Journal articles on the topic 'Bond strength'

To see the other types of publications on this topic, follow the link: Bond strength.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Bond strength.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

P. Nandurkar, B., and Dr A. M. Pande. "Critical studies on bond strengths of masonry units." International Journal of Engineering & Technology 7, no. 4 (September 17, 2018): 2250. http://dx.doi.org/10.14419/ijet.v7i4.15308.

Full text
Abstract:
Performance of masonry is normally attributed to compressive strength of individual units, water absorption of individual units, strength of masonry mortar and the bond between mortar and individual units. Many researches in the past have contributed towards the bond strength and relevance of compressive strength of mortar in achieving good bonds. However, the quality of bricks available in India significantly vary from region the region. Thus, a need is felt in understanding bond strength of masonry. In this paper three types of mortars(total nine combinations), two types of bricks (red clay brick and fly ash brick) are considered, tests such as compressive strength, water absorption of the bricks, compressive strength of various mortar combinations, flexure bond strength and shear bond strength are presented. Failure patterns of the masonry units are also discussed. Results of the two tests show noticeable variation in bond strengths, however the shear bond strength has significant relationship with the compressive strength of mortar. The research outcome also points towards using bricks in saturated condition for achieving adequate performance.
APA, Harvard, Vancouver, ISO, and other styles
2

Müller, M., P. Hrabě, R. Chotěborský, and D. Herák. "Evaluation of factors influencing adhesive bond strength." Research in Agricultural Engineering 52, No. 1 (February 7, 2012): 30–37. http://dx.doi.org/10.17221/4877-rae.

Full text
Abstract:
In the last ten years periods the bonding technology noted a great boom not only in manufacturing industry but in repairing industry, too. The expansion of chemical industry is the cause of this boom. In this way the use of bonding technology in industrial applications brings considerable cost savings. For the successful use of adhesives the knowledge of used adhesives and of further affecting factors is important. Respecting of this know-how is the presumption of the bonded joint successful design. The breaking of the technological procedure and the incorrect design are very often reasons of wrong joints. The paper contains theoretical in formation about the bonded joints creation and some results of laboratory tests inquiring into the reasons which affect the bonded joint strength. For tests the two-component epoxy adhesives were used.
APA, Harvard, Vancouver, ISO, and other styles
3

Murdza, Andrii, Arttu Polojärvi, Erland M. Schulson, and Carl E. Renshaw. "The flexural strength of bonded ice." Cryosphere 15, no. 6 (June 28, 2021): 2957–67. http://dx.doi.org/10.5194/tc-15-2957-2021.

Full text
Abstract:
Abstract. The flexural strength of ice surfaces bonded by freezing, termed freeze bond, was studied by performing four-point bending tests of bonded freshwater S2 columnar-grained ice samples in the laboratory. The samples were prepared by milling the surfaces of two ice pieces, wetting two of the surfaces with water of varying salinity, bringing these surfaces together, and then letting them freeze under a compressive stress of about 4 kPa. The salinity of the water used for wetting the surfaces to generate the bond varied from 0 to 35 ppt (parts per thousand). Freezing occurred in air under temperatures varying from −25 to −3 ∘C over periods that varied from 0.5 to ∼ 100 h. Results show that an increase in bond salinity or temperature leads to a decrease in bond strength. The trend for the bond strength as a function of salinity is similar to that presented in Timco and O'Brien (1994) for saline ice. No freezing occurs at −3 ∘C once the salinity of the water used to generate the bond exceeds ∼ 25 ppt. The strength of the saline ice bonds levels off (i.e., saturates) within 6–12 h of freezing; bonds formed from freshwater reach strengths that are comparable or higher than that of the parent material in less than 0.5 h.
APA, Harvard, Vancouver, ISO, and other styles
4

Scherf, Richard R. "Dentin Bond Strength." Journal of the American Dental Association 139, no. 2 (February 2008): 129. http://dx.doi.org/10.14219/jada.archive.2008.0117.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Anderson, G. P., and K. L. Devries. "Predicting Bond Strength." Journal of Adhesion 23, no. 4 (December 1987): 289–302. http://dx.doi.org/10.1080/00218468708075411.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Cardoso, Gabriela Cardoso de, Leina Nakanishi, Cristina Pereira Isolan, Patrícia dos Santos Jardim, and Rafael Ratto de Moraes. "Bond Stability of Universal Adhesives Applied To Dentin Using Etch-And-Rinse or Self-Etch Strategies." Brazilian Dental Journal 30, no. 5 (October 2019): 467–75. http://dx.doi.org/10.1590/0103-6440201902578.

Full text
Abstract:
Abstract This study evaluated the immediate and 6-month dentin bond strength of universal adhesives used in etch-and-rinse or self-etch bonding strategies. The adhesives tested were Ambar Universal, G-Bond, Single Bond Universal, Tetric N-Bond Universal, and Ybond Universal. Gold standard adhesives (Scotchbond Multipurpose Plus and Clearfil SE Bond) were controls. Microtensile dentin bond strength (n=5 teeth), pH, and C=C conversion (n=3) were evaluated. Data were analyzed at α=0.05. All adhesives showed differences in pH. Ybond had intermediately strong aggressiveness, whereas the others were ultra-mild. The C=C conversion was different in most adhesives. In the etch-and-rinse strategy, all adhesives showed similar results generally except for G-Bond, which had lower bond strength than most adhesives. G-Bond and Tetric-N-Bond showed lower bond strengths after 6 months compared with 24 h, whereas the other adhesives had stable dentin bonds. In the self-etch strategy, G-Bond had lower bond strength than most adhesives. After 6 months, Ambar was the only adhesive showing lower dentin bond strength compared with 24 h. Most adhesives had discreet drops in bond strength during aging when used in the self-etch strategy. The failure modes were also material dependent, with a general pattern of increased adhesive and/or pre-testing failures after storage. In conclusion, the bonding performance of universal adhesives to dentin is material dependent. Most adhesives had stable dentin bonds with results comparable to the gold standard materials, particularly when applied in the self-etch mode. In general, it seems the use of universal adhesives in dentin should not be preceded by phosphoric acid etching.
APA, Harvard, Vancouver, ISO, and other styles
7

Valente, Marco. "Bond Strength between Corroded Steel Rebar and Concrete." International Journal of Engineering and Technology 4, no. 5 (2012): 653–56. http://dx.doi.org/10.7763/ijet.2012.v4.454.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

De-Paula, Diego Martins, Alessandro D. Loguercio, Alessandra Reis, Natasha Marques Frota, Radamés Melo, Kumiko Yoshihara, and Victor Pinheiro Feitosa. "Micro-Raman Vibrational Identification of 10-MDP Bond to Zirconia and Shear Bond Strength Analysis." BioMed Research International 2017 (2017): 1–7. http://dx.doi.org/10.1155/2017/8756396.

Full text
Abstract:
So far, there is no report regarding the micro-Raman vibrational fingerprint of the bonds between 10-methacryloyloxy-decyl dihydrogen phosphate (10-MDP) and zirconia ceramics. Thus, the aim of this study was to identify the Raman vibrational peaks related to the bonds of 10-MDP with zirconia, as well as the influence on microshear bond strength. Micro-Raman spectroscopy was employed to assess the vibrational peak of 10-MDP binding to zirconia. Microshear bond strength of the dual-cure resin cement to zirconia with the presence of 10-MDP in composition of experimental ceramic primer and self-adhesive resin cement was also surveyed. Statistical analysis was performed by one-way ANOVA and Tukey’s test (p<0.05). Peaks at 1545 cm−1 and 1562 cm−1 were found to refer to zirconia binding with 10-MDP. The presence of 10-MDP in both experimental ceramic primer and self-adhesive resin cement improved microshear bond strength to zirconia ceramic. It can be concluded that the nondestructive method of micro-Raman spectroscopy was able to characterize chemical bonds of 10-MDP with zirconia, which improves the bond strengths of resin cement.
APA, Harvard, Vancouver, ISO, and other styles
9

Mazumdar, Paromita, Soumya Singh, and Debojyoti Das. "Method for Assessing the Bond Strength of Dental Restorative Materials — An Overview." Journal of Pierre Fauchard Academy (India Section) 35, no. 2 (October 14, 2021): 73. http://dx.doi.org/10.18311/jpfa/2021/27758.

Full text
Abstract:
<p>Bond strengths achieved while testing in laboratories are the key for selection of adhesive systems. Longevity of a restorations can be predicted to some extent based on bond strength of adhesives. There have been several discrepancies within the reported bond strengths of various materials. Bond strength of the adhesive system is affected by a large number of factors, which makes the comparison among studies difficult. Throughout the years, laboratory evaluations have been the basis for clinicians to choose the adhesive systems in their daily practice. However the validity of bond strength tests to predict clinical performance of dental adhesives is yet to be justified. The realization of an adequate and valid method for assessing bond strength is a difficult endeavor. Different types of test have been utilized to assess the strength of a bond, which has its own advantages and disadvantages. Bonding strength is the strength required to rupture a bond formed by an adhesive system and the adherent. Often, the test involves the measurement of the shear and flexural bond strength of the adhesive system. This review focuses on aspects associated to various bond strength test methods used to test the adhesion between tooth and the restorative materials and their mechanics.</p>
APA, Harvard, Vancouver, ISO, and other styles
10

Shenbagavalli, S., and Ramesh Babu Chokkalingam. "Flexural Strength of Fly ash Brick Masonry Wall with four different bond." Journal of Physics: Conference Series 2070, no. 1 (November 1, 2021): 012190. http://dx.doi.org/10.1088/1742-6596/2070/1/012190.

Full text
Abstract:
Abstract The strength of the masonry mainly depends on type of bond, types of bricks, compressive strength of the bricks and mortar used. The types of bonds play a major role in the properties of brick masonry wall. The most common types of bond used in practice are English bond, Flemish bond, Stretcher bond and Header bond. A lot of study has been performed on the load-carrying capacity of masonry walls. In this paper, effort has been taken to study the influence of different bonds on the flexural strength of the flyash brick masonry wall. For this wall of size 1m × 0.76m × 0.22m has been casted, cured for 28 days and tested in a loading frame. From the results, it was found the English bond gave higher flexural strength compared to other bonds such as Flemish, Stretcher and Header bond. The flexural strength of English bond was around 45 to 50% higher than the other bonds. The crack pattern at failure was also noted for all the masonry walls.
APA, Harvard, Vancouver, ISO, and other styles
11

Almuammar, Majed, Allen Schulman, and Fouad Salama. "Shear bond strength of six restorative materials." Journal of Clinical Pediatric Dentistry 25, no. 3 (April 1, 2001): 221–25. http://dx.doi.org/10.17796/jcpd.25.3.r8g48vn51l46421m.

Full text
Abstract:
The purpose of this study was to determine and compare the shear bond strength of a conventional glassionomer cement, a resin modified glass-ionomer, a composite resin and three compomer restorative materials. Dentin of the occlusal surfaces from sixty extracted human permanent molars were prepared for shear bond strength testing. The specimens were randomly divided into six groups of 10 each. Dentinal surfaces were treated according to the instructions of manufacturers for each material. Each restorative material was placed inside nylon cylinders 2 mm high with an internal diameter of 3 mm, which were placed perpendicular to dentin surfaces. Shear bond strengths were determined using an Universal Testing Machine at crosshead speed of 0.5 mm/min in a compression mode. Conventional glass-ionomer, Ketac-Molar aplicap showed the lowest mean shear bond strength 3.77 ± 1.76 (X ± SD MPa) and the composite resin, Heliomolar showed the highest mean shear bond strength 16.54 ± 1.65 while the mean bond strength of Fuji II LC was 9.55 ± 1.06. The shear bond strengths of compomer restorative materials were 12.83 ± 1.42, 10.64 ± 1.42 and 11.19 ± 1.19 for Compoglass, Hytac and Dyract respectively. ANOVA revealed statistically significant differences in the mean shear bond strengths of all groups (P&lt;0.001). No statistically significant difference was found between the three compomer materials (P&gt;0.5). Ketac-Molar and composite resin showed statistically significant difference (P&lt;0.0005). The mode of fracture varied between materials. It is concluded that the compomer restorative materials show higher shear bond strength than conventional glass-ionomer and resin modified glass-ionomer, but less than composite resin. The fracture mode is not related to the shear bond strengths values.
APA, Harvard, Vancouver, ISO, and other styles
12

Kim, J.-H., S.-Y. Chae, Y. Lee, G.-J. Han, and B.-H. Cho. "Effects of Multipurpose, Universal Adhesives on Resin Bonding to Zirconia Ceramic." Operative Dentistry 40, no. 1 (January 1, 2015): 55–62. http://dx.doi.org/10.2341/13-303-l.

Full text
Abstract:
SUMMARY This study evaluated the effects of single-bottle, multipurpose, universal adhesives on the bond strength of resin cement to zirconia ceramic. Polished zirconia ceramic (Cercon base) discs were randomly divided into four groups (n=40) according to the applied surface-conditioning agent: Single Bond 2, Single Bond Universal, All-Bond Universal, and Alloy Primer. Cured composite cylinders (Ø 0.8 mm × 1 mm) were cemented to the conditioned zirconia specimens with resin cement (RelyX ARC). The bonded specimens were subjected to a microshear bond-strength test after 24 hours of water storage and after 10,000 cycles of thermocycling. The surface-conditioning agent significantly influenced the bond strength (p&lt;0.05). Single Bond Universal showed the highest initial bond strength (37.7 ± 5.1 MPa), followed by All-Bond Universal (31.3 ± 5.6 MPa), Alloy Primer (26.9 ± 5.1 MPa), and Single Bond 2 (8.5 ± 4.6 MPa). Artificial aging significantly reduced the bond strengths of all the test groups (p&lt;0.05). After 10,000 cycles of thermocycling, All-Bond Universal showed the highest bond-strength value (26.9 ± 6.4 MPa). Regardless of artificial aging, Single Bond Universal and All-Bond Universal showed significantly higher bond strengths than Alloy Primer, a conventional metal primer.
APA, Harvard, Vancouver, ISO, and other styles
13

Chang, Hongtao, Hu Feng, Zeyu Guo, Aofei Guo, and Yongkang Wang. "Bond Properties of Magnesium Phosphate Cement-Based Engineered Cementitious Composite with Ordinary Concrete." Materials 15, no. 14 (July 12, 2022): 4851. http://dx.doi.org/10.3390/ma15144851.

Full text
Abstract:
A magnesium phosphate cement-based engineered cementitious composite (MPC-ECC) was developed using polyvinyl alcohol (PVA) fibers and fly ash. In this study, the bond behavior of MPC-ECC with ordinary concrete was evaluated through single and double shear bond strength tests. The effects of the water to solid mass ratio (W/S), the sand to binder mass ratio (S/B), the molar ratio of MgO to KH2PO4 (M/P), the fly ash content (F), the borax dosage (B), the volume fraction of PVA fibers (Vf), and curing age on the bond behavior of MPC-ECC with ordinary concrete were examined. The results showed that as the W/S increased, the single and double shear bond strengths were gradually reduced. As the S/B increased, the double shear bond strength increased; the single shear bond strength first decreased up to an S/B of 0.1 and then increased. With the increase of M/P, the single and double shear bond strengths increased. With the increase of F, the single shear bond strength first increased up to an F of 30% and then decreased; the double shear bond strength decreased. With the increase of B, the single and double shear bond strengths increased first and then decreased, and their strength reached its maximum at a B of 6%. The increase of Vf improved the single and double shear bond strengths. The research results can provide some technical guidance for repairing concrete structures with MPC-ECC.
APA, Harvard, Vancouver, ISO, and other styles
14

Mirhashmi, Seyyed Amir Hossein, Mohammad Sadegh Ahmad Akhundi, Saeed Mehdi Pour Ganji, Mehdi Allahdadi, Mohammad Norouzian, and Nasim Chiniforush. "Optimized Er: YAG Laser Irradiation Distance to Achieve the Strongest Bond Strength Between Orthodontic Brackets and Zirconia-Ceramics." Journal of Lasers in Medical Sciences 11, no. 3 (June 21, 2020): 287–91. http://dx.doi.org/10.34172/jlms.2020.48.

Full text
Abstract:
Introduction: In recent decades zirconium oxide has been introduced in the field of dentistry as a high-strength ceramic. Unlike its mechanical advantages, however, due to its inert chemical properties, it bonds poorly to other substrates, so improving bonding strength to an adhesive material is necessary. Methods: In this experimental study, 70 ceramic zirconia blocks were prepared and distributed randomly among 7 groups. Then the shear bond strengths were determined and the samples were examined by a scanning electron microscope (SEM). Statistical analysis was performed by one-way ANOVA and multiple Tukey comparisons. Results: One-way analysis of variance (ANOVA) showed that laser irradiation distance has a significant effect on orthodontics brackets bond strength to zirconia-ceramics. Based on the Tukey post hoc test, each group was compared with other groups and the contact mode and 2 mm distance groups showed significantly higher bond strength than other groups (P value <0.05). Conclusion: Orthodontic bracket bond strength to zirconia-ceramics will be reduced by increasing Er: YAG laser irradiation distance from samples. The highest bond strength will be achieved when laser irradiation distance is 2 mm or when the laser beam is in contact with samples.
APA, Harvard, Vancouver, ISO, and other styles
15

Iqbal, S., N. Ullah, and A. Ali. "Effect of Maximum Aggregate Size on the Bond Strength of Reinforcements in Concrete." Engineering, Technology & Applied Science Research 8, no. 3 (June 19, 2018): 2892–96. http://dx.doi.org/10.48084/etasr.1989.

Full text
Abstract:
The bond between reinforcements and concrete is the only mechanism that transfers the tensile stresses from concrete to reinforcements. Several factors including chemical adhesion, roughness and reinforcement interface and bar bearing affect the bond strength of reinforcements with concrete. This work was carried out considering another varying factor which is maximum aggregate size. Four mixes of concrete with similar compressive strengths but different maximum aggregate sizes of 25.4mm, 19.05mm, 12.7mm and 9.53mm were used with the same bar size of 16mm. Compressive strength, splitting tensile strength and bond strength for each concrete mix were studied. Test results depict a slight increase in compressive and splitting tensile strength with decrease in maximum aggregate size. The bond strength remained at the same level with decrease in maximum aggregate size except at maximum aggregate size of 9.53mm when there was a drop in bond strength, despite better compressive and splitting tensile strengths. ACI-318 and FIB-2010 codes equation for bond strength calculation work well only when the maximum aggregate size is 12.7mm and above. Therefore, maximum aggregate size is critical for bond strength when smaller size aggregates are used.
APA, Harvard, Vancouver, ISO, and other styles
16

Liu, Yi Bo, Wei Liu, Xia Huang, and Hai Peng Zhang. "Research on High Performance Vitrified Bond Diamond Wheel." Advanced Materials Research 497 (April 2012): 83–88. http://dx.doi.org/10.4028/www.scientific.net/amr.497.83.

Full text
Abstract:
Two kinds of vitrified bonds, bond L (low temperature bond) and bond G (high strength bond), were blended by ball-milling in producing diamond wheels. Proper Sintering techniques were employed by analyzing properties of the wheel through DSC testing and fracture strength. The wheels were applied in machining PDC in comparison with similar product on market. It showed that the blended bond reached its highest fracture strength when bond L amount was about 20wt~24wt%; combination of this two kinds of vitrified bonds would improve the wheel’s comprehensive properties by raising its fracture strength to 89MPa when bond L takes up 22wt%; service life of wheels prepared by this blended bond was 30~50% longer than similar market product and efficiency improved by more than 20%.
APA, Harvard, Vancouver, ISO, and other styles
17

Irie, Masao, Yukinori Maruo, Goro Nishigawa, Kumiko Yoshihara, and Takuya Matsumoto. "Flexural Strength of Resin Core Build-Up Materials: Correlation to Root Dentin Shear Bond Strength and Pull-Out Force." Polymers 12, no. 12 (December 9, 2020): 2947. http://dx.doi.org/10.3390/polym12122947.

Full text
Abstract:
The aims of this study were to investigate the effects of root dentin shear bond strength and pull-out force of resin core build-up materials on flexural strength immediately after setting, after one-day water storage, and after 20,000 thermocycles. Eight core build-up and three luting materials were investigated, using 10 specimens (n = 10) per subgroup. At three time periods—immediately after setting, after one-day water storage, and after 20,000 thermocycles, shear bond strengths to root dentin and pull-out forces were measured. Flexural strengths were measured using a 3-point bending test. For all core build-up and luting materials, the mean data of flexural strength, shear bond strength and pull-out force were the lowest immediately after setting. After one-day storage, almost all the materials yielded their highest results. A weak, but statistically significant, correlation was found between flexural strength and shear bond strength (r = 0.508, p = 0.0026, n = 33). As the pull-out force increased, the flexural strength of core build-up materials also increased (r = 0.398, p = 0.0218, n = 33). Multiple linear regression analyses were conducted using these three independent factors of flexural strength, pull-out force and root dentin shear bond strength, which showed this relationship: Flexural strength = 3.264 × Shear bond strength + 1.533 × Pull out force + 10.870, p = 0.002). For all the 11 core build-up and luting materials investigated immediately after setting, after one-day storage and after 20,000 thermocycles, their shear bond strengths to root dentin and pull-out forces were correlated to the flexural strength in core build-up materials. It was concluded that the flexural strength results of the core build-up material be used in research and quality control for the predictor of the shear bond strength to the root dentin and the retentive force of the post.
APA, Harvard, Vancouver, ISO, and other styles
18

Platt, Jeffrey A. "Decades of Bond Strength." Operative Dentistry 35, no. 2 (March 1, 2010): 137–38. http://dx.doi.org/10.2341/1559-2863-35.2.137.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Persson, Magnus, and Maud Bergman. "Metal-ceramic bond strength." Acta Odontologica Scandinavica 54, no. 3 (January 1996): 160–65. http://dx.doi.org/10.3109/00016359609003517.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Irie, Masao, Masahiro Okada, Yukinori Maruo, Goro Nishigawa, and Takuya Matsumoto. "Shear Bond Strength of Resin Luting Materials to Lithium Disilicate Ceramic: Correlation between Flexural Strength and Modulus of Elasticity." Polymers 15, no. 5 (February 23, 2023): 1128. http://dx.doi.org/10.3390/polym15051128.

Full text
Abstract:
This study investigates the effect of the curing mode (dual-cure vs. self-cure) of resin cements (four self-adhesive and seven conventional cements) on their flexural strength and flexural modulus of elasticity, alongside their shear bond strength to lithium disilicate ceramics (LDS). The study aims to determine the relationship between the bond strength and LDS, and the flexural strength and flexural modulus of elasticity of resin cements. Twelve conventional or adhesive and self-adhesive resin cements were tested. The manufacturer’s recommended pretreating agents were used where indicated. The shear bond strengths to LDS and the flexural strength and flexural modulus of elasticity of the cement were measured immediately after setting, after one day of storage in distilled water at 37 °C, and after 20,000 thermocycles (TC 20k). The relationship between the bond strength to LDS, flexural strength, and flexural modulus of elasticity of resin cements was investigated using a multiple linear regression analysis. For all resin cements, the shear bond strength, flexural strength, and flexural modulus of elasticity were lowest immediately after setting. A clear and significant difference between dual-curing and self-curing modes was observed in all resin cements immediately after setting, except for ResiCem EX. Regardless of the difference of the core-mode condition of all resin cements, flexural strengths were correlated with the LDS surface upon shear bond strengths (R2 = 0.24, n = 69, p < 0.001) and the flexural modulus of elasticity was correlated with them (R2 = 0.14, n = 69, p < 0.001). Multiple linear regression analyses revealed that the shear bond strength was 17.877 + 0.166, the flexural strength was 0.643, and the flexural modulus was (R2 = 0.51, n = 69, p < 0.001). The flexural strength or flexural modulus of elasticity may be used to predict the bond strength of resin cements to LDS.
APA, Harvard, Vancouver, ISO, and other styles
21

Talic, Yousef Fouad. "Method for Immediate Measurement of In Vitro Bond Strength of Bonded Direct Esthetic Restorations." Journal of Contemporary Dental Practice 4, no. 3 (2003): 11–23. http://dx.doi.org/10.5005/jcdp-4-3-11.

Full text
Abstract:
Abstract There are many different ways to measure the bond strength of direct esthetic restorations to various dental substrates. Unfortunately, most methods cannot measure bond strengths immediately after a restoration has been placed. This lack of clinically-relevant information seriously affects the clinician's ability to select and use various bonding agents and procedures. The aim of this article is to provide a very detailed method for immediate measurement of in vitro bond strengths of direct bonded esthetic restorations. It focuses on the steps that should be taken to select and prepare various tooth substrates for bond strength testing, the steps to “restore” various tooth substrates, and to measure the immediate in vitro bond strength. A fundamental understanding of a standardized testing protocol should provide clinicians with a clearer appreciation of bond strengths associated with various bonding procedures. Citation Talic YF. Method for Immediate Measurement of In Vitro Bond Strength of Bonded Direct Esthetic Restorations. J Contemp Dent Pract 2003 August;(4)3:011-023.
APA, Harvard, Vancouver, ISO, and other styles
22

Sul, Young-Taeg, Carina Johansson, and Tomas Albrektsson. "A novel in vivo method for quantifying the interfacial biochemical bond strength of bone implants." Journal of The Royal Society Interface 7, no. 42 (April 15, 2009): 81–90. http://dx.doi.org/10.1098/rsif.2009.0060.

Full text
Abstract:
Quantifying the in vivo interfacial biochemical bond strength of bone implants is a biological challenge. We have developed a new and novel in vivo method to identify an interfacial biochemical bond in bone implants and to measure its bonding strength. This method, named biochemical bond measurement (BBM), involves a combination of the implant devices to measure true interfacial bond strength and surface property controls, and thus enables the contributions of mechanical interlocking and biochemical bonding to be distinguished from the measured strength values. We applied the BBM method to a rabbit model, and observed great differences in bone integration between the oxygen (control group) and magnesium (test group) plasma immersion ion-implanted titanium implants (0.046 versus 0.086 MPa, n =10, p =0.005). The biochemical bond in the test implants resulted in superior interfacial behaviour of the implants to bone: (i) close contact to approximately 2 μm thin amorphous interfacial tissue, (ii) pronounced mineralization of the interfacial tissue, (iii) rapid bone healing in contact, and (iv) strong integration to bone. The BBM method can be applied to in vivo experimental models not only to validate the presence of a biochemical bond at the bone–implant interface but also to measure the relative quantity of biochemical bond strength. The present study may provide new avenues for better understanding the role of a biochemical bond involved in the integration of bone implants.
APA, Harvard, Vancouver, ISO, and other styles
23

Aksu, Muge, and Ilken Kocadereli. "Influence of Two Different Bracket Base Cleaning Procedures on Shear Bond Strength Reliability." Journal of Contemporary Dental Practice 14, no. 2 (2013): 250–54. http://dx.doi.org/10.5005/jp-journals-10024-1308.

Full text
Abstract:
ABSTRACT Purpose To search if the shear bond strengths of brackets would change after two different base-cleaning procedures such as sandblasting or carbide bur cleaning, and to determine if a previously bonded tooth surface had any effect on bond strength. Materials and methods A total of 120 new brackets were first bonded to 120 extracted premolars and then debonded and bond strength was recorded. The debonded brackets were divided into two groups and recycled either by sandblasting or tungsten-carbide bur cleaning. Sixty recycled brackets were divided into two subgroups: In each group; 30 recycled brackets were bonded to unused 30 extracted premolars. The remaining brackets were bonded to 30 previously used premolars. The brackets were debonded again and their bond strengths were remeasured. Results Bond strength of rebonded brackets after sandblasting was not significantly different from that of new brackets while the bond strength of rebonded brackets after carbide bur cleaning group significantly decreased. The previously bonded tooth surface did not affect the bond strength significantly. Clinical significance This study showed that rebonding the brackets after sandblasting supplies sufficient bond strength. Previously bonded tooth surface did not cause a decreasing effect on bond strength. However, when carbide bur cleaning procedure is chosen, the clinician should proceed cautiously. How to cite this article Aksu M, Kocadereli I. Influence of Two Different Bracket Base Cleaning Procedures on Shear Bond Strength Reliability. J Contemp Dent Pract 2013;14(2):250-254.
APA, Harvard, Vancouver, ISO, and other styles
24

Chourasia, Mukesh, Todd Cowen, Aviva Friedman-Ezra, Eden Rubanovich, and Avital Shurki. "The effect of immediate environment on bond strength of different bond types—A valence bond study." Journal of Chemical Physics 157, no. 24 (December 28, 2022): 244301. http://dx.doi.org/10.1063/5.0130020.

Full text
Abstract:
The ability to design catalysis largely depends on our understanding of the electrostatic effect of the surrounding on the bonds participating in the reaction. Here, we used a simplistic model of point charges (PCs) to determine a set of rules guiding how to construct PC-bond arrangement that can strengthen or weaken different chemical bonds. Using valence bond theory to calculate the in situ bond energies, we show that the effect of the PC mainly depends on the bond’s dipole moment irrespective of its type (being covalent or charge shift). That is, polar bonds are getting stronger or weaker depending on the sign and location of the PC, whereas non- or weakly polar bonds become stronger or weaker depending only on the location of the PC and to a smaller extent compared with polar bonds. We also show that for polar bonds, the maximal bond strengthening and weakening effect can be achieved when the PC is placed along the bond axis, as close as possible to the more and less polarizable atom/fragment, respectively. Finally, due to the stabilizing effects of polarizability, we show that, overall, it is easier to cause bond strengthening compared with bond weakening. Particularly, for polar bonds, bond strengthening is larger than bond weakening obtained by an oppositely signed PC. These rules should be useful in the future design of catalysis in, e.g., enzyme active sites.
APA, Harvard, Vancouver, ISO, and other styles
25

Tsujimoto, Akimasa, Nicholas G. Fischer, Wayne W. Barkmeier, and Mark A. Latta. "Bond Durability of Two-Step HEMA-Free Universal Adhesive." Journal of Functional Biomaterials 13, no. 3 (August 29, 2022): 134. http://dx.doi.org/10.3390/jfb13030134.

Full text
Abstract:
The purpose of this study is to compare bond durability, in terms of fatigue bond strength, of a two-step HEMA-free universal adhesive and representative adhesives in each systematic category. The adhesives used in this study were OptiBond FL, Prime&Bond NT, Clearfil SE Bond 2, G2-Bond Universal, and Scotchbond Universal Plus Adhesive. Fatigue bond strength testing and scanning electron microscopy analysis of adhesively bonded enamel and dentin interfaces were performed. For the adhesives in etch-and-rinse mode, the enamel fatigue bond strength of the G2-Bond Universal adhesive was significantly higher than those of other adhesives, and the dentin fatigue bond strength of Prime&Bond NT was significantly lower than the others. For adhesives in self-etch mode, the enamel fatigue bond strengths of Clearfil SE Bond 2 and G2-Bond Universal were significantly higher than that of the Scotchbond Universal Plus Adhesive, and the dentin fatigue bond strength of G2-Bond Universal was significantly higher than Clearfil SE Bond 2 and the Scotchbond Universal Plus Adhesive. The two-step HEMA-free universal adhesive showed higher enamel and higher or equal dentin fatigue bond strength than other selected representative adhesive systems in etch-and-rinse mode and higher or equal enamel and higher dentin fatigue bond strength than adhesive systems in self-etch mode.
APA, Harvard, Vancouver, ISO, and other styles
26

Gerlitzky, Christiane, Stefan Volz, and Peter Groche. "Brushing for High Performance Cold Pressure Welded Bonds." Key Engineering Materials 767 (April 2018): 309–15. http://dx.doi.org/10.4028/www.scientific.net/kem.767.309.

Full text
Abstract:
Joining of steel and aluminum is a commonly applied manufacturing process to obtain lightweight components. Cold pressure welding by means of direct extrusion allows gaining high bond strengths between these two materials. The contacting surfaces are usually prepared by using scratch brushing to enhance the bond strength. Most studies have shown the benefit of the brushing whereas the resulting bond strength scatters. Variations in the parameters of the brush treatment are presumed to be a major cause for the variations in strength. Within the presented work, scratch brushing parameters are adjusted to further improve the resulting bond strength. Cracking of the surfaces at low strains is a beneficial effect to enhance the bond strength. Therefore, the crack formation of the surfaces brushed under different conditions is analyzed in tensile tests. Roughness, residual stresses and microstructural changes of the aluminum surfaces resulting from brushing processes are evaluated to enhance the understanding of the cracking mechanism. Concluding, the brushing parameters are adjusted to improve bond strengths up to the material strength of the used aluminum.
APA, Harvard, Vancouver, ISO, and other styles
27

Shim, JS, YJ Park, ACF Manaloto, SW Shin, JY Lee, YJ Choi, and JJ Ryu. "Shear Bond Strength of Four Different Repair Materials Applied to Bis-acryl Resin Provisional Materials Measured 10 Minutes, One Hour, and Two Days After Bonding." Operative Dentistry 39, no. 4 (July 1, 2014): E147—E153. http://dx.doi.org/10.2341/13-196-l.

Full text
Abstract:
SUMMARY This study investigated the shear bond strength of repaired provisional restoration materials 1) to compare the bond strengths between bis-acryl resin and four different materials and 2) to investigate the effect of the amount of time elapsed after bonding on the bond strength. The self-cured bis-acryl resin (Luxatemp) was used as the base material, and four different types of resins (Luxatemp, Protemp, Z350 flowable, and Z350) were used as the repair materials. Specimens were divided into three groups depending on the point of time of shear bond strength measurement: 10 minutes, one hour, and 48 hours. Shear bond strengths were measured with a universal testing machine, and the fracture surface was examined with a video measuring system. Two-way analysis of variance revealed that the repair materials (p&lt;0.001) and the amount of time elapsed after bonding (p&lt;0.001) significantly affected the repair strength. All of the repaired materials showed increasing bond strength with longer storage time. The highest bond strength and cohesive failure were observed for bonding between Luxatemp base and Luxatemp at 48 hours after bonding.
APA, Harvard, Vancouver, ISO, and other styles
28

Coogan, Timothy J., and David O. Kazmer. "Healing simulation for bond strength prediction of FDM." Rapid Prototyping Journal 23, no. 3 (April 18, 2017): 551–61. http://dx.doi.org/10.1108/rpj-03-2016-0051.

Full text
Abstract:
Purpose The purpose of this paper is to present a diffusion-controlled healing model for predicting fused deposition modeling (FDM) bond strength between layers (z-axis strength). Design/methodology/approach Diffusion across layers of an FDM part was predicted based on a one-dimensional transient heat analysis of the interlayer interface using a temperature-dependent diffusion model determined from rheological data. Integrating the diffusion coefficient across the temperature history with respect to time provided the total diffusion used to predict the bond strength, which was compared to the measured bond strength of hollow acrylonitrile butadiene styr (ABS) boxes printed at various processing conditions. Findings The simulated bond strengths predicted the measured bond strengths with a coefficient of determination of 0.795. The total diffusion between FDM layers was shown to be a strong determinant of bond strength and can be similarly applied for other materials. Research limitations/implications Results and analysis from this paper should be used to accurately model and predict bond strength. Such models are useful for FDM part design and process control. Originality/value This paper is the first work that has predicted the amount of polymer diffusion that occurs across FDM layers during the printing process, using only rheological material properties and processing parameters.
APA, Harvard, Vancouver, ISO, and other styles
29

Sebben, Jader, Volni A. Canevese, Rodrigo Alessandretti, Gabriel K. R. Pereira, Rafael Sarkis-Onofre, Ataís Bacchi, and Aloísio O. Spazzin. "Effect of Surface Coating on Bond Strength between Etched Feldspar Ceramic and Resin-Based Luting Agents." BioMed Research International 2018 (July 24, 2018): 1–6. http://dx.doi.org/10.1155/2018/3039251.

Full text
Abstract:
This study evaluated adhesive protocols (silane, silane and unfilled resin, and universal adhesive) of bond strength between feldspar ceramic and resin-based luting agents (RBLAs). Thirty ceramic disks were embedded into acrylic resin, polished, etched, and randomly divided into 6 groups: S-RC: silane (S) and light-cured resin cement (RC) (RelyX Veneer; 3M ESPE); SB-RC: S followed by bond (B) (Clearfil SE Bond, Kuraray) and RC; UA-RC: universal adhesive (UA) (Single Bond Universal; 3M ESPE) and RC; flowable composite resin (F) was used on groups S-F, SB-F, and UA-F, and luting agent cylinders were built. The response variables (n=20) were microshear bond strength (MPa), characteristic strength (σ0, MPa), and Weibull modulus (m). The RC groups presented similar bond strengths regardless of whether or not bond was used. The S-F group with only silane application showed the highest bond strength, while the universal adhesive showed the lowest bond strength. The reliability was only affected in the UA-RC group, which was lower than the S-F group. Silane application is fundamental since the universal adhesive only decreased the bond strength between the feldspar ceramic and the RBLAs. Overall, the use of unfilled resin did not positively influence bond strength.
APA, Harvard, Vancouver, ISO, and other styles
30

Kristiawan, Stefanus, Bambang Santosa, Edy Purwanto, and Rachmad A. Caesar. "Slant shear strength of fibre reinforced polyvinyl acetate (PVA) modified mortar." MATEC Web of Conferences 195 (2018): 01016. http://dx.doi.org/10.1051/matecconf/201819501016.

Full text
Abstract:
Strengthening of reinforced concrete elements can be carried out using a variety of materials and techniques. One of such materials is textile reinforced concrete (TRC). This material consists of a matrix, usually made of mortar, and textile as reinforcement. This study aims to produce mortar that meets the characteristic of a TRC matrix with respect to an adequate bond strength. The type of mortar developed in this study was fibre reinforced polyvinyl acetate (PVA) modified mortar. The bond strength of this material to the parent concrete was tested by the slant shear method. The results indicate that the amount of PVA content affects the magnitude of the bond strength. The higher the PVA content, the higher the bond strength. The results also confirm that the relationship between the bond strengths and their corresponding compressive strengths tends to be linear.
APA, Harvard, Vancouver, ISO, and other styles
31

Cao, Rihong, Wenyu Tang, Hang Lin, and Xiang Fan. "Numerical Analysis for the Progressive Failure of Binary-Medium Interface under Shearing." Advances in Civil Engineering 2018 (2018): 1–11. http://dx.doi.org/10.1155/2018/4197172.

Full text
Abstract:
Binary-medium specimens were fabricated using the particle flow code, and the shear strength, dilatancy, and failure behavior of the binary-medium specimens with different bond strength ratios (0.25, 0.5, 0.75, and 1.0) under different normal stresses were studied. Numerical results show that the bond strength ratio and normal stresses considerably influence the shear strengths of binary-medium interface. Shear strength increases as the bond strength ratio and normal stress increase. The dilation of interfaces with high bond strength ratios is more evident than those of interfaces with lower bond strength ratios, and the curves for the high bond strength ratio exhibit remarkable fluctuations during the residual stage. At increased normal stress and bond strength ratio, the peak dilation angle shows decreasing and increasing trends successively. In this study, the specimens exhibited three kinds of failure modes. In mode II, the sawtooth experienced shear failure, but some tensile cracks appeared on the interface of the binary-medium. In mode III, no sawtooth was cut off, indicating tensile failure on the interface. At a low bond strength ratio, damage or failure is mostly concentrated in the upper part of the model. Failure parts gradually transfer to the lower part of the model when the bond strength ratio and normal stress increase. Furthermore, evident tensile cracks occur on the interface. When the bond strength ratio reaches 1.0, the failure mode of the specimen gradually transforms from sheared-off failure to chip-off failure. The number of microcracks in the specimens indicates that the lower the bond strength ratio, the more severe the damage on the specimens.
APA, Harvard, Vancouver, ISO, and other styles
32

Ateyah, Nasrien Z., and Ahmed A. Elhejazi. "Shear Bond Strengths and Microleakage of Four Types of Dentin Adhesive Materials." Journal of Contemporary Dental Practice 5, no. 1 (2004): 63–73. http://dx.doi.org/10.5005/jcdp-5-1-63.

Full text
Abstract:
Abstract The aim of this investigation was to compare the microleakage of composite resin (Z-100) and shear bond strength to bovine dentin using different types of adhesive systems (Scotch Bond Multi-Purpose, All-Bond 2, One-Step, and Perma Quick) to compare and correlate microleakage to shear bond strength. For the microleakage aspect of the study, 20 class V were prepared (bovine incisors) with 90-degree cavosurface margins and were located at the cemento-enamel junction using a template. Each dentin bonding system was applied to five cavities following the manufacturer's instructions and restored with Z-100 composite resin. After 24 hours of storage in distilled water at 37°C, the teeth were immersed in 2% basic fuchsin dye. All teeth were sectioned in a mesiodistal direction using a diamond saw, and each section was then inspected under a stereomacroscope. For the shear bond strength aspect of the study, 20 bovine incisors were centrally horizontally mounted in Teflon mold with cold cure acrylic resin. Flat labial dentin surfaces were prepared using different grit silicon carbide abrasive wheels. Five specimens were used for each of the bonding agent systems. Each specimen was bonded with restorative composite resin (Z-100) and applied to the treated dentinal surface through a split Teflon mold. All specimens were stored in distilled water at 37°C for 24 hours. The bonds were stressed using shear forces at a crosshead speed of 0.5mm/min using an Instron Universal testing machine. Findings indicate none of the systems tested in this study were free from microleakage. Scotch bond multipurpose achieved the best seal, with One-Step being second best, while All-Bond 2 and Perma Quick had the poorest seal. However, there were significant differences among the shear bond strengths of the four bonding systems tested. Scotch Bond Multi-Purpose has a higher bond strength to composite resin when compared to the other dentin adhesives. The study also concluded there is no association between microleakage and shear bond strength. Citation Ateyah AZ, Elhejazi AA. Shear Bond Strengths and Microleakage of Four Types of Dentin Adhesive Materials. J Contemp Dent Pract 2004 February;(5)1:063-073.
APA, Harvard, Vancouver, ISO, and other styles
33

Chidambaram, Aparna, Hawthorne Davis, and Subhash K. Batra. "Strength Loss in Thermally Bonded Polypropylene Fibers." International Nonwovens Journal os-9, no. 3 (September 2000): 1558925000OS—90. http://dx.doi.org/10.1177/1558925000os-900307.

Full text
Abstract:
Single fiber experiments to analyze the relationships between fiber structure, fiber properties and bonding conditions were performed with three structurally different polypropylene (PP) fibers. Bonds were formed between pairs of fibers at different temperatures and bond strengths were measured. Fiber strengths were measured before and after bonding to estimate changes caused by the bonding process. Bonded fiber strength was also compared with the strength of single fibers which experienced the same thermal conditions to assess the effect of mechanical damage from pressing fibers together with steel rolls while creating bonds. In all fiber types, significant bonding occurred simultaneously with degradation of fiber strength. The observed reduction in fiber strength was caused mainly by thermal, rather than mechanical, damage during bonding. Fibers with low birefringence skins formed strong interfiber bonds at lower temperatures with smaller losses in fiber strength.
APA, Harvard, Vancouver, ISO, and other styles
34

Tandon, Raghav, Sanjeev Maharjan, and Suraj Gautam. "Shear and tensile bond strengths of autoclaved aerated concrete (AAC) masonry with different mortar mixtures and thicknesses." Journal of Engineering Issues and Solutions 1, no. 1 (May 1, 2021): 20–31. http://dx.doi.org/10.3126/joeis.v1i1.36814.

Full text
Abstract:
Autoclaved aerated concrete (AAC) blocks are commonly used for masonry walls. In order to understand the strength of AAC masonry, it is essential to assess the tensile and shear bond strengths of the AAC block-mortar interface for various mortar combinations. This research investigates the bond strength of AAC block mortar interface made up of a) polymer modified mortar (PMM) and b) ordinary cement sand mortar of 1:4 or 1:6 ratio with thickness of 10mm, 15mm or 20mm. A thin cement slurry coating was applied on the block surface before placing the cement sand mortar in the masonry. For all types of interface, shear bond strength of masonry was studied using a triplet test, while the tensile bond strength was determined through a cross-couplet test. Among the cement sand mortar used in this study, cement sand mortar of ratio 1:4 and thickness 15mm showed the maximum shear strength of 0.13MPa with the failure of blocks as the predominant failure while the PMM had shear bond strength of 0.12MPa with the failure of blocks as the predominant failure type. However, in case of the tensile bond strength testing, PMM showed the tensile bond strength of 0.19MPa, which was highest among all the test specimens used in this study. Considering both the tensile and shear bond strengths of the AAC masonry and based on the observed failure pattern, among all the combinations used in the experiment, either PMM or cement-sand mortar of ratio 1:4 and thickness of 15mm can be chosen for the AAC masonry.
APA, Harvard, Vancouver, ISO, and other styles
35

Mendis, P., and C. French. "Bond Strength of Reinforcement in High-Strength Concrete." Advances in Structural Engineering 3, no. 3 (July 2000): 245–53. http://dx.doi.org/10.1260/1369433001502175.

Full text
Abstract:
The use of high-strength concrete is becoming popular around the world. The american code, ACI 318–95 is used in many countries to calculate the development length of deformed bars in tension. However, current design provisions of ACI 318–95 are based on empirical relationships developed from tests on normal strength concrete. The results of a series of tests on high-strength concrete, reported in the literature, from six research studies are used to review the existing recommendations in ACI 318–95 for design of splices and anchorage of reinforcement. It is shown that ACI 318–95 equations may be unconservative for some cases beyond 62 MPa (9 ksi).
APA, Harvard, Vancouver, ISO, and other styles
36

Zhang, Shao Feng, Jing Gao, Jie Mo Tian, Jun Jia, and Jiang Li. "Bond Strengths of Alumina-Glass and Zirconia-Glass Composites to Opaque-Dentine Porcelain." Key Engineering Materials 368-372 (February 2008): 1258–60. http://dx.doi.org/10.4028/www.scientific.net/kem.368-372.1258.

Full text
Abstract:
This study was conducted to evaluate shear bond strength between self-made alumina glass composites (AGC), zirconia glass composites (ZGC) and Vita alpha opaque-dentine porcelain, and also evaluate the effects of different surface treating methods on the bond strength. The AGC and ZGC specimens were treated differently and then bonded with Vita alpha opaque-dentine porcelain. The bond strength was measured by shear test and the surface treating methods included no-treatment, sandblasting and etching. The results showed that the values of bond strength for AGC groups were 30.1, 41.1 and 41.9 MPa respectively, and the bond strength of both sandblasting group and etching group were significantly higher than that of the untreated group; the bond strength of ZGC groups were 60.2, 63.6 and 35.5MPa respectively, and the value of sandblasting group was significantly higher while etching group was lower than that of untreated group. These results indicated that Vita alpha opaque-dentine porcelain can be well sintered to self-made AGC and ZGC. Sandblasting can greatly improve the bond strength for both AGC and ZGC, while the etching treating method had different effect on their bond strengths.
APA, Harvard, Vancouver, ISO, and other styles
37

Jeng, F. S., T. T. Wang, H. H. Li, and T. H. Huang. "Influences of Microscopic Factors on Macroscopic Strength and Stiffness of Inter-Layered Rocks — Revealed by a Bonded Particle Model." Journal of Mechanics 24, no. 4 (December 2008): 379–89. http://dx.doi.org/10.1017/s1727719100002501.

Full text
Abstract:
AbstractSince a conventional petrographic analysis does not allow a systematic and detailed study on how the microscopic factors affect the macroscopic behavior of inter-layered rocks, this research adopted a numerical model, the bonded particle model, to explore the micro-mechanisms associated with the strength and stiffness of inter-layered rocks. The model was first calibrated by comparing the simulations to the actual behavior until they tally with each other. Following, the microscopic factors, including the bond strength, the bond stiffness, type of bonds and friction of particles and type of bond stiffness, are varied to study their influences. As expected, the bond strength and the bond stiffness are found to have a direct and significant influence on the macroscopic uniaxial compressive strength and stiffness, respectively. Furthermore, close observations on the breaking of bonds during the loading process reveal interesting phenomena, including the transition of shear/normal bond breaking, the type of internal fracture and the factors controlling internal failure, etc. These phenomena enlighten the interpretations about the micromechanisms accounting for the macroscopic strength and stiffness of inter-layered rocks.
APA, Harvard, Vancouver, ISO, and other styles
38

Hinoura, K., M. Miyazaki, and H. Onose. "Dentin Bond Strength of Light-cured Glass-ionomer Cements." Journal of Dental Research 70, no. 12 (December 1991): 1542–44. http://dx.doi.org/10.1177/00220345910700121301.

Full text
Abstract:
The purpose of this study was to investigate the influence of surface treatments and irradiation conditions on the bond strength of light-cured glass-ionomer cements to dentin. The light-cured glass-ionomer cements used in this study were Vitrabond, XR Ionomer, and Fuji Lining LC. Three experiments were designed to study the influence of the following factors on bond strength to dentin: (1) effect of the surface treatment of the dentin, (2) effect of the irradiation time, (3) effect of an increase in the interval between mixing of the cement and irradiation. Samples were stored in water for 24 hours, after which shear bond testing was performed at a cross-head speed of 1 mm/min. For Vitrabond, the Scotchprep and Gluma 2 treatments gave the greatest shear bond strengths. For XR Ionomer and Fuji Lining LC, the Scotchprep treatment gave the greatest shear bond strengths. The bond strengths for all cements increased with prolonged irradiation time. Bond strengths decreased with a longer elapsed time between mixing and light-curing. This means that light-curing should be done soon after the cement is placed. The failure mode was found to be cohesive in the ionomer.
APA, Harvard, Vancouver, ISO, and other styles
39

Saxena, Kuldeep, Aryn Hays, Ivan Kao, and Zach Cole. "Analysis of Ultrasonic Welding of Mechanically Coupled Small Area Contacts for a Silicon Carbide Power Module." International Symposium on Microelectronics 2018, no. 1 (October 1, 2018): 000549–55. http://dx.doi.org/10.4071/2380-4505-2018.1.000549.

Full text
Abstract:
Abstract In this paper, the shear strength of small-area welded contacts was optimized. Using design of experiments (DoE) optimization techniques, copper power contacts with 4 mm2 feet were welded to both copper and nickel/gold-plated copper substrates where pressure, amplitude and deformation were the control factors. For the DoE, the setting combinations were chosen based on the concept of classical Screening Design. This design was best suited because it allowed an elementary exploration of interaction relationships for a wide range of factors. The pressure was varied from 1.4 to 1.8 bar, the amplitude from 90 to 100%, and the deformation from 0.06 to 0.1 mm. These limits were chosen based on prior welding experience with the equipment and similar parts. Each bond was optically inspected to confirm high-quality bonds. Using common shear techniques, over 200 bonds were sheared; strengths were recorded as the DoE dependent variable. The initial step found that pressure was the most significant factor. Pressure and deformation were also proven to affect strength inversely to each other. Interestingly, the planarity of the contacts had no measurable effect on the welding strength. Consistent bond strengths were achieved regardless of the bonds' place within the established bond order. For both copper-to-copper and copper-to-nickel/gold-plated bonds, the optimal settings were: pressure at 1.8 bar, amplitude at 90%, and deformation of 0.1 mm, while achieving shear strengths of 40.5 kgf and 39.1 kgf, respectively. This doubled the previous benchmark weld strengths.
APA, Harvard, Vancouver, ISO, and other styles
40

Huang, Dawei, Oriol Pons, and Albert Albareda. "Bond Strength Tests under Pure Shear and Tension between Masonry and Sprayed Mortar." Materials 13, no. 9 (May 9, 2020): 2183. http://dx.doi.org/10.3390/ma13092183.

Full text
Abstract:
Sprayed mortar or shotcrete is a construction technology that could enhance existing masonry buildings’ resilience by reinforcing low-safety load-bearing walls. Many factors affect the resistance of shotcrete-reinforced structures. One of the most important is the bond strength at the interface between the shotcrete and the reinforced wall. According to previous technical literature, bond strength usually has two evaluation criteria: shear and tensile strength. The experimental campaign described in this article focused on the bond strength between sprayed mortar and three masonry materials without the influence of normal force or constraint, as well as the roughness of these materials. The analysis of these tests focused on determining the relation between bond strength, roughness, and material strength. The analyses revealed that material strength has a more significant effect on bond strength than roughness, and bond strength is related to shrinkage of the materials. On the basis of previous theories, these researchers found that when there is no obvious influence due to normal force and constraint, the shear strength and tensile strength are different, and the shear strength is likely to be the cohesion force of the two materials. Finally, this article concludes with a novel logarithmic relationship between these strengths.
APA, Harvard, Vancouver, ISO, and other styles
41

Asmussen, E., and R. L. Bowen. "Effect of Acidic Pretreatment on Adhesion to Dentin Mediated by Gluma." Journal of Dental Research 66, no. 8 (August 1987): 1386–88. http://dx.doi.org/10.1177/00220345870660082001.

Full text
Abstract:
Tensile bond strengths between dentin and a typical restorative resin were measured after the dentin was treated with Gluma. Solutions of phosphoric, pyruvic, nitric, or oxalic acid, also containing various amino acids, were used as pretreatments. Without amino acids in the solutions, the pretreatments conferred bonds of low strength. Use of acidic solutions containing glycine or N-phenylglycine was found to give bonds of high strength to both dentin and enamel.
APA, Harvard, Vancouver, ISO, and other styles
42

Carrilho, M. R. O., R. M. Carvalho, M. F. de Goes, V. di Hipólito, S. Geraldeli, F. R. Tay, D. H. Pashley, and L. Tjäderhane. "Chlorhexidine Preserves Dentin Bond in vitro." Journal of Dental Research 86, no. 1 (January 2007): 90–94. http://dx.doi.org/10.1177/154405910708600115.

Full text
Abstract:
Loss of hybrid layer integrity compromises resin-dentin bond stability. Matrix metalloproteinases (MMPs) may be partially responsible for hybrid layer degradation. Since chlorhexidine inhibits MMPs, we hypothesized that chlorhexidine would decelerate the loss of resin-dentin bonds. Class I preparations in extracted third molars were sectioned into two halves. One half was customarily restored (etch-and-rinse adhesive/resin composite), and the other was treated with 2% chlorhexidine after being acid-etched before restoration. Specimens were stored in artificial saliva with/without protease inhibitors. Microtensile bond strengths and failure mode distribution under SEM were analyzed immediately after specimens’ preparation and 6 months later. With chlorhexidine, significantly better preservation of bond strength was observed after 6 months; protease inhibitors in the storage medium had no effect. Failure analysis showed significantly less failure in the hybrid layer with chlorhexidine, compared with controls after 6 months. In conclusion, this in vitro study suggests that chlorhexidine might be useful for the preservation of dentin bond strength.
APA, Harvard, Vancouver, ISO, and other styles
43

Shah, Arshad, Farhan, Raza, Khan, Imtiaz, Shahzadi, Qurashi, and Waseem. "Sustainable Brick Masonry Bond Design and Analysis: An Application of a Decision-Making Technique." Applied Sciences 9, no. 20 (October 14, 2019): 4313. http://dx.doi.org/10.3390/app9204313.

Full text
Abstract:
This research intends to explore the sustainable masonry bond formation and interface behaviour of brick masonry bonds with different cement mortar ratios. To test the sustainable behaviour of different brick bonds, different tests were applied to evaluate the performance of the developed five brick masonry structures with the help of four mortar ratios. Following that pattern, the methodologies of a prism triplet test, a bond wrench test, a shear bond test and strength tests for brick masonry were applied. The prism triplet test explained the bonding behaviour of mortar by producing a maximum strength (0.21 MPa) with a 1:3 mix ratio, and the minimum strength (0.095 MPa) with a 1:8 mix ratio. The bond wrench test showed a bond strength of maximum 0.0685 MPa with a mortar ratio of 1:3 and a minimum of 0.035 MPa with a mortar ratio of 1:8. The strength tests for masonry structures expressed that compressive strength (0.786 MPa) and flexural strength (0.352 MPa) were found to be at maximum level with a mortar ratio (1:3) with an English bond formation. For predictions of compressive and flexural strength, artificial neural networks (ANNs) were deployed, and successful predictions of these values along with the relationships between different properties of the material, mortar combinations and bond combinations are presented to complete the exploration of the relationship. This pattern can be helpful for the selection of sustainable brick masonry formations for housing development.
APA, Harvard, Vancouver, ISO, and other styles
44

Chiari, G., and G. Ferraris. "Bond valence VS bond length and Pauling's bond strength: The Ca – O bond." Zeitschrift für Kristallographie 191, no. 1-2 (January 1990): 39–43. http://dx.doi.org/10.1524/zkri.1990.191.1-2.39.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Sargison, Anne E., John McCabe, and Peter H. Gordon. "An Ex Vivo Study of Self-, Light-, and Dual-cured Composites for Orthodontic Bonding." British Journal of Orthodontics 22, no. 4 (November 1995): 319–23. http://dx.doi.org/10.1179/bjo.22.4.319.

Full text
Abstract:
It was postulated that using a dual-cured composite to bond orthodontic brackets could result in bond strengths comparable with those of chemically-activated materials and higher than those for light-activated materials. The shear bond strength of four composite resins used to attach mesh-backed orthodontic brackets was measured at 24 hours and following mechanical insult in the ball-mill. Analysis of variance and an SNK range test showed that at 24 hours Dual-cured Porcelite® gave a significantly higher mean bond strength than the other materials (P<0·05). However, following ball-milling the mean bond strength for Right-on® was apparently significantly higher than that of the other materials. (P<0·05). In this study, the mode of bond failure is also analysed and the use of Weibull analysis in bond strength testing is described.
APA, Harvard, Vancouver, ISO, and other styles
46

Endo, Toshiya, Rieko Ozoe, Koichi Shinkai, Makiko Aoyagi, Hiroomi Kurokawa, Yoshiroh Katoh, and Shohachi Shimooka. "Shear Bond Strength of Brackets Rebonded with a Fluoride-Releasing and -Recharging Adhesive System." Angle Orthodontist 79, no. 3 (May 1, 2009): 564–70. http://dx.doi.org/10.2319/061008-300.1.

Full text
Abstract:
Abstract Objective: To ascertain the effects of repeated bonding on the shear bond strength of orthodontic brackets bonded with a fluoride-releasing and -recharging adhesive system with a self-etching primer in comparison with two other types of adhesive system. Materials and Methods: A total of 48 premolars were collected and divided equally into three groups of 16. Each group was assigned one of three adhesive systems: Transbond XT, Transbond Plus, or a fluoride-releasing and -recharging adhesive system, Beauty Ortho Bond. Shear bond strength was measured 24 hours after bracket bonding, with the bonding/debonding procedures repeated twice after the first debonding. A universal testing machine was used to determine shear bond strengths, and bracket/adhesive failure modes were evaluated with the adhesive remnant index after each debonding. Results: At every debonding sequence, all of these three adhesive systems had a shear bond strength of 6 MPa, which is a minimum requirement for clinical use. Transbond XT and Transbond Plus had significantly higher mean shear bond strengths than did Beauty Ortho Bond at each debonding. No significant differences in mean bond strength were observed between the three debondings in each adhesive system. Bond failure at the enamel/adhesive interface occurred more frequently in Beauty Ortho Bond than in Transbond XT or Transbond Plus. Conclusions: The fluoride-releasing and -recharging adhesive system with the self-etching primer (Beauty Ortho Bond) had clinically sufficient shear bond strength in repeated bracket bonding; this finding can help orthodontists to decrease the risk of damage to enamel at debonding.
APA, Harvard, Vancouver, ISO, and other styles
47

Pujari-Palmer, Michael, Roger Giró, Philip Procter, Alicja Bojan, Gerard Insley, and Håkan Engqvist. "Factors That Determine the Adhesive Strength in a Bioinspired Bone Tissue Adhesive." ChemEngineering 4, no. 1 (March 21, 2020): 19. http://dx.doi.org/10.3390/chemengineering4010019.

Full text
Abstract:
Phosphoserine-modified cements (PMCs) are a family of wet-field tissue adhesives that bond strongly to bone and biomaterials. The present study evaluated variations in the adhesive strength using a scatter plot, failure mode, and a regression analysis of eleven factors. All single-factor, continuous-variable correlations were poor (R2 < 0.25). The linear regression model explained 31.6% of variation in adhesive strength (R2 = 0.316 p < 0.001), with bond thickness predicting an 8.5% reduction in strength per 100 μm increase. Interestingly, PMC adhesive strength was insensitive to surface roughness (Sa 1.27–2.17 μm) and the unevenness (skew) of the adhesive bond (p > 0.167, 0.171, ANOVA). Bone glued in conditions mimicking the operating theatre (e.g., the rapid fixation and minimal fixation force in fluids) produced comparable adhesive strength in laboratory conditions (2.44 vs. 1.96 MPa, p > 0.986). The failure mode correlated strongly with the adhesive strength; low strength PMCs (<1 MPa) failed cohesively, while high strength (>2 MPa) PMCs failed adhesively. Failure occurred at the interface between the amorphous surface layer and the PMC bulk. PMC bonding is sufficient for clinical application, allowing for a wide tolerance in performance conditions while maintaining a minimal bond strength of 1.5–2 MPa to cortical bone and metal surfaces.
APA, Harvard, Vancouver, ISO, and other styles
48

Poon, Clement, and Paul M. Mayer. "Electron-spin conservation and methyl-substitution effects on bonds in closed- and open-shell systems — A G3 ab initio study of small boron-containing molecules and radicals." Canadian Journal of Chemistry 80, no. 1 (January 1, 2002): 25–30. http://dx.doi.org/10.1139/v01-185.

Full text
Abstract:
High level ab initio molecular orbital theory calculations have been used to study the geometries and thermochemistry of molecules and free radicals substituted by BH2, BHCH3, and B(CH3)2. The heats of formation and RR'B—X bond strengths (RR' = H, H; H, CH3; CH3, CH3 and X = CH3, NH2, OH, F, SiH3, PH2, SH, and Cl) together with those for the open-shell systems RR'B—Y· (RR' = H, H; H, CH3; CH3, CH3 and Y = CH2, NH, O, SiH2, PH, and S) have been calculated at the G3 level of theory. The trends observed for the homolytic bond strengths in the closed-shell systems are those expected from electronegativity arguments, i.e., as the difference in electronegativity between the two atoms in the B—X bond increases, the bond strength increases. Methyl substitution on B in the closed- and open-shell species increases the ionic contribution to the bond thereby decreasing the bond strength. The lowest possible homolytic dissociation energy for the free radicals RR'BY· is lower than those of their closed-shell counterparts, yet the B—Y· bonds are shorter. This is due to the demands of spin conservation in the dissociation of the radicals favouring the formation of higher energy products.Key words: ab initio calculations, bond dissociation energy, organoboron compounds, thermochemistry.
APA, Harvard, Vancouver, ISO, and other styles
49

Cho, SD, P. Rajitrangson, BA Matis, and JA Platt. "Effect of Er,Cr:YSGG Laser, Air Abrasion, and Silane Application on Repaired Shear Bond Strength of Composites." Operative Dentistry 38, no. 3 (April 1, 2013): E58—E66. http://dx.doi.org/10.2341/11-054-l.

Full text
Abstract:
SUMMARY Aged resin composites have a limited number of carbon-carbon double bonds to adhere to a new layer of resin. Study objectives were to 1) evaluate various surface treatments on repaired shear bond strength between aged and new resin composites and 2) to assess the influence of a silane coupling agent after surface treatments. Methods Eighty disk-shape resin composite specimens were fabricated and thermocycled 5000 times prior to surface treatment. Specimens were randomly assigned to one of the three surface treatment groups (n=20): 1) air abrasion with 50-μm aluminum oxide, 2) tribochemical silica coating (CoJet), or 3) Er,Cr:YSGG (erbium, chromium: yttrium-scandium-gallium-garnet) laser or to a no-treatment control group (n=20). Specimens were etched with 35% phosphoric acid, rinsed, and dried. Each group was divided into two subgroups (n=10): A) no silanization and B) with silanization. The adhesive agent was applied and new resin composite was bonded to each conditioned surface. Shear bond strength was evaluated and data analyzed using two-way analysis of variance (ANOVA). Results Air abrasion with 50-μm aluminum oxide showed significantly higher repair bond strength than the Er,Cr:YSGG laser and control groups. Air abrasion with 50-μm aluminum oxide was not significantly different from tribochemical silica coating. Tribochemical silica coating had significantly higher repair bond strength than Er,Cr:YSGG laser and the control. Er,Cr:YSGG laser and the control did not have significantly different repair bond strengths. Silanization had no influence on repair bond strength for any of the surface treatment methods. Conclusion Air abrasion with 50-μm aluminum oxide and tribochemical silica followed by the application of bonding agent provided the highest repair shear bond strength values, suggesting that they might be adequate methods to improve the quality of repairs of resin composites.
APA, Harvard, Vancouver, ISO, and other styles
50

Zheng, Peng, Shin-ichi J. Takayama, A. Grant Mauk, and Hongbin Li. "Hydrogen Bond Strength Modulates the Mechanical Strength of Ferric-Thiolate Bonds in Rubredoxin." Journal of the American Chemical Society 134, no. 9 (February 24, 2012): 4124–31. http://dx.doi.org/10.1021/ja2078812.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography