Next Article in Journal
An Approach to Testing Antivandal Composite Materials as a Function of Their Thickness and Striker Shape—A Case Study
Next Article in Special Issue
Evaluation of the Mechanical and Adhesion Characteristics of Indirect Restorations Manufactured with Three-Dimensional Printing
Previous Article in Journal
Mucoadhesive Polymeric Polyologels Designed for the Treatment of Periodontal and Related Diseases of the Oral Cavity
Previous Article in Special Issue
Influence of Post-Processing on the Degree of Conversion and Mechanical Properties of 3D-Printed Polyurethane Aligners
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanical Performance of Extensive Restorations Made with Short Fiber-Reinforced Composites without Coverage: A Systematic Review of In Vitro Studies

by
András Jakab
1,†,
Dániel Palkovics
2,
Veronika T. Szabó
1,†,
Balázs Szabó
3,
Eszter Vincze-Bandi
4,
Gábor Braunitzer
5,
Lippo Lassila
6,
Pekka Vallittu
6,
Sufyan Garoushi
6,‡ and
Márk Fráter
1,*,‡
1
Department of Operative and Esthetic Dentistry, Faculty of Dentistry, University of Szeged, H-6720 Szeged, Hungary
2
Department of Periodontology, Faculty of Dentistry, Semmelweis University, H-1088 Budapest, Hungary
3
Department of Periodontology, Faculty of Dentistry, University of Szeged, H-6720 Szeged, Hungary
4
Dr. Borbáth Dental and Implantology Center, H-6800 Hódmezővásárhely, Hungary
5
dicomLAB Dental Ltd., H-6726 Szeged, Hungary
6
Department of Biomaterials Science and Turku Clinical Biomaterials Center—TCBC, Institute of Dentistry, University of Turku, FI-20520 Turku, Finland
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
These authors contributed equally to this work.
Polymers 2024, 16(5), 590; https://doi.org/10.3390/polym16050590
Submission received: 31 December 2023 / Revised: 13 February 2024 / Accepted: 15 February 2024 / Published: 21 February 2024
(This article belongs to the Special Issue Advanced Polymeric Materials for Dental Applications II)

Abstract

:
In recent years, composite resin materials have been the most frequently used materials for direct restorations of posterior teeth. These materials have some clinically relevant limitations due to their lack of fracture toughness, especially when used in larger cavities with high volume factors or when utilized as direct or indirect overlays or crown restorations. Recently, short-fiber-reinforced composite materials have been used in bi-structure restorations as a dentine substituting material due to their superior mechanical properties; however, there is no scientific consensus as to whether they can be used as full restorations. The aim of our review was to examine the available literature and gather scientific evidence on this matter. Two independent authors performed a thorough literature search using PubMed and ScienceDirect up until December 2023. This study followed the PRISMA guidelines, and the risk of bias was assessed using the QUIN tool. The authors selected in vitro studies that used short-fiber-reinforced composite materials as complete restorations, with a conventional composite material as a comparison group. Out of 2079 potentially relevant articles, 16 met our inclusion criteria. All of the included studies reported that the usage of short-fiber-reinforced composites improved the restoration’s load-bearing capacity. Fifteen of the included publications examined the fracture pattern, and thirteen of them reported a more favorable fracture outcome for the short-fiber-reinforced group. Only one article reported a more favorable fracture pattern for the control group; however, the difference between groups was not significant. Within the limitations of this review, the evidence suggests that short-fiber-reinforced composites can be used effectively as complete restorations to reinforce structurally compromised teeth.

1. Introduction

With the development of adhesive techniques, clinicians are routinely using composite fillings in both the anterior and posterior regions [1,2]. Modern restorations are expected to not only restore the aesthetics and function of teeth, but also to strengthen the remaining tooth material, protecting against fractures [3]. Excessive cavity preparations and root canal treatments make the teeth more prone to fractures [4,5]. The more dentine that is lost due to caries, previous restorations, or fractures, the less resistant teeth become to fracture. The loss of a single marginal ridge can lead to a 46% loss in tooth rigidity in premolars, and in the case of a mesio-occluso-distal (MOD) cavity, this ratio can be as high as 63% [6]. Besides the absence of these structurally important elements, the depth of the cavity plays a significant role in the mechanical integrity of the teeth. The volume factor, which refers to the size of the missing tooth material, is a factor with key importance regarding the weakening effect of the lesion [7,8]. It has been shown that while a shallower cavity (referring to minimal hard tissue loss) can be safely restored with a direct composite restoration [9], more extensive restorations made with such restorative techniques exhibit significantly lower fracture resistance [10].
In recent years, composite resin materials have become the most frequently used materials for direct restorations of posterior teeth. However, due to their physical limitations, clinical failure may occur in the case of extensive restorations. The two most concerning problems with conventional particulate filler composites (PFCs) are polymerization shrinkage [11,12,13,14] and that—even though they are brittle, strong materials—they lack fracture toughness [15]. The problem of low fracture toughness is particularly evident in larger direct restorations, as the volume of the material increases. Fracture toughness is a mechanical property that determines how well brittle materials resist the catastrophic propagation of flaws under an applied load [16,17,18]. This parameter is considered to be the measure of fatigue resistance, which predicts structural performance [19].
Short-fiber-reinforced composite (SFRC) materials seem to offer a solution to this problem, as these materials are much closer to dentine in terms of their physical parameters [16]. In previous studies, it has been shown that even deep MOD cavities can be reinforced by the application of SFRCs in direct restorations [15,20]. Due to this reinforcing effect, in recent years, numerous innovative restoration techniques and materials have appeared. The use of SFRCs has expanded the possible applications of direct extensive restorations both in vital and root-canal-treated teeth. For structurally compromised teeth, an indirect restoration was usually performed in the past; however, new materials and techniques may allow clinicians to perform a less time-consuming and more cost-effective extensive direct restoration in these cases (such as direct crowns or overlays). The question arises if maximizing the SFRC content in restorations could maximize the potential reinforcing effect in direct and indirect situations. Our study aims to collect and evaluate the available evidence on this subject.

2. Materials and Methods

This systematic review was carried out according to the Preferred Reporting Items for Systematic Review and Meta-Analysis guidelines (PRISMA) [21].

2.1. Search Strategy

Sources: A review of the literature was conducted using PubMed and ScienceDirect until December 2023. Articles were gathered using keywords such as “fiber reinforcement”, “fiber-reinforced composite”, “fiber-reinforced restoration”, “SFRC”, and “EverX”. While our emphasis was on articles from the current decade, a few older publications (n = 4) were incorporated due to their high relevance. Boolean operators (“AND” and “OR”) were employed to connect the keywords, and these terms were adjusted accordingly for each database. Table 1 shows the used search strategies.

2.2. Study Screening and Selection

The collected articles were organized in a reference management software (Mendeley Reference Manager 2.106.0). After scanning for duplicates, two independent examiners (A.J. and M.F.) carefully reviewed the title and the abstract of the articles. A title was discarded if both examiners agreed it was irrelevant. Eligible abstracts were selected for a full-text review.

2.3. Eligibility Criteria

Eligible studies were in vitro studies that used short-fiber-reinforced composite materials as complete restorations, compared to a control group restored with conventional particulate filler composite materials. Only studies testing load-bearing capacity were included. Further criteria for eligibility encompassed studies published in peer-reviewed journals and written in English. These studies were required to incorporate the search terms within either the title or the abstract.

2.4. Data Synthesis

The studies meeting the inclusion criteria underwent thorough examination, with relevant information systematically gathered in a Microsoft Excel document. For each included publication, authors’ names, article title, publication year, experimental and control groups, type of short fiber employed, restoration type, as well as the primary outcomes and conclusions were documented.

2.5. Quality Assessment

The risk of bias was determined for each included publication using the Quality Assessment Tool for In Vitro Studies (QUIN), which consists of 12 criteria [22]. Each criterion is scored from 0 to 2 (0: not specified; 1: inadequately specified; 2: adequately specified; NA: not applicable). The final score is determined by the following formula: final score = (total score × 100)/(2 × number of criteria applicable). According to the final score, the article is graded as having low, medium, or high risk of bias (>70%, low risk of bias; 50–70%, medium risk of bias; <50%, high risk of bias).

3. Results

After the application of our search strategy, 2079 records were identified. After the removal of the duplicates (n = 114), 1995 articles were submitted to title and abstract screening. After screening, 19 articles were examined on the full-text level. After the removal of articles that did not meet our inclusion criteria (n = 3), 16 articles were included in the final review. Figure 1 shows the screening and selection process in a PRISMA flow diagram [21].
Four of the sixteen studies used extracted human teeth [23,24,25,26], while twelve studies used artificial plastic [27,28,29], zirconia [30,31,32,33,34,35], or cobalt-chromium [36,37,38] models as a base for the restorations. All included studies conducted mechanical testing to measure the load-bearing capacity of the restorations. After loading, they examined the fracture patterns, too. Three of the included publications involved further material investigations and measured some of the physical parameters of the tested materials, such as fracture toughness [27,34,37]. In accordance with our inclusion criteria, all the selected studies tested restorations solely made with an SFRC material and included a comparison group made with conventional PFC. Ten of the selected studies used an SFRC with a paste consistency (experimental fiber composite, EverX Posterior) [23,24,25,29,30,31,32,33,34,35], six of them used a flowable SFRC [26,27,28,36,37,38], and four of them made restorations using a short-fiber-reinforced CAD-CAM composite material [30,32,36,37]. All the included studies reported significantly higher load-bearing capacities for the groups restored solely with SFRCs compared to the PFC control groups. In one of the selected publications, where the authors tested both a paste and a CAD-CAM SFRC material for inlay-retained fixed partial dentures (IRFPDs), the PFC group performed slightly better than the paste-consistency SFRC group [30]. However, in this case, the CAD-CAM SFRC material performed significantly better than all the other tested groups in terms of load-bearing capacity. In another publication, the authors tested the materials both with and without a conventional fiber-reinforced composite (FRC) post in the restoration [29]. In this case, the SFRC group without the FRC post outperformed the control group. However, when a conventional FRC post was applied in the restoration, the PFC group performed better. Furthermore, in those articles where a material analysis was performed, in all cases, significantly higher fracture toughness was measured in the SFRC groups. Fifteen of the included studies examined the fracture pattern in the failed specimens [23,24,25,26,28,29,30,31,32,33,34,35,36,38]. In thirteen cases, a more favorable fracture pattern was reported for the SFRC groups compared to their controls. Bielic et al. found that both the SFRC and the control groups had favorable fracture patterns [29]. In one case, the PFC group showed a slightly more favorable pattern, but the difference was statistically not significant [24]. Table 2 summarizes the details of the included publications.

3.1. Quality Assessment of the Included In Vitro Trials

The details of the methodological quality of the publications are shown in Table 3. Thirteen of the publications had medium risk of bias, while three of the studies had low risk [24,26,38].

3.2. Quantitative Assessment of the Included Publications

The highest median value for LBC was measured in a group restored by flowable SFRC (EverX Flow) onlays at 3990 N (±330 N) [28]. The highest fracture toughness was exhibited by an FRC CAD-CAM material at 2.9 MPa/m2 [37], and the lowest was 1.0 MPa/m2, which was measured for a conventional composite filling material [34]. Table 4 summarizes the methodology of the mechanical tests applied in the assessed studies, while Table 5 summarizes the load-bearing results from the assessed studies.

4. Discussion

Fiber reinforcement for restoring structurally compromised teeth is a frequently discussed subject. Restorations after endodontic treatments, large decays, and replacement of large fillings are part of the everyday dental routine, and there is a constant need for the reinforcement of these teeth. Although PFC materials have some limitations in terms of their physical properties, they are routinely used to restore large MOD cavities in molars and premolars [39,40]. One approach to tackling the issue of polymerization shrinkage is layering the material while considering the variable bondability across various dental tissues. With the “decoupling with time” concept, clinicians can counter the stresses caused by the material’s shrinkage during and after polymerization [40]. A generally accepted biomimetic restorative approach suggests replacing enamel with feldspathic porcelain or glass ceramic and dentine with hybrid composites [41]. Even though this technique shows promising results in the case of performing indirect restorations, it does not take ideal fracture toughness into consideration, and the lack of toughness can lead to catastrophic fractures. Although certain mechanical properties of conventional composites, like flexural strength, flexural modulus, and thermal expansion coefficient, may resemble those of dentine from a biomimetic perspective, the material’s microstructure significantly differs from that of dentine [41]. Conventional PFC materials consist of filler particles embedded in a resin matrix, while dentine consists of collagen fibers in a hydroxyapatite matrix. As a result, PFC materials are rigid, brittle materials and they are incapable of stopping crack propagation under loading, which can lead to irreparable, catastrophic fractures [16]. For the above-mentioned reasons, direct composite restorations might not be the ideal solution to higher-volume-factor cavities. Fiber reinforcement in composite restorations can strengthen the compromised tooth and the restoration. These fibers can make the composite’s microstructure more dentine-like, as they provide the material a much higher fracture toughness than PFC; thus, they are more suitable for restoring missing dentine. The volume of the material, the size and the orientation of the fibers, and the fibers’ adhesion to the matrix can all influence the mechanical performance of the material. The transfer of stress from the polymer matrix to the fibers greatly affects the material’s mechanical performance [42]. Also, critical fiber length, which is calculated from the diameter from the fibers, is of paramount importance. For advanced fiber-reinforced composites, the length of the fibers should be as much as 50 times the diameter of the fiber [16]. In the past, there have been two low-aspect-ratio packable SFRC materials (Alert, Pentron; Nulite F, NSI Dental) available on the market. Although they exhibited acceptable mechanical performance, some clinical issues, including surface roughness and wear resistance, as well as material and cusp fractures, hindered their widespread adoption as a general dentine substitute [43,44]. Alert has a fiber length of 60–80 μm, with a 6–10-micrometer diameter, and Nulite F has a fiber length of 150–200 μm with a 9 μm diameter. In both cases, the fibers are well below the critical fiber length, which explains the above-mentioned problems [42]. This is also in accordance with the difference in fracture toughness between these earlier SFRCs and the more advanced, modern materials. The fracture toughness of Alert is measured at 1.7 MPa/m2, while that of EverX Flow is measured at 2.8 MPa/m2 [37]. In modern SFRC materials, the fibers are randomly oriented, which means they are capable of strengthening the material in every direction. Longer, conventional, unidirectional fiber-containing posts often deliver a stronger reinforcement, yet achieving a precise orientation can pose practical challenges. Nowadays, SFRC materials are often used to replace dentine because they are easy to use and provide a quick solution to the restoration of heavily compromised teeth [20,45]. Excessive wear and surface roughness problems associated with earlier SFRC materials appear to be resolved. EverX Flow had significantly lower wear value after 15,000 chewing simulation cycles than conventional PFCs, and it did not exhibit a coarse surface after the test. This SFRC material fulfilled the American Dental Association’s criterion for wear [46]. In an earlier study, after 100,000 cycles of brushing simulation, it was found that SFRC had a similar wear and surface roughness to that of PFC [47]. During the polishing of these materials, the fibers on the surface suffer microfractures; thus, they can be polished together with the resin matrix [46]. As a future perspective, with the advancement of these SFRC materials, it might be possible to use them not just as a dentine substitute, but rather for complete restorations. A usual concern regarding these materials is their water absorption after restoration. Lassila et al. compared the water absorption of flowable SFRC to conventional bulk-fill PFC. After 36 days, the material’s water absorption was the second best (0.5 wt%), while other regular bulk-fill composite materials had much higher values (Estelite Bulk Fill Flow 1.1 wt%) [46]. Regarding fracture patterns, SFRCs show a promising performance in stopping and deflecting crack formation [48,49,50]. As it is the fracture pattern that will determine the restorability of teeth in case of a fracture, it is of high importance. According to Scotti and co-workers, a distinction can be made between restorable or non-restorable fractures based on optical microscopic observation (preferably with a two-examiner agreement). A restorable fracture is above the cemento-enamel junction (CEJ), meaning that in case of fracture, the tooth can be restored, while a non-restorable fracture extends below the CEJ and the tooth will likely need to be extracted [51]. The energy-absorbing and stress-distributing fibers allow crack propagation to be deflected toward the peripheries of the material, which can result in a much more favorable fracture pattern.
For the above reasons, this review aimed to gather information on whether the maximization of SFRCs can further strengthen restorations and if their use contributes to a more favorable fracture pattern.

4.1. Fracture Resistance

Fracture resistance is one of the most important characteristics of materials used in posterior region as it determines the longevity of the restorations. All the included studies tested the restorations’ load-bearing capacity. In clinical situations, the load-bearing capacity is highly dependent on the quantity and quality of the remaining tooth structure. Four of the included studies used extracted human teeth for these tests [23,24,25,26]. While these in vitro investigations have a high clinical relevance, the usage of human teeth adds another variable to this equation in terms of mechanical parameters and survival of the restoration–tooth unit. In these cases, the measured load-bearing capacity values came out much lower than in the rest of the included publications, where they used standardized, artificial models.
In all but one of the studies, the uncovered SFRC group performed significantly better than the control PFC group. Bijelic et al. did not compare the SFRC and the control groups directly [29]. Even though the SFRC group performed better as an anterior crown restoration without the application of an FRC post (238 ± 43 N), with the insertion of a post, the PFC group performed slightly better (265 ± 59 N). The authors did not report a significant difference between the groups [29]. Apart from this study, all the other articles reported a significantly higher fracture resistance for the SFRC group compared to the PFC control group. Eight of the included studies compared an uncovered, plain SFRC group to a bi-layered SFRC + PFC group as well [25,26,28,31,34,35,38]. In all eight publications, the uncovered SFRC performed better than the bi-layered group. Lassila et al. compared EverX Flow covered by 2 mm PFC on the occlusal surface to uncovered flowable SFRC in posterior crown restorations. They found that the uncovered group (3866 ± 263 N) performed significantly better (p < 0.05) than the covered ones (2000–2250 N) [27]. In another publication, they compared posterior crown restorations made with a flowable SFRC (EverX Flow) core covered by PFC layers of different thicknesses (0.5, 1, 1.5, 2). They also included an uncovered, plain SFRC group and a fully PFC-made group (1907.7 ± 179 N). Regression analysis showed that by decreasing the thickness of the PFC layer, the load-bearing capacity increased. The uncovered, plain SFRC group had a significantly higher load-bearing capacity than all the tested groups (3990 ± 330 N) [28]. Garoushi et al. tested direct onlay restorations made of plain PFC, SFRC + PFC, and plain SFRC [35]. Similarly to the crown restorations, the uncovered SFRC group exhibited the highest load-bearing capacity, which was significantly higher (1733 N) than that of the control PFC group (1081 N).

4.2. Fracture Pattern

Structurally compromised teeth are prone to fracture, mainly due to the high amount of lost tooth material, the presence of cracks, and the loss of structurally important parts, such as marginal ridges [52,53]. If a fracture occurs, the fracture pattern will determine the restorability of the tooth, and for that reason, it is of high importance.
Fifteen of the included articles investigated fracture patterns. Thirteen reported a favorable pattern for the SFRC groups [23,25,26,28,30,31,32,33,34,35,36,38]. Cekic-Nagas et al. tested both an SFRC paste composite (EverX Posterior) and a CAD-CAM SFRC material [30]. They found that while EverX Posterior did not perform significantly better than the control group, the CAD-CAM material showed a significantly more favorable fracture pattern than the rest of the groups. This could be explained by the amount of fibers in the material. The CAD-CAM version is more similar in terms of type of fibers to the flowable version of EverX, which means that because of the shorter fibers, the material has a higher fiber ratio. Bielic et al. found that both the SFRC and the control group had favorable fracture patterns [29]. In another publication, they reported that the control group showed a slightly more favorable fracture pattern, but the difference from the SFRC group was not statistically significant [24].
Despite the manufacturer’s recommendation to use SFRCs exclusively as a substitute for missing dentin in direct and indirect restorations while avoiding contact with the oral environment, the study’s findings suggest that employing SFRCs without surface coverage in various simulated clinical scenarios yielded satisfactory outcomes in terms of restoration fracture behavior. This might improve the clinical performance of extensive direct composite restorations; however, it is necessary to conduct clinical trials to obtain more reliable and conclusive findings.

5. Conclusions

Within the limitations of this review, in vitro evidence suggests that short-fiber-reinforced composite materials effectively reinforce structurally compromised teeth. These materials exhibit higher fracture toughness, akin to dentine, imparting greater resilience, and the restorations created with SFRCs demonstrate heightened load-bearing capacities. Regarding fracture patterns, most publications documented more favorable outcomes when using SFRCs. Some studies compared groups using SFRCs with and without a PFC covering, and in these instances, the uncovered SFRC groups demonstrated superior performance. For these reasons and due to advancements in areas such as two-body wear and surface roughness, SFRCs may hold promise for future clinical use as complete restorations; however, further research is warranted in this area.

Author Contributions

Conceptualization, A.J., S.G. and M.F.; methodology, A.J.; software, D.P. and A.J.; validation, E.V.-B. and M.F.; investigation, B.S., V.T.S., D.P. and A.J.; data curation, E.V.-B. and M.F.; writing—original draft preparation, A.J. and G.B.; writing—review and editing, L.L., P.V., S.G. and M.F.; visualization, S.G. and A.J.; supervision, S.G. and M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the ÚNKP-23-3-SZTE-147 New National Excellence Program of the Ministry of Hungary for Innovation and Technology from the Source of the National Research, Development and Innovation Fund.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

Gábor Braunitzer was employed by the dicomLAB Dental Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Bonilla, E.D.; Hayashi, M.; Pameijer, C.H.; Le, N.V.; Morrow, B.R.; Garcia-Godoy, F. The Effect of Two Composite Placement Techniques on Fracture Resistance of MOD Restorations with Various Resin Composites. J. Dent. 2020, 101, 103348. [Google Scholar] [CrossRef]
  2. Haak, R.; Näke, T.; Park, K.-J.; Ziebolz, D.; Krause, F.; Schneider, H. Internal and Marginal Adaptation of High-Viscosity Bulk-Fill Composites in Class II Cavities Placed with Different Adhesive Strategies. Odontology 2019, 107, 374–382. [Google Scholar] [CrossRef]
  3. Valizadeh, S.; Ranjbar Omrani, L.; Deliperi, S.; Sadeghi Mahounak, F. Restoration of a Nonvital Tooth with Fiber Reinforce Composite (Wallpapering Technique). Case Rep. Dent. 2020, 2020, 9619787. [Google Scholar] [CrossRef]
  4. Gaeta, C.; Marruganti, C.; Mignosa, E.; Franciosi, G.; Ferrari, E.; Grandini, S. Influence of Methodological Variables on Fracture Strength Tests Results of Premolars with Different Number of Residual Walls. A Systematic Review with Meta-Analysis. Dent. J. 2021, 9, 146. [Google Scholar] [CrossRef] [PubMed]
  5. Patel, S.; Bhuva, B.; Bose, R. Present Status and Future Directions: Vertical Root Fractures in Root Filled Teeth. Int. Endod. J. 2022, 55, 804–826. [Google Scholar] [CrossRef] [PubMed]
  6. Daher, R.; Ardu, S.; Di Bella, E.; Rocca, G.T.; Feilzer, A.J.; Krejci, I. Fracture Strength of Non-Invasively Reinforced MOD Cavities on Endodontically Treated Teeth. Odontology 2021, 109, 368–375. [Google Scholar] [CrossRef]
  7. Magne, P.; Carvalho, M.A.; Milani, T. Shrinkage-Induced Cuspal Deformation and Strength of Three Different Short Fiber-Reinforced Composite Resins. J. Esthet. Restor. Dent. 2023, 35, 56–63. [Google Scholar] [CrossRef] [PubMed]
  8. Magne, P.; Milani, T. Short-Fiber Reinforced MOD Restorations of Molars with Severely Undermined Cusps. J. Adhes. Dent. 2023, 25, 99–106. [Google Scholar] [CrossRef] [PubMed]
  9. Sidhu, P.; Sultan, O.S.; Math, S.Y.; Malik, N.A.; Wilson, N.H.F.; Lynch, C.D.; Blum, I.R.; Daood, U. Current and Future Trends in the Teaching of Direct Posterior Resin Composites in Malaysian Dental Schools: A Cross-Sectional Study. J. Dent. 2021, 110, 103683. [Google Scholar] [CrossRef] [PubMed]
  10. European Society of Endodontology developed by:; Mannocci, F.; Bhuva, B.; Roig, M.; Zarow, M.; Bitter, K. European Society of Endodontology Position Statement: The Restoration of Root Filled Teeth. Int. Endod. J. 2021, 54, 1974–1981. [Google Scholar] [CrossRef]
  11. Sadr, A.; Bakhtiari, B.; Hayashi, J.; Luong, M.N.; Chen, Y.-W.; Chyz, G.; Chan, D.; Tagami, J. Effects of Fiber Reinforcement on Adaptation and Bond Strength of a Bulk-Fill Composite in Deep Preparations. Dent. Mater. 2020, 36, 527–534. [Google Scholar] [CrossRef]
  12. Lempel, E.; Szebeni, D.; Őri, Z.; Kiss, T.; Szalma, J.; Lovász, B.V.; Kunsági-Máté, S.; Böddi, K. The Effect of High-Irradiance Rapid Polymerization on Degree of Conversion, Monomer Elution, Polymerization Shrinkage and Porosity of Bulk-Fill Resin Composites. Dent. Mater. 2023, 39, 442–453. [Google Scholar] [CrossRef]
  13. Tseng, P.-C.; Chuang, S.-F.; Kaisarly, D.; Kunzelmann, K.-H. Simulating the Shrinkage-Induced Interfacial Damage around Class I Composite Resin Restorations with Damage Mechanics. Dent. Mater. 2023, 39, 513–521. [Google Scholar] [CrossRef]
  14. Ohmori, K.; Tasaki, T.; Kimura, S.; Hori, A.; Sakaeda, N.; Hanabusa, M.; Yamamoto, T. Residual Polymerization Stresses in Human Premolars Generated with Class II Composite Restorations. J. Mech. Behav. Biomed. Mater. 2020, 104, 103643. [Google Scholar] [CrossRef]
  15. Tsertsidou, V.; Mourouzis, P.; Dionysopoulos, D.; Pandoleon, P.; Tolidis, K. Fracture Resistance of Class II MOD Cavities Restored by Direct and Indirect Techniques and Different Materials Combination. Polymers 2023, 15, 3413. [Google Scholar] [CrossRef]
  16. Lassila, L.; Keulemans, F.; Säilynoja, E.; Vallittu, P.K.; Garoushi, S. Mechanical Properties and Fracture Behavior of Flowable Fiber Reinforced Composite Restorations. Dent. Mater. 2018, 34, 598–606. [Google Scholar] [CrossRef] [PubMed]
  17. Yang, J.; Silikas, N.; Watts, D.C. Polymerization and Shrinkage Kinetics and Fracture Toughness of Bulk-Fill Resin-Composites. Dent. Mater. 2022, 38, 1934–1941. [Google Scholar] [CrossRef]
  18. Fráter, M.; Sáry, T.; Jókai, B.; Braunitzer, G.; Säilynoja, E.; Vallittu, P.K.; Lassila, L.; Garoushi, S. Fatigue Behavior of Endodontically Treated Premolars Restored with Different Fiber-Reinforced Designs. Dent. Mater. 2021, 37, 391–402. [Google Scholar] [CrossRef] [PubMed]
  19. Heintze, S.D.; Ilie, N.; Hickel, R.; Reis, A.; Loguercio, A.; Rousson, V. Laboratory Mechanical Parameters of Composite Resins and Their Relation to Fractures and Wear in Clinical Trials—A Systematic Review. Dent. Mater. 2017, 33, e101–e114. [Google Scholar] [CrossRef] [PubMed]
  20. Escobar, L.B.; Pereira da Silva, L.; Manarte-Monteiro, P. Fracture Resistance of Fiber-Reinforced Composite Restorations: A Systematic Review and Meta-Analysis. Polymers 2023, 15, 3802. [Google Scholar] [CrossRef]
  21. Page, M.J.; McKenzie, J.E.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. The PRISMA 2020 Statement: An Updated Guideline for Reporting Systematic Reviews. BMJ 2021, 372, n71. [Google Scholar] [CrossRef]
  22. Sheth, V.H.; Shah, N.P.; Jain, R.; Bhanushali, N.; Bhatnagar, V. Development and Validation of a Risk-of-Bias Tool for Assessing in Vitro Studies Conducted in Dentistry: The QUIN. J. Prosthet. Dent. 2022, S0022391322003456. [Google Scholar] [CrossRef] [PubMed]
  23. Garoushi, S.; Vallittu, P.K.; Lassila, L.V.J. Direct Restoration of Severely Damaged Incisors Using Short Fiber-Reinforced Composite Resin. J. Dent. 2007, 35, 731–736. [Google Scholar] [CrossRef] [PubMed]
  24. Bijelic, J.; Garoushi, S.; Vallittu, P.K.; Lassila, L.V.J. Short Fiber Reinforced Composite in Restoring Severely Damaged Incisors. Acta Odontol. Scand. 2013, 71, 1221–1231. [Google Scholar] [CrossRef] [PubMed]
  25. Garoushi, S.K.; Hatem, M.; Lassila, L.V.J.; Vallittu, P.K. The Effect of Short Fiber Composite Base on Microleakage and Load-Bearing Capacity of Posterior Restorations. Acta Biomater. Odontol. Scand. 2015, 1, 6–12. [Google Scholar] [CrossRef] [PubMed]
  26. Garoushi, S.; Akbaşak-Sungur, A.Ö.; Erkut, S.; Vallittu, P.K.; Uctasli, S.; Lassila, L. Evaluation of Fracture Behavior in Short Fiber-Reinforced Direct and Indirect Overlay Restorations. Clin. Oral Investig. 2023, 27, 5449–5458. [Google Scholar] [CrossRef] [PubMed]
  27. Lassila, L.; Säilynoja, E.; Prinssi, R.; Vallittu, P.K.; Garoushi, S. Fracture Behavior of Bi-Structure Fiber-Reinforced Composite Restorations. J. Mech. Behav. Biomed. Mater. 2020, 101, 103444. [Google Scholar] [CrossRef] [PubMed]
  28. Lassila, L.; Säilynoja, E.; Prinssi, R.; Vallittu, P.K.; Garoushi, S. Bilayered Composite Restoration: The Effect of Layer Thickness on Fracture Behavior. Biomater. Investig. Dent. 2020, 7, 80–85. [Google Scholar] [CrossRef]
  29. Bijelic, J. Fracture Load of Tooth Restored with Fiber Post and Experimental Short Fiber Composite. Open Dent. J. 2011, 5, 58–65. [Google Scholar] [CrossRef]
  30. Cekic-Nagas, I.; Egilmez, F.; Ergun, G.; Vallittu, P.K.; Lassila, L.V.J. Load-Bearing Capacity of Novel Resin-Based Fixed Dental Prosthesis Materials. Dent. Mater. J. 2018, 37, 49–58. [Google Scholar] [CrossRef]
  31. Garoushi, S.; Vallittu, P.K.; Lassila, L.V.J. Use of Short Fiber-Reinforced Composite with Semi-Interpenetrating Polymer Network Matrix in Fixed Partial Dentures. J. Dent. 2007, 35, 403–408. [Google Scholar] [CrossRef]
  32. Nagata, K.; Garoushi, S.K.; Vallittu, P.K.; Wakabayashi, N.; Takahashi, H.; Lassila, L.V.J. Fracture Behavior of Single-Structure Fiber-Reinforced Composite Restorations. Acta Biomater. Odontol. Scand. 2016, 2, 118–124. [Google Scholar] [CrossRef] [PubMed]
  33. Keulemans, F.; Lassila, L.V.J.; Garoushi, S.; Vallittu, P.K.; Kleverlaan, C.J.; Feilzer, A.J. The Influence of Framework Design on the Load-Bearing Capacity of Laboratory-Made Inlay-Retained Fibre-Reinforced Composite Fixed Dental Prostheses. J. Biomech. 2009, 42, 844–849. [Google Scholar] [CrossRef]
  34. Bijelic-Donova, J.; Garoushi, S.; Vallittu, P.K.; Lassila, L.V.J. Mechanical Properties, Fracture Resistance, and Fatigue Limits of Short Fiber Reinforced Dental Composite Resin. J. Prosthet. Dent. 2016, 115, 95–102. [Google Scholar] [CrossRef] [PubMed]
  35. Garoushi, S.; Mangoush, E.; Vallittu, M.; Lassila, L. Short Fiber Reinforced Composite: A New Alternative for Direct Onlay Restorations. Open Dent. J. 2013, 7, 181–185. [Google Scholar] [CrossRef]
  36. Lassila, L.; Mangoush, E.; Vallittu, P.K.; Garoushi, S. Fracture Behavior of Discontinuous Fiber-Reinforced Composite Inlay-Retained Fixed Partial Denture before and after Fatigue Aging. J. Prosthodont. Res. 2023, 67, 271–277. [Google Scholar] [CrossRef]
  37. Mangoush, E. Influence of Short Fiber Reinforced Composites on Fracture Resistance of Single-Structure Restorations. Eur. J. Prosthodont. Restor. Dent. 2020, 189. [Google Scholar] [CrossRef]
  38. Lassila, L.; Haapsaari, A.; Vallittu, P.K.; Garoushi, S. Fracture Resistance of Anterior Crowns Reinforced by Short-Fiber Composite. Polymers 2022, 14, 1809. [Google Scholar] [CrossRef]
  39. Ferracane, J.L.; Lawson, N.C. Probing the Hierarchy of Evidence to Identify the Best Strategy for Placing Class II Dental Composite Restorations Using Current Materials. J. Esthet. Restor. Dent. 2021, 33, 39–50. [Google Scholar] [CrossRef]
  40. Hofsteenge, J.W.; Scholtanus, J.D.; Özcan, M.; Nolte, I.M.; Cune, M.S.; Gresnigt, M.M.M. Clinical Longevity of Extensive Direct Resin Composite Restorations after Amalgam Replacement with a Mean Follow-up of 15 Years. J. Dent. 2023, 130, 104409. [Google Scholar] [CrossRef]
  41. Magne, P.; Belser, U. Biomimetic Restorative Dentistry, 2nd ed.; Quintessence Publishing Co., Inc.: Batavia, IL, USA, 2021; ISBN 978-0-86715-572-3. [Google Scholar]
  42. Alshabib, A.; Jurado, C.A.; Tsujimoto, A. Short Fiber-Reinforced Resin-Based Composites (SFRCs); Current Status and Future Perspectives. Dent. Mater. J. 2022, 41, 647–654. [Google Scholar] [CrossRef]
  43. Fagundes, T.C.; Barata, T.J.E.; Carvalho, C.A.R.; Franco, E.B.; Van Dijken, J.W.V.; Navarro, M.F.L. Clinical Evaluation of Two Packable Posterior Composites. J. Am. Dent. Assoc. 2009, 140, 447–454. [Google Scholar] [CrossRef]
  44. Van Dijken, J.W.V.; Sunnegårdh-Grönberg, K. Fiber-Reinforced Packable Resin Composites in Class II Cavities. J. Dent. 2006, 34, 763–769. [Google Scholar] [CrossRef] [PubMed]
  45. Aram, A.; Hong, H.; Song, C.; Bass, M.; Platt, J.A.; Chutinan, S. Physical Properties and Clinical Performance of Short Fiber Reinforced Resin-Based Composite in Posterior Dentition: Systematic Review and Meta-Analysis. Oper. Dent. 2023, 48, E119–E136. [Google Scholar] [CrossRef] [PubMed]
  46. Lassila, L.; Säilynoja, E.; Prinssi, R.; Vallittu, P.; Garoushi, S. Characterization of a New Fiber-Reinforced Flowable Composite. Odontology 2019, 107, 342–352. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, L.; Garcia, F.C.P.; De Araújo, P.A.; Franco, E.B.; Mondelli, R.F.L. Wear Resistance of Packable Resin Composites after Simulated Toothbrushing Test. J Esthet Restor Dent 2004, 16, 303–314. [Google Scholar] [CrossRef] [PubMed]
  48. Garoushi, S.; Gargoum, A.; Vallittu, P.K.; Lassila, L. Short Fiber-Reinforced Composite Restorations: A Review of the Current Literature. J. Investig. Clin. Dent. 2018, 9, e12330. [Google Scholar] [CrossRef] [PubMed]
  49. Tiu, J.; Belli, R.; Lohbauer, U. Rising R-Curves in Particulate/Fiber-Reinforced Resin Composite Layered Systems. J. Mech. Behav. Biomed. Mater. 2020, 103, 103537. [Google Scholar] [CrossRef] [PubMed]
  50. Selvaraj, H.; Krithikadatta, J.; Shrivastava, D.; Onazi, M.A.A.; Algarni, H.A.; Munaga, S.; Hamza, M.O.; Saad Al-Fridy, T.; Teja, K.V.; Janani, K.; et al. Systematic Review Fracture Resistance of Endodontically Treated Posterior Teeth Restored with Fiber Reinforced Composites—A Systematic Review. BMC Oral Health 2023, 23, 566. [Google Scholar] [CrossRef]
  51. Scotti, N.; Coero Borga, F.A.; Alovisi, M.; Rota, R.; Pasqualini, D.; Berutti, E. Is Fracture Resistance of Endodontically Treated Mandibular Molars Restored with Indirect Onlay Composite Restorations Influenced by Fibre Post Insertion? J. Dent. 2012, 40, 814–820. [Google Scholar] [CrossRef] [PubMed]
  52. Soares, L.M.; Razaghy, M.; Magne, P. Optimization of Large MOD Restorations: Composite Resin Inlays vs. Short Fiber-Reinforced Direct Restorations. Dent. Mater. 2018, 34, 587–597. [Google Scholar] [CrossRef] [PubMed]
  53. Aregawi, W.A.; Fok, A.S.L. Shrinkage Stress and Cuspal Deflection in MOD Restorations: Analytical Solutions and Design Guidelines. Dent. Mater. 2021, 37, 783–795. [Google Scholar] [CrossRef] [PubMed]
Figure 1. PRISMA flow diagram of the screening and selection process.
Figure 1. PRISMA flow diagram of the screening and selection process.
Polymers 16 00590 g001
Table 1. Search strategy in each database.
Table 1. Search strategy in each database.
DatabaseSearch Strategy
Pubmed((((Fiber reinforcement [Title/Abstract]) OR (fiber-reinforced composite [Title/Abstract])) OR (fiber-reinforced restoration [Title/Abstract])) OR (SFRC [Title/Abstract])) OR (EverX [Title/Abstract])
ScienceDirectin title/abstract: (“tooth” OR “teeth”) AND (“Fiber reinforcement” OR “fiber-reinforced composite” OR “fiber-reinforced restoration” OR “sfrc” OR “everx”)
Table 2. Details of the included publications.
Table 2. Details of the included publications.
First AuthorType of SFRCType of RestorationType of Mechanical TestingTested ParametersMain Conclusion
L. Lassila [27]EverX FlowPosterior crownStatical loadingLBC + FT + FP“Restorations combining a fiber-reinforced composite core and a surface layer of conventional composite, displayed promising performance related to fracture-behavior.”
I. Cekic-Nagas [30]EverX Posterior, SFRC CAD-CAMIRFPDStatical loadingLBC + FP“FRC had significantly affected load-bearing capacity of tested inlay-retained FDPs”
L. Lassila [36]EverX Flow, SFRC CAD-CAMIRFDPDynamic and statical loadingLBC + FP“CAD/CAM-fabricated IRFPDs made of experimental SFRC blocks have shown promising performance in clinical testing in terms of fracture behavior”.
S. Garoushi [23]Experimental FRCAnterior crownStatical loadingLBC + FP“Short glass fiber reinforced semi-IPN composite resin demonstrated improved load-bearing capacity compared with conventional particulate filler restorative composite resin”.
L. Lassila [28]EverX FlowPosterior crownStatical loadingLBC + FP“Restorations combining a thick FRC-core and a thin surface layer of PFC (0.5–1 mm), displayed promising performance related to fracture-behavior and load-bearing capacity”.
E. Mangoush [37]EverX Flow, SFRC CAD-CAMAnterior crownStatical loadingLBC + FT + FM“Within the limitation of this study, single-structure CAD/CAM fabricated restorations made of experimental SFRC blocks displayed promising performance related to fracture-behavior”.
S. Garoushi [31]Experimental FRCFDPStatical loadingLBC + FP“Short glass fiber reinforced semi-IPN composite resin revealed improvement in load-bearing capacity compared with the conventional particulate filler veneering composite”.
K. Nagata [32]EverX Posterior, SFRC CAD-CAMAnterior crownStatical loadingLBC + FP“Single-structure FRC restorations showed higher fracture resistance than the restorations made from PFC, leucite-reinforced ceramic, and lithium disilicate”.
J. Bijelic [29]Experimental FRCAnterior crownStatical loadingLBC + FP“…individually formed fiber-reinforced (FRC) root canal post improved the fracture load of the post-crown system and significantly contributed to the reinforcement and strengthening the restored teeth”.
S. Garoushi [35]EverX PosteriorOnlayStatical loadingLBC + FP“Onlay restorations combining base of short fiber reinforced composite resin as substructure and surface layer of conventional composite resin displayed promising performance in high load-bearing areas”.
L. Lassila [38]EverX FlowAnterior crownStatical loadingLBC + FP“Using SFC as core material with conventional PFC veneering composite to strengthen anterior crown restoration proved to be a promising strategy for further testing”.
J. Bijelic [24]Experimental FRCAnterior crownStatical loadingLBC + FP“...the use of SFC as a restorative material of choice for fabricating the direct composite post-core-crown restorations of severely damaged incisors provided an improved load-bearing capacity greater than CC…”
F. Keulemans [33]Experimental FRCInlay-retained FDPStatical loadingLBC + FP“…S-FRC seems to be a viable material for improving the framework of FRC-FDPs...”
J. Bijelic [34]EverX PosteriorAnterior crownDynamic and statical loadingLBC + FT + FM + FP“SFC crowns showed good performance under static and fatigue loading. FT was the only in vitro test method that filtered as a clinically relevant parameter”.
S. Garoushi [25]EverX PosteriorClass II filling and onlayStatical loadingMGF + LBC + FP“Based on the microleakage and compressive loading tests, base of short FRC resin and surface layer of conventional composite resin is the best combination”.
S. Garoushi [26]EverX FlowOverlayDynamic and statical loadingLBC + FP“The most effective method for restoring large MOD cavities was found to be direct restoration using SFC either alone or as a bulk core in combination with PFC composite”.
LBC: load-bearing capacity; FT: fracture toughness; FP: fracture pattern; FM: flexural modulus; MGF: marginal gap formation; FDP: fixed dental prosthesis.
Table 3. Quality assessment of the in vitro studies using the QUIN tool.
Table 3. Quality assessment of the in vitro studies using the QUIN tool.
First AuthorClearly Stated Aims/ObjectivesSample Size CalculationExplanation of Sampling TechniqueDetails of Control GroupExplanation of MethodologyOperator DetailsRandomizationMethod of Measurement of OutcomeOutcome Assessor DetailsBlindingStatistical AnalysisPresentation of ResultsRisk of Bias
L. Lassila [27]202220NA20NA22Medium
I. Cekic-Nagas [30]202120NA20NA22Medium
L. Lassila [36]202120NA20NA22Medium
S. Garoushi [23]202220NA20NA22Medium
L. Lassila [28]202220NA20NA22Medium
E. Mangoush [37]202120NA20NA22Medium
S. Garoushi [31]202220NA20NA22Medium
K. Nagata [32]202120NA20NA22Medium
J. Bijelic [29]202220NA20NA22Medium
S. Garoushi [35]202220NA20NA22Medium
L. Lassila [38]202122NA21NA22Low
J. Bijelic [24]202221NA21NA22Low
F. Keulemans [33]202120NA20NA22Medium
J. Bijelic [34]202220NA20NA22Medium
S. Garoushi [25]202220NA20NA22Medium
S. Garoushi [26]202222NA22NA22Low
0–2 (0: not specified; 1: inadequately specified; 2: adequately specified; NA: not applicable). Final score = (total score × 100)/(2 × number of criteria applicable). According to the final score the article is graded low, medium, or high risk of bias (>70%, low risk of bias; 50–70%, medium risk of bias; <50%, high risk of bias).
Table 4. Methodology of the mechanical tests performed in the included articles.
Table 4. Methodology of the mechanical tests performed in the included articles.
First AuthorStatical LoadingDynamical Loading
Direction of LoadingCrosshead SpeedLoading Tip Diameter
L. Lassila [27]Vertical (long axis)1 mm/min5 mmNo
I. Cekic-Nagas [30]Vertical (long axis)1 mm/min6 mmNo
L. Lassila [36]Vertical (long axis)1 mm/minNA10,000 cycles, Fmax = 500 N for 20 s, 1.2 Hz
S. Garoushi [23]45° Oblique1 mm/min2 mmNo
L. Lassila [28]Vertical (long axis)1 mm/min5 mmNo
E. Mangoush [37]45° Oblique1 mm/minNANo
S. Garoushi [31]Vertical (long axis)1 mm/min6 mmNo
K. Nagata [32]45° Oblique1 mm/minNANo
J. Bijelic [29]45° Oblique1 mm/minNANo
S. Garoushi [35]Vertical (long axis)1 mm/min3 and 6 mmNo
L. Lassila [38]45° Oblique1 mm/minNANo
J. Bijelic [24]45° Oblique1 mm/minNANo
F. Keulemans [33]Vertical (long axis) and Oblique (buccal cusp)1 mm/min6 mmNo
J. Bijelic [34]45° Oblique1 mm/minNADynamic: 10,000 cycles, 1.0 Hz
S. Garoushi [25]Vertical (long axis)1 mm/min5 mmNo
S. Garoushi [26]Vertical (long axis)1 mm/min5 mm500,000 cycles, Fmax = 150 N, 1.5 Hz
NA: not available
Table 5. Load-bearing capacity test results.
Table 5. Load-bearing capacity test results.
First AuthorLBC SFRC GroupLBC Control GroupSignificance
L. Lassila [27]3866 ± 263 N1750–2250 Np < 0.05
I. Cekic-Nagas [30]CAD-CAM FRC: 896.1 ± 30.3 N476.9 ± 20.3 Np < 0.05
L. Lassila [36]Paste SFRC: 476.9 ± 20.3 N-
S. Garoushi [23]Before aging SFRC CAD-CAM: 2624 ± 463 N1427 ± 409 Np < 0.05
L. Lassila [28]Before aging SFRC Flow: 2521 ± 371 Np < 0.05
E. Mangoush [37]After aging SFRC CAD-CAM: 2775 ± 297 N1599 ± 397 Np < 0.05
S. Garoushi [31]After aging SFRC Flow: 2404 ± 357 Np < 0.05
K. Nagata [32]349 N173 Np < 0.05
J. Bijelic [29]3990 ± 330 N1098 ± 179 Np < 0.05
S. Garoushi [35]SFRC CAD-CAM: 1650 ± 230 N850–950 Np < 0.05
L. Lassila [38]SFRC Flow: 1310 ± 397 Np < 0.05
J. Bijelic [24]2171 N1482 Np < 0.05
F. Keulemans [33]Paste SFRC: 1145 ± 89.6580 ± 40 Np < 0.05
J. Bijelic [34]SFRC CAD-CAM: 913.6 ± 86.3p < 0.05
S. Garoushi [25]Without post: 238 NWithout post: 158 N
With post: 199 NWith post: 265 N
L. Lassila [27]1733 N1081 Np < 0.05
I. Cekic-Nagas [30]~1700 N~900 Np < 0.05
L. Lassila [36]515.8 ± 241.6 N164.8 ± 95.1 Np < 0.05
S. Garoushi [23]Central fossa: 800 NCentral fossa: 702 ± 86 N
L. Lassila [28]Buccal cusp: 643 ± 8 NBuccal cusp: 403 ± 62 Np < 0.05
E. Mangoush [37]954 ± 121 N415 ± 75 Np < 0.05
S. Garoushi [31]1528 N~900 Np < 0.05
K. Nagata [32]2674 ± 465 N~1800 Np < 0.05
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jakab, A.; Palkovics, D.; T. Szabó, V.; Szabó, B.; Vincze-Bandi, E.; Braunitzer, G.; Lassila, L.; Vallittu, P.; Garoushi, S.; Fráter, M. Mechanical Performance of Extensive Restorations Made with Short Fiber-Reinforced Composites without Coverage: A Systematic Review of In Vitro Studies. Polymers 2024, 16, 590. https://doi.org/10.3390/polym16050590

AMA Style

Jakab A, Palkovics D, T. Szabó V, Szabó B, Vincze-Bandi E, Braunitzer G, Lassila L, Vallittu P, Garoushi S, Fráter M. Mechanical Performance of Extensive Restorations Made with Short Fiber-Reinforced Composites without Coverage: A Systematic Review of In Vitro Studies. Polymers. 2024; 16(5):590. https://doi.org/10.3390/polym16050590

Chicago/Turabian Style

Jakab, András, Dániel Palkovics, Veronika T. Szabó, Balázs Szabó, Eszter Vincze-Bandi, Gábor Braunitzer, Lippo Lassila, Pekka Vallittu, Sufyan Garoushi, and Márk Fráter. 2024. "Mechanical Performance of Extensive Restorations Made with Short Fiber-Reinforced Composites without Coverage: A Systematic Review of In Vitro Studies" Polymers 16, no. 5: 590. https://doi.org/10.3390/polym16050590

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop