care. Br Dent
J. 2018;225(2):171-6.
[7] Mutter J, Naumann J, Sadaghiani C, Walach H, Drasch G. Amalgam studies: disregarding
basic principles of mercury toxicity. Int J Hyg Environ Health. 2004;207(4):391-7.
[8] Zhou X, Huang X, Li M, Peng X, Wang S, Zhou X, et al. Development and status of resin
composite as dental restorative materials. J Appl Polym Sci. 2019;136(44):48180.
[9] Stahl F, Ashworth SH, Jandt KD, Mills RW. Light-emitting diode (LED) polymerisation
of dental composites: flexural properties and polymerisation potential. Biomaterials.
2000;21(13):1379-85.
[10] Peutzfeldt A. Resin composites in dentistry: the monomer systems. Eur J Oral Sci.
1997;105(2):97-116.
[11] Rodríguez HA, Kriven WM, Casanova H. Development of mechanical properties in dental
resin composite: Effect of filler size and filler aggregation state. Mater Sci Eng C.
2019;101:274-82.
[12] Yang D-L, Sun Q, Niu H, Wang R-L, Wang D, Wang J-X. The properties of dental resin
composites reinforced with silica colloidal nanoparticle clusters: Effects of heat treatment and
filler composition. Composites, Part B. 2020;186:107791.
[13] Cho K, Sul J-H, Stenzel MH, Farrar P, Prusty BG. Experimental cum computational
investigation on interfacial and mechanical behavior of short glass fiber reinforced dental
composites. Composites, Part B. 2020;200:108294.
[14] Aminoroaya A, Neisiany RE, Khorasani SN, Panahi P, Das O, Madry H, et al. A review
of dental composites: Challenges, chemistry aspects, filler influences, and future insights.
Composites, Part B. 2021:108852.
[15] Behl S, Farahani AD, Rajan G, Ellakwa A, Farrar P, Thordarson P, et al. Evaluation of
rheological behaviour of flowable dental composites reinforced with low aspect ratio micro-
sized glass fibres. Dent Mater. 2021;37(1):131-42.
[16] Raju R, Rajan G, Farrar P, Prusty BG. Dimensional stability of short fibre reinforced
flowable dental composites. Sci Rep. 2021;11(1):1-11.

48
[17] Sousa-Lima R, Silva L, Chaves L, Geraldeli S, Alonso R, Borges B. Extensive assessment
of the physical, mechanical, and adhesion behavior of a low-viscosity bulk fill composite and
a traditional resin composite in tooth cavities. Oper Dent. 2017;42(5):E159-E66.
[18] Buelvas DDA, Besegato JF, Vicentin BLS, Jussiani EI, Hoeppner MG, Andrello AC, et
al. Impact of light-cure protocols on the porosity and shrinkage of commercial bulk fill dental
resin composites with different flowability. J Polym Res. 2020;27(9):1-10.
[19] Kobayashi S, Nakajima M, Furusawa K, Tichy A, Hosaka K, Tagami J. Color adjustment
potential of single-shade resin composite to various-shade human teeth: Effect of structural
color phenomenon. Dent Mater J. 2021:2020-364.
[20] Salas M, Lucena C, Herrera LJ, Yebra A, Della Bona A, Pérez MM. Translucency
thresholds for dental materials. Dent Mater. 2018;34(8):1168-74.
[21] Cândea Ciurea A, Şurlin P, Stratul ŞI, Soancă A, Roman A, Moldovan M, et al. Evaluation
of the biocompatibility of resin composite‐based dental materials with gingival mesenchymal
stromal cells. Microsc Res Tech. 2019;82(10):1768-78.
[22] Shahi S, Özcan M, Maleki Dizaj S, Sharifi S, Al-Haj Husain N, Eftekhari A, et al. A
review on potential toxicity of dental material and screening their biocompatibility. Toxicol
Mech Methods. 2019;29(5):368-77.
[23] Çimen C, ÖZALP N. Biocompatibility evaluation of resin-based restorative materials: A
review. European Annals of Dental Sciences.48(1).
[24] Chen H, Wang R, Qian L, Liu H, Wang J, Zhu M. Surface modification of urchin-like
serried hydroxyapatite with sol-gel method and its application in dental composites.
Composites, Part B. 2020;182:107621.
[25] Wang Y, Hua H, Liu H, Zhu M, Zhu X-X. Surface modification of ZrO2 nanoparticles
and its effects on the properties of dental resin composites. ACS Appl Bio Mater.
2020;3(8):5300-9.
[26] Shah PK, Stansbury JW. Photopolymerization shrinkage-stress reduction in polymer-
based dental restoratives by surface modification of fillers. Dent Mater. 2021;37(4):578-87.
[27] Cho K, Wang G, Fang J, Rajan G, Stenzel MH, Farrar P, et al. Selective atomic-level
etching on short s-glass fibres to control interfacial properties for restorative dental composites.
Sci Rep. 2019;9(1):1-10.
[28] Rameshbabu AP, Mohanty S, Bankoti K, Ghosh P, Dhara S. Effect of alumina, silk and
ceria short fibers in reinforcement of Bis-GMA/TEGDMA dental resin. Composites, Part B.
2015;70:238-46.
[29] Sadr A, Bakhtiari B, Hayashi J, Luong MN, Chen Y-W, Chyz G, et al. Effects of fiber
reinforcement on adaptation and bond strength of a bulk-fill composite in deep preparations.
Dent Mater. 2020;36(4):527-34.
[30] Herrera-Franco P, Valadez-Gonzalez A. A study of the mechanical properties of short
natural-fiber reinforced composites. Composites, Part B. 2005;36(8):597-608.
[31] Cho K, Wang G, Raju R, Rajan G, Fang J, Stenzel MH, et al. Influence of surface treatment
on the interfacial and mechanical properties of short S-glass fiber-reinforced dental composites.
ACS Appl Mater Interfaces. 2019;11(35):32328-38.
[32] Sharma S, Kumar P, Chandra R, Setia P. Prediction of properties of silica
nanoparticle/hydroxyapatite fiber reinforced Bis-GMA/TEGDMA composites using molecular
dynamics. Comput Mater Sci. 2019;158:32-41.
49
[33] Chen G, Li A, Liu H, Huang S, Zhang Z, Liu W, et al. Mechanical and dynamic properties
of resin blend and composite systems: A molecular dynamics study. Compos Struct.
2018;190:160-8.
[34] Seo S, Lee DW, Ahn JS, Cunha K, Filippidi E, Ju SW, et al. Significant performance
enhancement of polymer resins by bioinspired dynamic bonding. Adv Mater.
2017;29(39):1703026.
[35] Khvostenko D, Salehi S, Naleway S, Hilton T, Ferracane J, Mitchell J, et al. Cyclic
mechanical loading promotes bacterial penetration along composite restoration marginal gaps.
Dent Mater. 2015;31(6):702-10.
[36] Maske TT, Kuper NK, Cenci MS, Huysmans M-CD. Minimal gap size and dentin wall
lesion development next to resin composite in a microcosm biofilm model. Caries Res.
2017;51(5):475-81.
[37] Ferracane JL. Resin composite-state of the art. Dent Mater. 2011;27(1):29-38.
[38] Marghalani HY. Resin-based dental composite materials. 13. 2016.
[39] Wang Y, Wang R, Habib E, Wang R, Zhang Q, Sun B, et al. Surface modification of
quartz fibres for dental composites through a sol-gel process. Mater Sci Eng C. 2017;74:21-6.
[40] Chen H, Wei S, Wang R, Zhu M. Improving the physical-mechanical property of dental
composites by grafting methacrylate-polyhedral oligomeric silsesquioxane onto a filler surface.
ACS Biomater Sci Eng. 2021;7(4):1428-37.
[41] Schnaider L, Ghosh M, Bychenko D, Grigoriants I, Ya’ari S, Shalev Antsel T, et al.
Enhanced nanoassembly-incorporated antibacterial composite materials. ACS Appl Mater
Interfaces. 2019;11(24):21334-42.
[42] Yang Y, Xu Z, Guo Y, Zhang H, Qiu Y, Li J, et al. Novel core–shell CHX/ACP
nanoparticles effectively improve the mechanical, antibacterial and remineralized properties of
the dental resin composite. Dent Mater. 2021;37(4):636-47.
[43] Wu Z, Xu H, Xie W, Wang M, Wang C, Gao C, et al. Study on a novel antibacterial light-
cured resin composite containing nano-MgO. Colloids Surf B Biointerfaces. 2020;188:110774.
[44] Bai X, Lin C, Wang Y, Ma J, Wang X, Yao X, et al. Preparation of Zn doped mesoporous
silica nanoparticles (Zn-MSNs) for the improvement of mechanical and antibacterial properties
of dental resin composites. Dent Mater. 2020;36(6):794-807.
[45] Makvandi P, Gu JT, Zare EN, Ashtari B, Moeini A, Tay FR, et al. Polymeric and inorganic
nanoscopical antimicrobial fillers in dentistry. Acta Biomater. 2020;101:69-101.
[46] Ali S, Sangi L, Kumar N, Kumar B, Khurshid Z, Zafar MS. Evaluating antibacterial and
surface mechanical properties of chitosan modified dental resin composites. Technol Health
Care. 2020;28(2):165-73.
[47] Jandt KD, Watts DC. Nanotechnology in dentistry: Present and future perspectives on
dental nanomaterials. Dent Mater. 2020.
[48] Vasudeva G. Monomer systems for dental composites and their future: a review. J Calif
Dent Assoc. 2009;37(6):389-98.
[49] Fugolin A, Pfeifer C. New resins for dental composites. J Dent Res. 2017;96(10):1085-
91.
[50] Xu X, He L, Zhu B, Li J, Li J. Advances in polymeric materials for dental applications.
Polym Chem. 2017;8(5):807-23.

50
[51] Bowen R. Properties of a silica-reinforced polymer for dental restorations. J Am Dent
Assoc. 1963;66(1):57-64.
[52] Dickens SH, Stansbury J, Choi K, Floyd C. Photopolymerization kinetics of methacrylate
dental resins. Macromolecules. 2003;36(16):6043-53.
[53] Moszner N, Fischer UK, Angermann J. New diluents for dental composites. Macromol.
Mater. Eng. 2016;301(6):750-9.
[54] He J, Kopperud HM. Preparation and characterization of Bis-GMA-free dental composites
with dimethacrylate monomer derived from 9, 9-Bis [4-(2-hydroxyethoxy) phenyl] fluorene.
Dent Mater. 2018;34(7):1003-13.
[55] Pérez-Mondragón AA, Cuevas-Suárez CE, González-López JA, Trejo-Carbajal N,
Meléndez-Rodríguez M, Herrera-González AM. Preparation and evaluation of a BisGMA-free
dental composite resin based on a novel trimethacrylate monomer. Dent Mater.
2020;36(4):542-50.
[56] Fugolin AP, de Paula AB, Dobson A, Huynh V, Consani R, Ferracane JL, et al. Alternative
monomer for BisGMA-free resin composites formulations. Dent Mater. 2020;36(7):884-92.
[57] Klauer E, Belli R, Petschelt A, Lohbauer U. Mechanical and hydrolytic degradation of an
Ormocer®-based Bis-GMA-free resin composite. Clin Oral Investig. 2019;23(5):2113-21.
[58] Kerby RE, Knobloch LA, Schricker S, Gregg B. Synthesis and evaluation of modified
urethane dimethacrylate resins with reduced water sorption and solubility. Dent Mater.
2009;25(3):302-13.
[59] Yoshinaga K, Yoshihara K, Yoshida Y. Development of new diacrylate monomers as
substitutes for Bis-GMA and UDMA. Dent Mater. 2021;37(6):e391-e8.
[60] Catel Y, Angermann J, Fässler P, Fischer U, Schnur T, Moszner N. High refractive index
monofunctional monomers as promising diluents for dental composites. Dent Mater.
2021;37(2):351-358.
[61] Pérez-Mondragón AA, Cuevas-Suárez CE, Castillo ORS, González-López JA, Herrera-
González AM. Evaluation of biocompatible monomers as substitutes for TEGDMA in resin-
based dental composites. Mater Sci Eng C. 2018;93: 80-87.
[62] Wang X, Huyang G, Palagummi SV, Liu X, Skrtic D, Beauchamp C, et al. High
performance dental resin composites with hydrolytically stable monomers. Dent Mater.
2018;34(2):228-237.
[63] He J, Garoushi S, Säilynoja E, Vallittu PK, Lassila L. The effect of adding a new monomer
“Phene” on the polymerization shrinkage reduction of a dental resin composite. Dent Mater.
2019;35(4):627-635.
[64] Cuevas-Suárez CE, González-López JA, da Silva AF, Piva E, Herrera-González AM.
Synthesis of an allyl carbonate monomer as alternative to TEGDMA in the formulation of
dental composite resins. J Mech Behav Biomed Mater. 2018;87:148-154.
[65] González-López JA, Pérez-Mondragón AA, Cuevas-Suárez CE, González SCE, Herrera-
González AM. Dental composite resins with low polymerization stress based on a new allyl
carbonate monomer. J Mech Behav Biomed Mater. 2020;110:103955.
[66] González‐López JA, Cuevas‐Suárez CE, Pérez‐Mondragón AA, Berlanga Duarte ML,
Herrera‐González AM. Photopolymerizable multifunctional monomers and their evaluation as
reactive Bis‐GMA eluents. J Appl Polym Sci. 2018;135(19):46240.

51
[67] Pratap B, Gupta RK, Bhardwaj B, Nag M. Modeling based experimental investigation on
polymerization shrinkage and micro-hardness of nano alumina filled resin based dental
material. J Mech Behav Biomed Mater. 2019;99:86-92.
[68] Amirouche-Korichi A, Mouzali M, Watts D. Shrinkage strain-rates study of dental
composites based on (BisGMA/TEGDMA) monomers. Arab J Chem. 2017;10:S190-S5.
[69] Srivastava R, Liu J, He C, Sun Y. BisGMA analogues as monomers and diluents for dental
restorative composite materials. Mater Sci Eng C. 2018;88:25-31.
[70] Gangwar S, Yadav S, Pathak VK. Optimized selection of nanohydroxyapatite‐filled dental
restorative composites formulation for best physico‐mechanical, chemical, and thermal
properties using hybrid analytical hierarchy process‐multi‐objective optimization on the basis
of ratio analysis approach. Polym Compos. 2021.
[71] Gupta R, Tomer AK, Kumari A, Mullick S, Dubey S. Bulkfill flowable composite resins-
A review. Int J Appl Dent Sci. 2017;3(2):38-40.
[72] Fronza BM, Rueggeberg FA, Braga RR, Mogilevych B, Soares LES, Martin AA, et al.
Monomer conversion, microhardness, internal marginal adaptation, and shrinkage stress of
bulk-fill resin composites. Dent Mater. 2015;31(12):1542-51.
[73] Ilie N, Hickel R. Investigations on mechanical behaviour of dental composites. Clin Oral
Investig. 2009;13(4):427-38.
[74] Mirică I-C, Furtos G, Bâldea B, Lucaciu O, Ilea A, Moldovan M, et al. Influence of filler
loading on the mechanical properties of flowable resin composites. Materials. 2020;13(6):1477.
[75] Behl S, Rajan G, Ellakwa A, Farrar P, Prusty BG. Physical and mechanical
characterisation of flowable dental composites reinforced with short aspect ratio micro-sized
S-Glass fibres. Mater Sci Eng C. 2020;111:110771.
[76] Pieniak D, Walczak A, Walczak M, Przystupa K, Niewczas AM. Hardness and wear
resistance of dental biomedical nanomaterials in a humid environment with non-stationary
temperatures. Materials. 2020;13(5):1255.
[77] Pratap B, Gupta RK, Denis L, Goswami D. Evaluation of polymerization shrinkage and
Vickers hardness for restorative dental composites. Mater Today: Proc. 2020;21:1563-5.
[78] Chadda H, Satapathy BK, Patnaik A, Ray AR. Mechanistic interpretations of fracture
toughness and correlations to wear behavior of hydroxyapatite and silica/hydroxyapatite filled
bis-GMA/TEGDMA micro/hybrid dental restorative composites. Composites, Part B.
2017;130:132-46.
[79] Kolb C, Gumpert K, Wolter H, Sextl G. Highly translucent dental resin composites
through refractive index adaption using zirconium dioxide nanoparticles and organic
functionalization. Dent Mater. 2020;36(10):1332-42.
[80] Faria-e-Silva AL, Dos Santos A, Tang A, Girotto EM, Pfeifer CS. Effect of thiourethane
filler surface functionalization on stress, conversion and mechanical properties of restorative
dental composites. Dent Mater. 2018;34(9):1351-8.
[81] Fu W, Wang L, He J. Evaluation of mechanical properties and shrinkage stress of thiol-
ene-methacrylate dental composites with synthesized fluorinated allyl ether. J Mech Behav
Biomed Mater. 2019;95:53-9.
[82] Yan Y, Chen C, Chen B, Shen J, Zhang H, Xie H. Effects of hydrothermal aging, thermal
cycling, and water storage on the mechanical properties of a machinable resin-based composite
containing nano-zirconia fillers. J Mech Behav Biomed Mater. 2020;102:103522.
52
[83] Boaro LCC, Campos LM, Varca GHC, Dos Santos TMR, Marques PA, Sugii MM, et al.
Antibacterial resin-based composite containing chlorhexidine for dental applications. Dent
Mater. 2019;35(6):909-18.
[84] Abaszadeh M, Mohammadi M, Mohammadzadeh I. Biocompatibility of a new
antibacterial compound and its effect on the mechanical properties of flowable dental
composites (animal study). J Dent. 2020;21(1):56.
[85] Yaghmoor RB, Xia W, Ashley P, Allan E, Young AM. Effect of novel antibacterial
composites on bacterial biofilms. J Funct Biomater. 2020;11(3):55.
[86] Yang D-L, Cui Y-N, Sun Q, Liu M, Niu H, Wang J-X. Antibacterial activity and
reinforcing effect of SiO 2–ZnO complex cluster fillers for dental resin composites. Biomater
Sci. 2021;9(5):1795-804.
[87] Wang Y, Zhu M, Zhu X. Functional fillers for dental resin composites. Acta Biomater.
2020.
[88] Habib E, Wang R, Zhu X. Correlation of resin viscosity and monomer conversion to filler
particle size in dental composites. Dent Mater. 2018;34(10):1501-8.
[89] Wang R, Habib E, Zhu X. Evaluation of the filler packing structures in dental resin
composites: from theory to practice. Dent Mater. 2018;34(7):1014-23.
[90] Alrahlah A, Khan R, Al-Odayni A-B, Saeed WS, Bautista LS, Vohra F. Evaluation of
synergic potential of rGO/SiO2 as hybrid filler for BisGMA/TEGDMA dental composites.
Polymers. 2020;12(12):3025.
[91] Sun C, Xu D, Hou C, Zhang H, Li Y, Zhang Q, et al. Core-shell structured SiO2@ ZrO2@
SiO2 filler for radiopacity and ultra-low shrinkage dental composite resins. J Mech Behav
Biomed Mater. 2021:104593.
[92] Zhang Y, Huang C, Chang J. Ca-Doped mesoporous SiO2/dental resin composites with
enhanced mechanical properties, bioactivity and antibacterial properties. J Mater Chem B.
2018;6(3):477-86.
[93] Siddiqui U, Khalid H, Ghafoor S, Javaid A, Asif A, Khan AS. Analyses on mechanical
and physical performances of nano-apatite grafted glass fibers based dental composites. Mater
Chem Phys. 2021;263:124188.
[94] Odermatt R, Mohn D, Wiedemeier DB, Attin T, Tauböck TT. Bioactivity and physico-
chemical properties of dental composites functionalized with nano-vs. micro-sized bioactive
glass. J Clin Med. 2020;9(3):772.
[95] Li Q, Tang C, Liu F, He J. The physiochemical properties of dental resin composites
reinforced with milled E-glass fibers. Silicon. 2018;10(5):1999-2007.
[96] Aguiar AE, da Silva LG, de Paula Barbosa HF, Gloria RF, Espanhol-Soares M, Gimenes
R. Synthesis of Al2O3-0.5 B2O3-SiO2 fillers by sol-gel method for dental resin composites. J
Non-Cryst Solids. 2017;458:86-96.
[97] Yang J, Shen J, Wu X, He F, Xie H, Chen C. Effects of nano-zirconia fillers conditioned
with phosphate ester monomers on the conversion and mechanical properties of Bis-GMA-and
UDMA-based resin composites. J Dent. 2020;94:103306.
[98] Barot T, Rawtani D, Kulkarni P. Physicochemical and biological assessment of silver
nanoparticles immobilized Halloysite nanotubes-based resin composite for dental applications.
Heliyon. 2020;6(3):e03601.

53
[99] de Freitas Guimaraes GM, Bronze-Uhle ES, Lisboa-Filho PN, Fugolin APP, Borges AFS,
Gonzaga CC, et al. Effect of the addition of functionalized TiO2 nanotubes and nanoparticles
on properties of experimental resin composites. Dent Mater. 2020;36(12):1544-56.
[100] Elgamily HM, El-Sayed HS, Abdelnabi A. The antibacterial effect of two cavity
disinfectants against one of cariogenic pathogen: An In vitro comparative study. Contemp Clin
Dent. 2018;9(3):457.
[101] Bukovinszky K, Szalóki M, Csarnovics I, Bonyár A, Petrik P, Kalas B, et al.
Optimization of plasmonic gold nanoparticle concentration in green LED light active dental
photopolymer. Polymers. 2021;13(2):275.
[102] Cho K, Yasir M, Jung M, Willcox MD, Stenzel MH, Rajan G, et al. Hybrid engineered
dental composites by multiscale reinforcements with chitosan-integrated halloysite nanotubes
and S-glass fibers. Composites, Part B. 2020;202:108448.
[103] Zhao S-N, Yang D-L, Wang D, Pu Y, Le Y, Wang J-X, et al. Design and efficient
fabrication of micro-sized clusters of hydroxyapatite nanorods for dental resin composites. J
Mater Sci. 2019;54(5):3878-92.
[104] Zhou W, Peng X, Zhou X, Weir MD, Melo MAS, Tay FR, et al. In vitro evaluation of
composite containing DMAHDM and calcium phosphate nanoparticles on recurrent caries
inhibition at bovine enamel-restoration margins. Dent Mater. 2020;36(10):1343-55.
[105] Al-Dulaijan YA, Cheng L, Weir MD, Melo MAS, Liu H, Oates TW, et al. Novel
rechargeable calcium phosphate nanocomposite with antibacterial activity to suppress biofilm
acids and dental caries. J Dent. 2018;72:44-52.
[106] Alania Y, Natale LC, Nesadal D, Vilela H, Magalhães AC, Braga RR. In vitro
remineralization of artificial enamel caries with resin composites containing calcium phosphate
particles. J Biomed Mater Res Part B Appl Biomater. 2019;107(5):1542-50.
[107] Buchwald Z, Sandomierski M, Voelkel A. Calcium-rich 13X zeolite as a filler with
remineralizing potential for dental composites. ACS Biomater Sci Eng. 2020;6(7):3843-54.
[108] Okulus Z, Sandomierski M, Zielińska M, Buchwald T, Voelkel A. Zeolite fillers for
resin-based composites with remineralizing potential. Spectrochim Acta A Mol Biomol
Spectrosc. 2019;210:126-35.
[109] Fanfoni L, De Biasi M, Antollovich G, Di Lenarda R, Angerame D. Evaluation of degree
of conversion, rate of cure, microhardness, depth of cure, and contraction stress of new
nanohybrid composites containing pre-polymerized spherical filler. J Mater Sci Mater Med.
2020;31(12):1-11.
[110] Cevik P, Yildirim‐Bicer AZ. The effect of silica and prepolymer nanoparticles on the
mechanical properties of denture base acrylic resin. J Prosthodont. 2018;27(8):763-70.
[111] Sabir M, Ali A, Siddiqui U, Muhammad N, Khan AS, Sharif F, et al. Synthesis and
characterization of cellulose/hydroxyapatite based dental restorative composites. J Biomater
Sci Polym Ed. 2020;31(14):1806-19.
[112] Wang Y, Hua H, Li W, Wang R, Jiang X, Zhu M. Strong antibacterial dental resin
composites containing cellulose nanocrystal/zinc oxide nanohybrids. J Dent. 2019;80:23-9.
[113] Djustiana N, Hasratiningsih Z, Karlina E, Febrida R, Takarini V, Cahyanto A, et al.
Hardness evaluation of dental composite with ceramic fillers. Key Eng Mater. 2016. p. 74-9.
[114] Wang L, Liu Y, Si W, Feng H, Tao Y, Ma Z. Friction and wear behaviors of dental
ceramics against natural tooth enamel. J Eur Ceram Soc. 2012;32(11):2599-606.
54
[115] Cao W, Zhang Y, Wang X, Chen Y, Li Q, Xing X, et al. Development of a novel resin-
based dental material with dual biocidal modes and sustained release of Ag+ ions based on
photocurable core-shell AgBr/cationic polymer nanocomposites. J Mater Sci Mater Med.
2017;28(7):103.
[116] Dias HB, Bernardi MIB, Bauab TM, Hernandes AC, de Souza Rastelli AN. Titanium
dioxide and modified titanium dioxide by silver nanoparticles as an anti biofilm filler content
for composite resins. Dent Mater. 2019;35(2):e36-e46.
[117] Sarosi C, Biris AR, Antoniac A, Boboia S, Alb C, Antoniac I, et al. The nanofiller effect
on properties of experimental graphene dental nanocomposites. J Adhes Sci Technol.
2016;30(16):1779-94.
[118] Zafar MS, Alnazzawi AA, Alrahabi M, Fareed MA, Najeeb S, Khurshid Z.
Nanotechnology and nanomaterials in dentistry. Adv Dent Biomater. 2019. p. 477-505.
[119] Teughels W, Van Assche N, Sliepen I, Quirynen M. Effect of material characteristics
and/or surface topography on biofilm development. Clin Oral Implants Res. 2006;17(S2):68-
81.
[120] Wu YR, Chang CW, Chang KC, Lin DJ, Ko CL, Wu HY, et al. Effect of micro‐/nano‐
hybrid hydroxyapatite rod reinforcement in composite resins on strength through thermal
cycling. Polym Compos. 2019;40(9):3703-10.
[121] Huang Q, Qin W, Garoushi S, He J, Lin Z, Liu F, et al. Physicochemical properties of
discontinuous S2-glass fiber reinforced resin composite. Dent Mater J. 2017.
[122] Huang Q, Garoushi S, Lin Z, He J, Qin W, Liu F, et al. Properties of discontinuous S2-
glass fiber-particulate-reinforced resin composites with two different fiber length distributions.
J Prosthodont Res. 2017;61(4):471-9.
[123] Lassila L, Keulemans F, Säilynoja E, Vallittu PK, Garoushi S. Mechanical properties
and fracture behavior of flowable fiber reinforced composite restorations. Dent Mater.
2018;34(4):598-606.
[124] Jardim RN, Rocha AA, Rossi AM, de Almeida Neves A, Portela MB, Lopes RT, et al.
Fabrication and characterization of remineralizing dental composites containing
hydroxyapatite nanoparticles. J Mech Behav Biomed Mater. 2020;109:103817.
[125] Du M, Chen J, Liu K, Xing H, Song C. Recent advances in biomedical engineering of
nano-hydroxyapatite including dentistry, cancer treatment and bone repair. Composites, Part
B. 2021:108790.
[126] Reis DP, Noronha Filho JD, Rossi AL, de Almeida Neves A, Portela MB, da Silva EM.
Remineralizing potential of dental composites containing silanized silica-hydroxyapatite (Si-
HAp) nanoporous particles charged with sodium fluoride (NaF). J Dent. 2019;90:103211.
[127] Karczewski A, Kalagi S, Viana ÍEL, Martins VM, Duarte S, Gregory RL, et al. Resin-
based dental materials containing 3-aminopropyltriethoxysilane modified halloysite-clay
nanotubes for extended drug delivery. Dent Mater. 2021;37(3):508-15.
[128] Barot T, Rawtani D, Kulkarni P, Hussain CM, Akkireddy S. Physicochemical and
biological assessment of flowable resin composites incorporated with farnesol loaded
halloysite nanotubes for dental applications. J Mech Behav Biomed Mater. 2020;104:103675.
[129] Cunha DA, Rodrigues NS, Souza LC, Lomonaco D, Rodrigues FP, Degrazia FW, et al.
Physicochemical and microbiological assessment of an experimental composite doped with
triclosan-loaded halloysite nanotubes. Materials. 2018;11(7):1080.

55
[130] Cha J, Kim J, Ryu S, Hong SH. Strengthening effect of melamine functionalized low-
dimension carbon at fiber reinforced polymer composites and their interlaminar shear behavior.
Composites, Part B. 2019;173:106976.
[131] Chen H, Wang R, Qian L, Ren Q, Jiang X, Zhu M. Dental restorative resin composites:
modification technologies for the matrix/filler interface. Macromol Mater.Eng.
2018;303(10):1800264.
[132] Dai S, Chen Y, Yang J, He F, Chen C, Xie H. Surface treatment of nanozirconia fillers
to strengthen dental bisphenol A-glycidyl methacrylate–based resin composites. Int J
Nanomedicine. 2019;14:9185.
[133] Zeng S, Zhang T, Nie M, Fei G, Wang Q. Effect of root-like mechanical-interlocking
interface in polypropylene/aramid fiber composites from experimental to numerical study.
Composites, Part B. 2021;216:108868.
[134] Sahin M, Schlögl S, Kalinka G, Wang J, Kaynak B, Mühlbacher I, et al. Tailoring the
interfaces in glass fiber-reinforced photopolymer composites. Polymer. 2018;141:221-31.
[135] Shouha PS, Ellakwa AE. Effect of short glass fibers on the polymerization shrinkage
stress of dental composite. J Biomed Mater Res Part B Appl Biomater. 2017;105(7):1930-7.
[136] Liu M, Chen J, Hu X, Du Y, Xia Y, Gu N, et al. Enhanced properties of fiberglass-
reinforced photocurable resin pile by introducing different fiberglass surface treatments and
their biological evolution. RSC Adv. 2015;5(85):69690-7.
[137] Jang H, Chung Y, Whangbo S, Kim T, Whang C, Lee S, et al. Effects of chemical etching
with nitric acid on glass surfaces. J Vac Sci Technol A: Vac Surf Films. 2001;19(1):267-74.
[138] Amiri P, Talebi Z, Semnani D, Bagheri R, Fashandi H. Improved performance of Bis-
GMA dental composites reinforced with surface-modified PAN nanofibers. J Mater Sci: Mater.
Med. 2021;32(7):1-8.
[139] Bukhari JH, Khan AS, Ijaz K, Zahid S, Chaudhry AA, Kaleem M. Low-temperature
flow-synthesis-assisted urethane-grafted zinc oxide-based dental composites: physical,
mechanical, and antibacterial responses. J Mater Sci: Mater. Med. 2021; 32(8):1-11.
[140] Matinlinna JP, Lung CYK, Tsoi JKH. Silane adhesion mechanism in dental applications
and surface treatments: A review. Dent Mater. 2018;34(1):13-28.
[141] Aydınoğlu A, Yoruç ABH. Effects of silane-modified fillers on properties of dental
composite resin. Mater Sci Eng C. 2017;79:382-9.
[142] Yadav S, Gangwar S. The effectiveness of functionalized nano-hydroxyapatite filler on
the physical and mechanical properties of novel dental restorative composite. Int J Polym
Mater Polym Biomater. 2019.
[143] Fugolin APP, Costa AR, Correr-Sobrinho L, Chaw RC, Lewis S, Ferracane JL, Pfeifer
CS. Toughening and polymerization stress control in composites using thiourethane-treated
fillers. Sci Rep. 2021:11(1):1-12.
[144] Lewis S, Fugolin A, Lam S, Scanlon C, Ferracane J, Pfeifer C. Effects of systematically
varied thiourethane-functionalized filler concentration on polymerization behavior and
relevant clinical properties of dental composites. Mater Des. 2021;197:109249.
[145] Bittmann B, Haupert F, Schlarb AK. Ultrasonic dispersion of inorganic nanoparticles in
epoxy resin. Ultrason Sonochemistry. 2009;16(5):622-8.

56
[146] Yang D-L, Sun Q, Duan Y-H, Niu H, Wang R-L, Wang D, et al. Efficient construction
of SiO2 colloidal nanoparticle clusters as novel fillers by a spray-drying process for dental
composites. Ind Eng Chem Res. 2019;58(39):18178-86.
[147] Scaffaro R, Maio A. Enhancing the mechanical performance of polymer based
nanocomposites by plasma-modification of nanoparticles. Polym Test. 2012;31(7):889-94.
[148] de Menezes BRC, da Graça Sampaio A, da Silva DM, do Amaral Montanheiro TL, Koga-
Ito CY, Thim GP. AgVO3 nanorods silanized with γ-MPS: An alternative for effective
dispersion of AgVO3 in dental acrylic resins improving the mechanical properties. Appl Surf
Sci. 2021;543:148830.
[149] Felix M, Martinez I, Romero A, Partal P, Guerrero A. Effect of pH and nanoclay content
on the morphology and physicochemical properties of soy protein/montmorillonite
nanocomposite obtained by extrusion. Composites, Part B. 2018;140:197-203.
[150] Goldstein GR. The longevity of direct and indirect posterior restorations is uncertain and
may be affected by a number of dentist-, patient-, and material-related factors. J Evid Based
Dent Pract. 2010;10(1):30-31.
[151] Ercalik‐Yalcinkaya S, Özcan M. Association between oral mucosal lesions and hygiene
habits in a population of removable prosthesis wearers. J Prosthodont. 2015;24(4):271-278.
[152] Turkistani KA, Alkayyal MA, Abbassy MA, Al-Dharrab AA, Zahran MH, Melis M, et
al. Comparison of occlusal bite force distribution in subjects with different occlusal
characteristics. CRANIO®. 2020:1-8.
[153] Ge J, Cui F, Wang X, Feng H. Property variations in the prism and the organic sheath
within enamel by nanoindentation. Biomaterials. 2005;26(16):3333-9.
[154] Rees J, Jacobsen P. The elastic moduli of enamel and dentine. Clin Mater. 1993;14(1):35-
9.
[155] Marshall Jr G, Balooch M, Gallagher R, Gansky S, Marshall S. Mechanical properties of
the dentinoenamel junction: AFM studies of nanohardness, elastic modulus, and fracture. J
Biomed Mater Res Part B Appl Biomater. 2001;54(1):87-95.
[156] Abe Y, Lambrechts P, Inoue S, Braem M, Takeuchi M, Vanherle G, et al. Dynamic
elastic modulus of ‘packable’composites. Dent Mater. 2001;17(6):520-5.
[157] Shouha P, Swain M, Ellakwa A. The effect of fiber aspect ratio and volume loading on
the flexural properties of flowable dental composite. Dent Mater. 2014;30(11):1234-44.
[158] Oja J, Lassila L, Vallittu PK, Garoushi S. Effect of accelerated aging on some mechanical
properties and wear of different commercial dental resin composites. Materials.
2021;14(11):2769.
[159] Lassila L, Säilynoja E, Prinssi R, Vallittu P, Garoushi S. Characterization of a new fiber-
reinforced flowable composite. Odontology. 2019;107(3):342-52.
[160] Imai A, Takamizawa T, Sugimura R, Tsujimoto A, Ishii R, Kawazu M, et al. Interrelation
among the handling, mechanical, and wear properties of the newly developed flowable resin
composites. J Mech Behav Biomed Mater. 2019;89:72-80.
[161] Gamstedt E, Sjögren B. Micromechanisms in tension-compression fatigue of composite
laminates containing transverse plies. Compos Sci Technol. 1999;59(2):167-78.
[162] International Standards Organization (1992) ISO 4049: 1988/Cor.1: 1992, Dentistry –
resin-based filling materials, pp. 6–8. International Standards Organization, Geneva.

57
[163] Manhart J, Kunzelmann K-H, Chen H, Hickel R. Mechanical properties and wear
behavior of light-cured packable composite resins. Dent Mater. 2000;16(1):33-40.
[164] Chen Q, Zhao Y, Wu W, Xu T, Fong H. Fabrication and evaluation of Bis-
GMA/TEGDMA dental resins/composites containing halloysite nanotubes. Dent Mater.
2012;28(10):1071-9.
[165] Ilie N, Hilton T, Heintze S, Hickel R, Watts D, Silikas N, et al. Academy of dental
materials guidance—resin composites: Part I-mechanical properties. Dent Mater.
2017;33(8):880-94.
[166] Aminoroaya A, Esmaeely Neisiany R, Nouri Khorasani S, Panahi P, Das O, Ramakrishna
S. A review of dental composites: methods of characterizations. ACS Biomater Sci Eng.
2020;6(7):3713-44.
[167] Brunthaler A, König F, Lucas T, Sperr W, Schedle A. Longevity of direct resin composite
restorations in posterior teeth: a review. Clin Oral Investig. 2003;7(2):63-70.
[168] Van Dijken J. Direct resin composite inlays/onlays: an 11 year follow-up. J Dent.
2000;28(5):299-306.
[169] da Rosa Rodolpho PA, Cenci MS, Donassollo TA, Loguércio AD, Demarco FF. A
clinical evaluation of posterior composite restorations: 17-year findings. J Dent.
2006;34(7):427-35.
[170] Ástvaldsdóttir Á, Dagerhamn J, van Dijken JW, Naimi-Akbar A, Sandborgh-Englund G,
Tranæus S, et al. Longevity of posterior resin composite restorations in adults-A systematic
review. J Dent. 2015;43(8):934-54.
[171] Alshabib A, Silikas N, Watts DC. Hardness and fracture toughness of resin-composite
materials with and without fibers. Dent Mater. 2019;35(8):1194-203.
[172] Fujishima A, Ferracane JL. Comparison of four modes of fracture toughness testing for
dental composites. Dent Mater. 1996;12(1):38-43.
[173] Soderholm K-J. Review of the fracture toughness approach. Dent Mater. 2010;26(2):e63-
e77.
[174] Ornaghi BP, Meier MM, Lohbauer U, Braga RR. Fracture toughness and cyclic fatigue
resistance of resin composites with different filler size distributions. Dent Mater.
2014;30(7):742-51.
[175] Thomaidis S, Kakaboura A, Mueller WD, Zinelis S. Mechanical properties of
contemporary composite resins and their interrelations. Dent Mater. 2013;29(8):e132-e41.
[176] El Mowafy O, Watts D. Fracture toughness of human dentin. J Dent Res.
1986;65(5):677-81.
[177] Ilie N, Hickel R, Valceanu AS, Huth KC. Fracture toughness of dental restorative
materials. Clin. Oral Investig. 2012;16(2):489-98.
[178] Lassila L, Keulemans F, Vallittu PK, Garoushi S. Characterization of restorative short-
fiber reinforced dental composites. Dent Mater J. 2020;39(6):992-9.
[179] Garoushi S, Gargoum A, Vallittu PK, Lassila L. Short fiber‐reinforced composite
restorations: a review of the current literature. J Investig Clin Dent. 2018;9(3):e12330.
[180] Soares LM, Razaghy M, Magne P. Optimization of large MOD restorations: Composite
resin inlays vs. short fiber-reinforced direct restorations. Dent Mater. 2018;34(4):587-97.

58
[181] Tiu J, Belli R, Lohbauer U. R-curve behavior of a short-fiber reinforced resin composite
after water storage. J Mech Behav Biomed Mater. 2020;104:103674.
[182] Garoushi S, Sungur S, Boz Y, Ozkan P, Vallittu P, Uctasli S, et al. Influence of short-
fiber composite base on fracture behavior of direct and indirect restorations. Clin. Oral Investig.
2021:1-10.
[183] Fráter M, Sáry T, Vincze-Bandi E, Volom A, Braunitzer G, Szabó P B, et al. Fracture
behavior of short fiber-reinforced direct restorations in large MOD cavities. Polymers.
2021;13(13):2040.
[184] Watts D, Marouf A, Al-Hindi A. Photo-polymerization shrinkage-stress kinetics in resin-
composites: methods development. Dent Mater. 2003;19(1):1-11.
[185] Maas MS, Alania Y, Natale LC, Rodrigues MC, Watts DC, Braga RR. Trends in
restorative composites research: what is in the future? Braz Oral Res. 2017;31.
[186] Schneider LFJ, Cavalcante LM, Silikas N. Shrinkage stresses generated during resin-
composite applications: a review. J Dent Biomech. 2010;2010.
[187] Soares CJ, Rodrigues MdP, Vilela ABF, Pfeifer CS, Tantbirojn D, Versluis A.
Polymerization shrinkage stress of composite resins and resin cements–What do we need to
know? Braz Oral Res. 2017;31.
[188] Al Sunbul H, Silikas N, Watts DC. Polymerization shrinkage kinetics and shrinkage-
stress in dental resin-composites. Dent Mater. 2016;32(8):998-1006.
[189] Opdam N, Van De Sande F, Bronkhorst E, Cenci M, Bottenberg P, Pallesen U, et al.
Longevity of posterior composite restorations: a systematic review and meta-analysis. J Dent
Res. 2014;93(10):943-9.
[190] Guimaraes GF, Marcelino E, Cesarino I, Vicente FB, Grandini CR, Simoes RP.
Minimization of polymerization shrinkage effects on composite resins by the control of
irradiance during the photoactivation process. J Appl Oral Sci. 2018;26.
[191] He J, Garoushi S, Vallittu PK, Lassila L. Effect of low-shrinkage monomers on the
physicochemical properties of experimental composite resin. Acta Biomater Odontol Scand.
2018;4(1):30-7.
[192] Kwon Y, Ferracane J, Lee I-B. Effect of layering methods, composite type, and flowable
liner on the polymerization shrinkage stress of light cured composites. Dent Mater.
2012;28(7):801-9.
[193] Ferracane J, Hilton T, Stansbury J, Watts D, Silikas N, Ilie N, et al. Academy of dental
materials guidance-resin composites: Part II-technique sensitivity (handling, polymerization,
dimensional changes). Dent Mater. 2017;33(11):1171-91.
[194] Sakaguchi R, Sasik C, Bunczak M, Douglas WH. Strain gauge method for measuring
polymerization contraction of composite restoratives. J Dent. 1991;19(5):312-6.
[195] Rajan G, Raju R, Jinachandran S, Farrar P, Xi J, Prusty BG. Polymerisation shrinkage
profiling of dental composites using optical fibre sensing and their correlation with degree of
conversion and curing rate. Sci Rep. 2019;9(1):1-10.
[196] Behl S, Rajan G, Farrar P, Prentice L, Prusty BG. Evaluation of depth-wise post-gel
polymerisation shrinkage behaviour of flowable dental composites. J Mech Behav Biomed
Mater. 2021;104860.

59
[197] Lau A, Li J, Heo YC, Fok A. A study of polymerization shrinkage kinetics using digital
image correlation. Dent Mater. 2015;31(4):391-8.
[198] Rajan G, Wong A, Farrar P, Prusty GB. Post-gel polymerisation shrinkage profiling of
polymer biomaterials using a chirped fibre Bragg grating. Sci Rep. 2021;11(1):1-10.
[199] Amaireh A, Al-Jundi S, Alshraideh H. In vitro evaluation of microleakage in primary
teeth restored with three adhesive materials: ACTIVA™, composite resin, and resin-modified
glass ionomer. Eur Arch Paediatr Dent. 2019;20(4):359-67.
[200] Zhou Y, Hiraishi N, Shimada Y, Wang G, Tagami J, Feng X. Evaluation of tooth
demineralization and interfacial bacterial penetration around resin composites containing
surface pre-reacted glass-ionomer (S-PRG) filler. Dent Mater. 2021;37(5):849-62.
[201] Harper RA, Shelton RM, James JD, Salvati E, Besnard C, Korsunsky AM, et al. Acid-
induced demineralisation of human enamel as a function of time and pH observed using X-ray
and polarised light imaging. Acta Biomater. 2021;120:240-8.
[202] Robinson C, Shore RC, Brookes SJ, Strafford S, Wood S, Kirkham J. The chemistry of
enamel caries. Crit Rev Oral Biol Med. 2000;11(4):481-95.
[203] Kaur R, Liu S. Antibacterial surface design–Contact kill. Prog Surf Sci. 2016;91(3):136-
53.
[204] Mitwalli H, Alsahafi R, Balhaddad AA, Weir MD, Xu HH, Melo MAS. Emerging
contact-killing antibacterial strategies for developing anti-biofilm dental polymeric restorative
materials. Bioengineering. 2020;7(3):83.
[205] Ai M, Du Z, Zhu S, Geng H, Zhang X, Cai Q, et al. Composite resin reinforced with
silver nanoparticles-laden hydroxyapatite nanowires for dental application. Dent Mater.
2017;33(1):12-22.
[206] Zajdowicz S, Song HB, Baranek A, Bowman CN. Evaluation of biofilm formation on
novel copper-catalyzed azide-alkyne cycloaddition (CuAAC)-based resins for dental
restoratives. Dent Mater. 2018;34(4):657-66.
[207] Sodagar A, Akhoundi MSA, Bahador A, Jalali YF, Behzadi Z, Elhaminejad F, et al.
Effect of TiO2 nanoparticles incorporation on antibacterial properties and shear bond strength
of dental composite used in Orthodontics. Dental Press J Orthod. 2017;22:67-74.
[208] Chen H, Wang R, Zhang J, Hua H, Zhu M. Synthesis of core-shell structured ZnO@ m-
SiO2 with excellent reinforcing effect and antimicrobial activity for dental resin composites.
Dent Mater. 2018;34(12):1846-55.
[209] Dias HB, Bernardi MIB, Marangoni VS, de Abreu Bernardi AC, de Souza Rastelli AN,
Hernandes AC. Synthesis, characterization and application of Ag doped ZnO nanoparticles in
a composite resin. Mater Sci Eng C. 2019;96:391-401.
[210] Fei X, Li Y, Weir MD, Baras BH, Wang H, Wang S, et al. Novel pit and fissure sealant
containing nano-CaF2 and dimethylaminohexadecyl methacrylate with double benefits of
fluoride release and antibacterial function. Dent Mater. 2020;36(9):1241-53.
[211] Barot T, Rawtani D, Kulkarni P. Development of chlorhexidine loaded halloysite
nanotube based experimental resin composite with enhanced physico-mechanical and
biological properties for dental applications. J Compos Sci. 2020;4(2):81.
[212] Luo D, Shahid S, Hasan SM, Whiley R, Sukhorukov GB, Cattell MJ. Controlled release
of chlorhexidine from a HEMA-UDMA resin using a magnetic field. Dent Mater.
2018;34(5):764-75.
60
[213] Zane A, Zuo R, Villamena FA, Rockenbauer A, Foushee AMD, Flores K, et al.
Biocompatibility and antibacterial activity of nitrogen-doped titanium dioxide nanoparticles
for use in dental resin formulations. Int J Nanomedicine. 2016;11:6459.
[214] Tanaka CB, Lopes DP, Kikuchi LN, Moreira MS, Catalani LH, Braga RR, et al.
Development of novel dental restorative composites with dibasic calcium phosphate loaded
chitosan fillers. Dent Mater. 2020;36(4):551-9.
[215] Ceci M, Viola M, Rattalino D, Beltrami R, Colombo M, Poggio C. Discoloration of
different esthetic restorative materials: A spectrophotometric evaluation. Eur J Dent.
2017;11(02):149-56.
[216] Stenhagen IS, Rukke HV, Dragland IS, Kopperud HM. Effect of methacrylated chitosan
incorporated in experimental composite and adhesive on mechanical properties and biofilm
formation. Eur J Oral Sci. 2019;127(1):81-8.
[217] de Souza Araujo IJ, de Paula AB, Bruschi Alonso RC, Taparelli JR, Innocentini Mei LH,
Stipp RN, et al. A novel Triclosan Methacrylate-based composite reduces the virulence of
Streptococcus mutans biofilm. PLoS One. 2018;13(4):e0195244.
[218] Sun Q, Zhang L, Bai R, Zhuang Z, Zhang Y, Yu T, et al. Recent progress in antimicrobial
strategies for resin-based restoratives. Polymers. 2021;13(10):1590.
[219] Monteiro JC, Garcia IM, Leitune VCB, Visioli F, de Souza Balbinot G, Samuel SMW,
et al. Halloysite nanotubes loaded with alkyl trimethyl ammonium bromide as antibacterial
agent for root canal sealers. Dent Mater. 2019;35(5):789-96.
[220] Cherchali FZ, Attik N, Mouzali M, Tommasino JB, Abouelleil H, Decoret D, et al.
Structural stability of DHMAI antibacterial dental composite following in vitro biological
aging. Dent Mater. 2020;36(9):1161-9.
[221] Wang W, Wu F, Zhang G, Zhu S, Ban J, Wang L. Preparation of a highly crosslinked
biosafe dental nanocomposite resin with a tetrafunctional methacrylate quaternary ammonium
salt monomer. RSC Adv. 2019;9(71):41616-27.
[222] Zhang L, Ma Z, Wang R, Zhu M. Synthesis and Characterization of Methacrylate-
Functionalized Betulin Derivatives as Antibacterial Comonomer for Dental Restorative Resins.
ACS Biomater Sci Eng. 2021;7 (7):3132-3140.
[223] Koyama J, Fukazawa K, Ishihara K, Mori Y. In situ surface modification on dental
composite resin using 2-methacryloyloxyethyl phosphorylcholine polymer for controlling
plaque formation. Mater Sci Eng C. 2019;104:109916.
[224] Lee M-J, Kwon J-S, Kim J-Y, Ryu J-H, Seo J-Y, Jang S, et al. Bioactive resin-based
composite with surface pre-reacted glass-ionomer filler and zwitterionic material to prevent the
formation of multi-species biofilm. Dent Mater. 2019;35(9):1331-41.
[225] Zhang N, Chen C, Melo MA, Bai Y-X, Cheng L, Xu HH. A novel protein-repellent
dental composite containing 2-methacryloyloxyethyl phosphorylcholine. Int J Oral Sci.
2015;7(2):103-9.
[226] Van Houte J. Role of micro-organisms in caries etiology. J Dent Res. 1994;73(3):672-
81.
[227] Gross EL, Beall CJ, Kutsch SR, Firestone ND, Leys EJ, Griffen AL. Beyond
Streptococcus mutans: dental caries onset linked to multiple species by 16S rRNA community
analysis. PLoS One. 2012.

61
[228] Karpiński TM, Szkaradkiewicz AK. Microbiology of dental caries. J Biol Earth Sci.
2013;3(1):M21-4.
[229] Hajishengallis G. Periodontitis: from microbial immune subversion to systemic
inflammation. Nat Rev Immunol. 2015;15(1):30-44.
[230] Bourgeois D, Inquimbert C, Ottolenghi L, Carrouel F. Periodontal pathogens as risk
factors of cardiovascular diseases, diabetes, rheumatoid arthritis, cancer, and chronic
obstructive pulmonary disease-is there cause for consideration? Microorganisms.
2019;7(10):424.
[231] Nakamura K, Oaki Y, Imai H. Multistep crystal growth of oriented fluorapatite nanorod
arrays for fabrication of enamel-like architectures on a polymer sheet. CrystEngComm.
2017;19(4):669-74.
[232] Shao C, Jin B, Mu Z, Lu H, Zhao Y, Wu Z, et al. Repair of tooth enamel by a biomimetic
mineralization frontier ensuring epitaxial growth. Sci Adv. 2019;5(8):eaaw9569.
[233] Hibino Y, Oyane A, Shitomi K, Miyaji H. Technique for simple apatite coating on a
dental resin composite with light-curing through a micro-rough apatite layer. Mater Sci Eng C.
2020;116:111146.
[234] Weir MD, Ruan J, Zhang N, Chow LC, Zhang K, Chang X, et al. Effect of calcium
phosphate nanocomposite on in vitro remineralization of human dentin lesions. Dent Mater.
2017;33(9):1033-44.
[235] Liang K, Wang S, Tao S, Xiao S, Zhou H, Wang P, et al. Dental remineralization via
poly (amido amine) and restorative materials containing calcium phosphate nanoparticles. Int
J Oral Sci. 2019;11(2):1-12.
[236] Langhorst S, O’donnell J, Skrtic D. In vitro remineralization of enamel by polymeric
amorphous calcium phosphate composite: quantitative microradiographic study. Dent Mater.
2009;25(7):884-91.
[237] Dong Z, Ni Y, Yang X, Hu C, Sun J, Li L, et al. Characterization and analysis of fluoride
calcium silicate composite interface in remineralization of dental enamel. Composites, Part B.
2018;153:393-7.
[238] Rabee M, Nomaan K, Abdelhady A. Effect of remineralizing agents on bond strength of
resin composite to dentin: An in-vitro study. Al-Azhar J. Dent. Sci. 2020;23(1-15).
[239] Xie X, Wang L, Xing D, Zhang K, Weir MD, Liu H, et al. Novel dental adhesive with
triple benefits of calcium phosphate recharge, protein-repellent and antibacterial functions.
Dent Mater. 2017;33(5):553-63.
[240] Tao S, He L, Xu HH, Weir MD, Fan M, Yu Z, et al. Dentin remineralization via adhesive
containing amorphous calcium phosphate nanoparticles in a biofilm-challenged environment.
J Dent. 2019;89:103193.
[241] Braga RR. Calcium phosphates as ion-releasing fillers in restorative resin-based
materials. Dent Mater. 2019;35(1):3-14.
[242] Khan AS, Syed MR. A review of bioceramics-based dental restorative materials. Dent
Mater J. 2019;38(2):163-76.
[243] Arifa MK, Ephraim R, Rajamani T. Recent advances in dental hard tissue
remineralization: a review of literature. Int J Clin Pediatr Dent. 2019;12(2):139.

62
[244] Ronay FC, Wegehaupt FJ, Becker K, Wiedemeier DB, Attin T, Lussi A, et al. Pure
hydroxyapatite as a substitute for enamel in erosion experiments. J Dent. 2019;84:89-94.
[245] Altaie A, Bubb N, Franklin P, German MJ, Marie A, Wood DJ. Development and
characterisation of dental composites containing anisotropic fluorapatite bundles and rods.
Dent Mater. 2020;36(8):1071-85.
[246] Xu HH, Sun L, Weir MD, Takagi S, Chow LC, Hockey B. Effects of incorporating
nanosized calcium phosphate particles on properties of whisker‐reinforced dental composites.
J Biomed Mater Res Part B Appl Biomater. 2007;81(1):116-25.
[247] Hench LL. The story of Bioglass®. J Mater Sci Mater Med. 2006;17(11):967-78.
[248] Al-Eesa N, Fernandes SD, Hill R, Wong F, Jargalsaikhan U, Shahid S. Remineralising
fluorine containing bioactive glass composites. Dent Mater. 2021;37(4):672-81.
[249] Gupta T, Nagaraja S, Mathew S, Narayana IH, Madhu K, Dinesh K. Effect of
desensitization using bioactive glass, hydroxyapatite, and diode laser on the shear bond
strength of resin composites measured at different time intervals: An In vitro Study. Contemp
Clin Dent. 2017;8(2):244.
[250] Par M, Mohn D, Attin T, Tarle Z, Tauböck TT. Polymerization shrinkage behaviour of
resin composites functionalized with unsilanized bioactive glass fillers. Sci Rep. 2020;10(1):1-
10.
[251] Attin T, Tarle Z, Tauböck TT. A new customized bioactive glass filler to functionalize
resin composites: acid-neutralizing capability, degree of conversion, and apatite precipitation.
J Clin Med. 2020;9(4):1173.
[252] Han X, Chen Y, Jiang Q, Liu X, Chen Y. Novel bioactive glass-modified hybrid
composite resin: mechanical properties, biocompatibility, and antibacterial and remineralizing
activity. Front bioeng biotechnol. 2021;9.
[253] Medjedovic E, Medjedovic S, Deljo D, Sukalo A. Impact of fluoride on dental health
quality. Mater Sociomed. 2015;27(6):395.
[254] Marquis RE. Antimicrobial actions of fluoride for oral bacteria. Can J Microbiol.
1995;41(11):955-964.
[255] Philip N. State of the art enamel remineralization systems: the next frontier in caries
management. Caries Res. 2019;53(3):284-295.
[256] Cao J, Yang DL, Pu Y, Wang D, Wang JX. CaF2/SiO2 core–shell nanoparticles as novel
fillers with reinforced mechanical properties and sustained fluoride ion release for dental resin
composites. J Mater Sci. 2021;56(29):16648-16660.
[257] Su LW, Lin DJ, Uan JY. Novel dental resin composites containing LiAl-F layered double
hydroxide (LDH) filler: Fluoride release/recharge, mechanical properties, color change, and
cytotoxicity. Dent Mater. 2019;35(5):663-672.
[258] Hoxha A, Gillam DG, Agha A, Karpukhina N, Bushby AJ, Patel MP. Novel fluoride
rechargeable dental composites containing MgAl and CaAl layered double hydroxide (LDH).
Dent Mater. 2020;36(8):973-986.
[259] Mitwalli H, Balhaddad AA, AlSahafi R, Oates TW, Melo MAS, Xu HH, Weir MD.
Novel CaF2 nanocomposites with antibacterial function and fluoride and calcium ion release
to inhibit oral biofilm and protect teeth. J Funct Biomater. 2020;11(3):56.

63
[260] Dai Q, Weir MD, Ruan J, Liu J, Gao J, Lynch CD, et al. Effect of co-precipitation plus
spray-drying of nano-CaF2 on mechanical and fluoride properties of nanocomposite. Dent
Mater. 2021;37(6):1009-1019.
[261] Reis DP, Noronha Filho JD, Rossi AL, de Almeida Neves A, Portela MB, da Silva EM.
Remineralizing potential of dental composites containing silanized silica-hydroxyapatite (Si-
HAp) nanoporous particles charged with sodium fluoride (NaF). J Dent. 2019;90:103211.
[262] Hesaraki S, Karimi M, Nezafati N. The synergistic effects of SrF2 nanoparticles, YSZ
nanoparticles, and poly-ε-l-lysin on physicomechanical, ion release, and antibacterial-cellular
behavior of the flowable dental composites. Mater Sci Eng C. 2020;109:110592.
[263] Rekow ED. Digital dentistry: The new state of the art-Is it disruptive or destructive? Dent
Mater. 2020;36(1):9-24.
[264] Sulaiman TA. Materials in digital dentistry-A review. J Esthet Restor Dent.
2020;32(2):171-81.
[265] Mai H-N, Lee K-B, Lee D-H. Fit of interim crowns fabricated using photopolymer-
jetting 3D printing. J Prosthet Dent. 2017;118(2):208-15.
[266] Park G-S, Kim S-K, Heo S-J, Koak J-Y, Seo D-G. Effects of printing parameters on the
fit of implant-supported 3D printing resin prosthetics. Materials. 2019;12(16):2533.
[267] Tahayeri A, Morgan M, Fugolin AP, Bompolaki D, Athirasala A, Pfeifer CS, et al. 3D
printed versus conventionally cured provisional crown and bridge dental materials. Dent Mater.
2018;34(2):192-200.
[268] Yeung M, Abdulmajeed A, Carrico CK, Deeb GR, Bencharit S. Accuracy and precision
of 3D-printed implant surgical guides with different implant systems: An in vitro study. J
Prosthet Dent. 2020;123(6):821-8.
[269] Jindal P, Juneja M, Siena FL, Bajaj D, Breedon P. Mechanical and geometric properties
of thermoformed and 3D printed clear dental aligners. Am J Orthod Dentofacial Orthop.
2019;156(5):694-701.
[270] Sa L, Kaiwu L, Shenggui C, Junzhong Y, Yongguang J, Lin W, Li R. 3D printing dental
composite resins with sustaining antibacterial ability. J Mater Sci. 2019;54(4):3309-3318.
[271] Alsandi Q, Ikeda M, Arisaka Y, Nikaido T, Tsuchida Y, Sadr A, et al. Evaluation of
Mechanical and Physical Properties of Light and Heat Polymerized UDMA for DLP 3D Printer.
Sensors. 2021;21(10):3331.
[272] Rosentritt M, Huber C, Strasser T, Schmid A. Investigating the mechanical and optical
properties of novel Urethandimethacrylate (UDMA) and Urethanmethacrylate (UMA) based
rapid prototyping materials. Dent Mater. 2021;37(10):1584-1591.
[273] Sodeyama MK, Ikeda H, Nagamatsu Y, Masaki C, Hosokawa R, Shimizu H. Printable
PICN Composite Mechanically Compatible with Human Teeth. J Dent Res. 2021.
[274] Kessler A, Reichl FX, Folwaczny M, Högg C. Monomer release from surgical guide
resins manufactured with different 3D printing devices. Dent Mater. 2020;36(11):1486-1492.
[275] Lin C-H, Lin Y-M, Lai Y-L, Lee S-Y. Mechanical properties, accuracy, and cytotoxicity
of UV-polymerized 3D printing resins composed of Bis-EMA, UDMA, and TEGDMA. J
Prosthet Dent. 2020;123(2):349-54.

64
[276] Derban P, Negrea R, Rominu M, Marsavina L. Influence of the printing angle and load
direction on flexure strength in 3D printed materials for provisional dental restorations.
Materials. 2021;14(12):3376.
[277] Xia J, Li Y, Cai D, Shi X, Zhao S, Jiang Q, et al. Direct resin composite restoration of
maxillary central incisors using a 3D-printed template: two clinical cases. BMC oral health.
2018;18(1):1-8.
[278] Liao W, Xu L, Wangrao K, Du Y, Xiong Q, Yao Y. Three-dimensional printing with
biomaterials in craniofacial and dental tissue engineering. PeerJ. 2019;7:e7271.
[279] Obregon F, Vaquette C, Ivanovski S, Hutmacher D, Bertassoni L. Three-dimensional
bioprinting for regenerative dentistry and craniofacial tissue engineering. J Dent Res.
2015;94(9_suppl):143S-52S.
[280] Xu W, Jambhulkar S, Zhu Y, Ravichandran D, Kakarla M, Vernon B, et al. 3D printing
for polymer/particle-based processing: A review. Composites, Part B. 2021:109102.
[281] Lin L, Fang Y, Liao Y, Chen G, Gao C, Zhu P. 3D printing and digital processing
techniques in dentistry: A review of literature. Adv Eng Mater. 2019;21(6):1801013.
[282] Della Bona A, Cantelli V, Britto VT, Collares KF, Stansbury JW. 3D printing restorative
materials using a stereolithographic technique: A systematic review. Dent Mater. 2021.

65

View publication stats

You might also like