To see the other types of publications on this topic, follow the link: Surface bond.

Journal articles on the topic 'Surface bond'

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 'Surface bond.'

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

Jones, Paul M., Huan Tang, Yiao-Tee Hsia, Xiaoping Yan, James D. Kiely, Junwei Huang, Christopher Platt, Xiaoding Ma, Michael Stirniman, and Lang Dinh. "Atomistic Frictional Properties of the C(100)2x1-H Surface." Advances in Tribology 2013 (2013): 1–11. http://dx.doi.org/10.1155/2013/850473.

Full text
Abstract:
Density functional theory- (DFT-) based ab initio calculations were used to investigate the surface-to-surface interaction and frictional behavior of two hydrogenated C(100) dimer surfaces. A monolayer of hydrogen atoms was applied to the fully relaxed C(100)2x1 surface having rows of C=C dimers with a bond length of 1.39 Å. The obtained C(100)2x1-H surfaces (C–H bond length 1.15 Å) were placed in a large vacuum space and translated toward each other. A cohesive state at a surface separation of 4.32 Å that is stabilized by approximately 0.42 eV was observed. An increase in the charge separation in the surface dimer was calculated at this separation having a 0.04 e transfer from the hydrogen atom to the carbon atom. The Mayer bond orders were calculated for the C–C and C–H bonds and were found to be 0.962 and 0.947, respectively.σC–H bonds did not change substantially from the fully separated state. A significant decrease in the electron density difference between the hydrogen atoms on opposite surfaces was seen and assigned to the effects of Pauli repulsion. The surfaces were translated relative to each other in the (100) plane, and the friction force was obtained as a function of slab spacing, which yielded a 0.157 coefficient of friction.
APA, Harvard, Vancouver, ISO, and other styles
2

Wang, Tian, Matthew H. Pelletier, Nicky Bertollo, Alan Crosky, and William R. Walsh. "Cement-Implant Interface Contamination: Possible Reason of Inferior Clinical Outcomes for Rough Surface Cemented Stems." Open Orthopaedics Journal 7, no. 1 (June 28, 2013): 250–57. http://dx.doi.org/10.2174/1874325001307010250.

Full text
Abstract:
Background: Shape-closed cemented implants rely on a stronger bond and have displayed inferior clinical outcomes when compared to force-closed designs. Implant contamination such as saline, bone marrow and blood prior to cement application has the potential to affect the cement-implant bond. The consequences of implant contamination were investigated in this study. Methods: Fifty Titanium alloy (Ti-6Al-4V) dowels were separated into ten groups based on surface roughness and contaminant, and then cemented in polyvinyl chloride tubes. Push-out testing was performed at 1mm per minute. The roughness of the dowel surface was measured before and after the testing. The dowel surface and cement mantel were analyzed using a Scanning Electron Microscopy (SEM) to determine the distribution and characteristics of any debris and contaminants on the surface. Results: Contaminants largely decreased stem-cement interfacial shear strength, especially for rough surfaces. Saline produced the greatest decrease, followed by blood. The effect of bone marrow was less pronounced and similar to that of oil. Increasing surface roughness increased the interfacial bonding strength, even with contaminants. There was a non-significant increase in mean bonding strength for smooth surfaces with bone marrow and oil contamination. SEM showed that contaminants influence the interfacial bond by different mechanisms. More debris was found on rough samples following testing. Conclusions: The results of this study underscore the importance of keeping an implant free from contamination, and suggest if contamination does occur, a saline rinse may further decrease the stability of an implant. The deleterious effects of contamination on rough surface cement bonding were considerable, and indicate that contamination at the time of surgery may, in part, contribute to inferior clinical outcomes for rough surfaced cemented stems.
APA, Harvard, Vancouver, ISO, and other styles
3

Nagasawa, Takahiro, and Koji Sueoka. "First-Principles Calculation on Initial Stage of Oxidation of Si (110)-(1 × 1) Surface." Advances in Condensed Matter Physics 2011 (2011): 1–5. http://dx.doi.org/10.1155/2011/216065.

Full text
Abstract:
The initial stage of oxidation of an Si (110)-(1 × 1) surface was analyzed by using the first-principles calculation. Two calculation cells with different surface areas were prepared. In these cells, O atoms were located at the Si–Si bonds in the first layer (A-bonds) and at the Si–Si bonds between the first and second layers (B-bonds). We found that (i) the most stable site of one O atom was the A-bond, and (ii) an O (A-bond) –Si–O (A-bond) was the most stable for two O atoms with a coverage ratio of while an O (A-bond) –Si–O (B-bond) was the most stable for . The stability of O (A-bond) –Si–Si–O (A-bond) was less than the structures obtained in (ii). The other calculations showed that the unoxidized A-bonds should be left when a coverage ratio of is close to 1. These simulations suggest that the O atoms will form clusters in the initial stage of oxidation, and the preferential oxidation will change from the A-bonds to the B-bonds up to the formation of 1 monolayer (ML) oxide. The results obtained here support the reported experimental results.
APA, Harvard, Vancouver, ISO, and other styles
4

Girish, PV, Uma Dinesh, CS Ramachandra Bhat, and Pradeep Chandra Shetty. "Comparison of Shear Bond Strength of Metal Brackets Bonded to Porcelain Surface using Different Surface Conditioning Methods: An in vitro Study." Journal of Contemporary Dental Practice 13, no. 4 (2012): 487–93. http://dx.doi.org/10.5005/jp-journals-10024-1174.

Full text
Abstract:
ABSTRACT Aim To evaluate and compare the shear bond strength of metal brackets bonded to ceramic surfaces using different conditioning methods and to assess the site of bond failure after debonding. Materials and methods A total of 70 ceramic surfaces were produced with uniform shape, size and composition. The samples were divided into 7 groups (each of 10 samples). Group 1 was the control group (untreated surface); in group 2 the surface were roughened with a diamond bur; in group 3 the surface were etched with hydrofluoric acid; in group 4 the surfaces were sandblasted; in group 5 the surfaces roughened with bur and silane applied; in group 6 the surfaces were etched with hydrofluoric acid and silane applied and in group 7 the surfaces were sandblasted and silane applied. To all the above groups, metal orthodontic brackets were bonded with light cure adhesive. The brackets were later stored in artificial saliva and incubated at 37°C (24 hours). The samples were then subjected to shear bond strength test using an Instron universal testing machine. The debonded porcelain surfaces were then studied under stereomicroscope to assess site of bond failure. Results Sandblasting the ceramic surface and silane application showed the highest bond strength. Stereomicroscope examination after debonding showed that the bond failure is at bracket-adhesive interface in four groups namely hydrofluoric acid, sandblasting, hydrofluoric acid with silane and sandblasting with silane. Conclusion Sandblasting with silane combination produced the highest shear bond strength, so it is a clinically suitable method for bonding orthodontic metal brackets onto ceramic surface. Clinical relevance Bonding orthodontic brackets to ceramic crowns of patients has been a tough task. In this study, different conditioning methods were used to treat the ceramic surfaces before bonding. The results showed that sandblasting the ceramic surface prior to application of silane produced the highest shear bond strength which is clinically suitable to reduce bond failures. How to cite this article Girish PV, Dinesh U, Bhat CSR, Shetty PC. Comparison of Shear Bond Strength of Metal Brackets Bonded to Porcelain Surface using Different Surface Conditioning Methods: An in vitro Study. J Contemp Dent Pract 2012;13(4):487-493.
APA, Harvard, Vancouver, ISO, and other styles
5

Kato, Shigenori, Kenta Chokawa, Katsumasa Kamaiya, and Kenji Shiraishi. "Theoretical Study of N Incorporation Effect during SiC Oxidation." Materials Science Forum 740-742 (January 2013): 455–58. http://dx.doi.org/10.4028/www.scientific.net/msf.740-742.455.

Full text
Abstract:
We investigated the atomistic mechanism of N incorporation during SiC oxidation by the first principles calculation. We found that N atoms play two characteristic roles in NO oxidation of SiC surface. One is that N atoms tend to form three-fold coordinated covalent bonds on a SiC(0001) surface, which assist the termination of surface dangling bonds, leading to improve the interface properties. The other is that N atoms form N-N bond like a double bond. The N2 molecule is desorbed from SiC surface, which do not disturb the oxidation process of SiC surfaces. These results indicate that N incorporation is effective to suppress defect state generation at SiO2/SiC interfaces during SiC oxidation.
APA, Harvard, Vancouver, ISO, and other styles
6

Ritchie, James P., and Steven M. Bachrach. "Bond paths and bond properties of carbon-lithium bonds." Journal of the American Chemical Society 109, no. 20 (September 1987): 5909–16. http://dx.doi.org/10.1021/ja00254a004.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Nørskov, Jens K., and Frank Abild-Pedersen. "Bond control in surface reactions." Nature 461, no. 7268 (October 2009): 1223–25. http://dx.doi.org/10.1038/4611223a.

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

Özarslan, Mustafa Mehmet, Özlem Üstün, Ulviye Sebnem Buyukkaplan, Çağatay Barutcigil, Nurullah Türker, and Kubilay Barutcigil. "Assessment the Bond Strength of Ceramic Brackets to CAD/CAM Nanoceramic Composite and Interpenetrating Network Composite after Different Surface Treatments." BioMed Research International 2018 (May 30, 2018): 1–6. http://dx.doi.org/10.1155/2018/1871598.

Full text
Abstract:
Adult orthodontics may confront problems related to the bonding performance of orthodontic brackets to new generation restorative materials used for crown or laminate restorations. The aim of the present study was to investigate the shear bond strength of ceramic brackets to two new generation CAD/CAM interpenetrating network composite and nanoceramic composite after different surface treatments. Er,Cr:YSGG Laser, hydrofluoric acid (9%), sandblasting (50 μm Al2O3), and silane were applied to the surfaces of 120 CAD/CAM specimens with 2 mm thickness and then ceramic brackets were bonded to the treated surfaces of the specimens. Bond strength was evaluated using the shear bond strength test. According to the results, CAD/CAM block types and surface treatment methods have significant effects on shear bond strength. The lowest bond strength values were found in the specimens treated with silane (3.35 ± 2.09 MPa) and highest values were found in the specimens treated with sandblast (8.92 ± 2.77 MPa). Sandblasting and hydrofluoric acid surface treatment led to the most durable bonds for the two types of CAD/CAM blocks in the present study. In conclusion, different surface treatments affect the shear bond strength of ceramic brackets to CAD/CAM interpenetrating network composite and nanoceramic composite. Among the evaluated treatments, sandblasting and hydrofluoric acid application resulted in sufficient bonding strength to ceramic brackets for both of the CAD/CAM materials.
APA, Harvard, Vancouver, ISO, and other styles
9

Filonenko, О. V., E. M. Demianenko, and V. V. Lobanov. "Quantum chemical modeling of orthophosphoric acid adsorption sites on hydrated anatase surface." Surface 12(27) (December 30, 2020): 20–35. http://dx.doi.org/10.15407/surface.2020.12.020.

Full text
Abstract:
Quantum chemical modeling of orthophosphoric acid adsorption sites on the hydrated surface of anatase was performed by the method of density functional theory (exchange-correlation functional PBE0, basis set 6-31 G(d,p)). The influence of the aqueous medium was taken into account within the framework of the continual solvent model. The work uses a cluster approach. The anatase surface is simulated by a neutral Ti(OH)4(H2O)2 cluster. The results of analysis of the geometry and energy characteristics of all the calculated complexes show that the highest interaction energy is inherent to the intermolecular complex of orthophosphoric acid and hydrated surface of anatase, where the oxygen atom of the phosphoryl group (О=Р≡) forms a hydrogen bond with a hydrogen atom of the coordinated water molecule of Ti(OH)4(H2O)2 cluster and two hydrogen atoms of the hydroxyl groups of the orthophosphoric acid molecule form two hydrogen bonds with two oxygen atoms of the titanol groups. The formation energy effect of this complex is -134.0 kJ/mol. The formation energy effect of the complex with separated charges by the proton transfer from the molecule H3PO4 to the Ti(OH)4(H2O)2 cluster with the formation of dihydrogen phosphate anion and the protonated form of the titanol group (º) is -131.1 kJ/mol, so indicating less thermodynamic probability of such intermolecular interaction. The smallest thermodynamic probability (-123.9 kJ/mol) of complexation between orthophosphoric acid and hydrated anatase surface where a water molecule moves from the coordination sphere of the titanium atom. The calculation results indicate a possible adsorption of the H3PO4 molecule in an aqueous solution on the hydrated anatase surface. Taking into account the effect of the solvent within the polarization continuum insignificantly changes the adsorption energy, which is -44.5 kJ/mol; for vacuum conditions this value is -49.0 kJ/mol.
APA, Harvard, Vancouver, ISO, and other styles
10

Fan, Bing Bing, Hai Long Wang, Li Guan, De Liang Chen, and Rui Zhang. "Ab Initio Study of Water Clusters Adsorption on Graphite Surface." Advanced Materials Research 105-106 (April 2010): 499–501. http://dx.doi.org/10.4028/www.scientific.net/amr.105-106.499.

Full text
Abstract:
Using the density functional theory method, we have characterized the geometrical structures and adsorption energy of water clusters adsorption on graphite surface. When one water molecule inter- acts with graphite surface, one of the H-O bonds formed hydrogen-bond with carbon atom in graphite sheet; in the two water molecules structure, the linear dimmer nearly parallel to the graphite surface, and also formed the hydrogen-bond; when the number of water molecules increased to six, all the H-O bond that point to the graphite surface has formed Hydrogen-bond with it. The binding energy of the water clusters with a graphite surface depends only on the number of water molecules that form hydrogen bond.
APA, Harvard, Vancouver, ISO, and other styles
11

Kwon, Taek-Ka, Jung-Yoo Choi, Jae-Il Park, and In-Sung Luke Yeo. "A Clue to the Existence of Bonding between Bone and Implant Surface: An In Vivo Study." Materials 12, no. 7 (April 11, 2019): 1187. http://dx.doi.org/10.3390/ma12071187.

Full text
Abstract:
We evaluated the shear bond strength of bone–implant contact, or osseointegration, in the rabbit tibia model, and compared the strength between grades 2 and 4 of commercially pure titanium (cp-Ti). A total of 13 grades 2 and 4 cp-Ti implants were used, which had an identical cylinder shape and surface topography. Field emission scanning electron microscopy, X-ray photoelectron spectroscopy, and confocal laser microscopy were used for surface analysis. Four grades 2 and 4 cp-Ti implants were inserted into the rabbit tibiae with complete randomization. After six weeks of healing, the experimental animals were sacrificed and the implants were removed en bloc with the surrounding bone. The bone–implant interfaces were three-dimensionally imaged with micro-computed tomography. Using these images, the bone–implant contact area was measured. Counterclockwise rotation force was applied to the implants for the measurement of removal torque values. Shear bond strength was calculated from the measured bone–implant contact and removal torque data. The t-tests were used to compare the outcome measures between the groups, and statistical significance was evaluated at the 0.05 level. Surface analysis showed that grades 2 and 4 cp-Ti implants have similar topographic features. We found no significant difference in the three-dimensional bone–implant contact area between these two implants. However, grade 2 cp-Ti implants had a higher shear bond strength than grade 4 cp-Ti implants (p = 0.032). The surfaces of the grade 2 cp-Ti implants were similar to those of the grade 4 implants in terms of physical characteristics and the quantitative amount of attachment to the bone, whereas the grade 2 surfaces were stronger than the grade 4 surfaces in the bone–surface interaction, indicating osseointegration quality.
APA, Harvard, Vancouver, ISO, and other styles
12

Sauli, Z., V. Retnasamy, W. M. W. Norhaimi, J. Adnan, and M. Palianysamy. "Wire Bond Shear Test Simulation on Hemispherical Surface Bond Pad." Advanced Materials Research 622-623 (December 2012): 643–46. http://dx.doi.org/10.4028/www.scientific.net/amr.622-623.643.

Full text
Abstract:
Wire bonding process is an interconnection method adopted in the semiconductor packaging manufactory. One of the method used to assess the reliability and bond strength of the bonded wires are wire bond shear test .In this study, simulation on wire bond shear test is done to evaluate the stress response of the bonded wire when sheared on a hemispherical surface bond pad. The contrast between three types of wire material:gold(Au), aluminum(Al) and copper(Cu) were carry out to examine the effects of wire material on the stress response of bonded wire during wire bond shear test. The simulation results showed that copper wire bond induces highest stress and gold wire exhibits the least stress response.
APA, Harvard, Vancouver, ISO, and other styles
13

Sauli, Zaliman, Vithyacharan Retnasamy, Rajendaran Vairavan, Nazuhusna Khalid, and Nooraihan Abdullah. "Wire Bond Shear Test Simulation on Flat Surface Bond Pad." Procedia - Social and Behavioral Sciences 129 (May 2014): 328–33. http://dx.doi.org/10.1016/j.sbspro.2014.03.684.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Satyanarayan and K. N. Prabhu. "Solder Joint Reliability of Sn-Cu and Sn-Ag-Cu Lead-Free Solder Alloys Solidified on Copper Substrates with Different Surface Roughnesses." Materials Science Forum 830-831 (September 2015): 265–69. http://dx.doi.org/10.4028/www.scientific.net/msf.830-831.265.

Full text
Abstract:
In the present work, the bond strength of Sn-0.7Cu, Sn-0.3Ag-0.7Cu, Sn-2.5Ag-0.5Cu and Sn-3Ag-0.5Cu lead free solders solidified on Cu substrates was experimentally determined. The bond shear test was used to assess the integrity of Sn–Cu and Sn–Ag–Cu lead-free solder alloy drops solidified on smooth and rough Cu substrate surfaces. The increase in the surface roughness of Cu substrates improved the wettability of solders. The wettability was not affected by the Ag content of solders. Solder bonds on smooth surfaces yielded higher shear strength compared to rough surfaces. Fractured surfaces revealed the occurrence of ductile mode of failure on smooth Cu surfaces and a transition ridge on rough Cu surfaces. Though rough Cu substrate improved the wettability of solder alloys, solder bonds were sheared at a lower force leading to decreased shear energy density compared to the smooth Cu surface. A smooth surface finish and the presence of minor amounts of Ag in the alloy improved the integrity of the solder joint. Smoother surface is preferable as it favors failure in the solder matrix.
APA, Harvard, Vancouver, ISO, and other styles
15

Holmberg, Kenneth, Anssi Laukkanen, Helena Ronkainen, and Kim Wallin. "Surface stresses in coated steel surfaces—influence of a bond layer on surface fracture." Tribology International 42, no. 1 (January 2009): 137–48. http://dx.doi.org/10.1016/j.triboint.2008.04.013.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Koh, S. K., Y.-B. Son, J.-S. Gam, K.-S. Han, W. K. Choi, and H.-J. Jung. "Formation of New Surface Layers on Ceramics by Ion Assisted Reaction." Journal of Materials Research 13, no. 9 (September 1998): 2560–64. http://dx.doi.org/10.1557/jmr.1998.0357.

Full text
Abstract:
Ar+ ions with 1 keV energy were irradiated on aluminum nitride in an O2 environment and on aluminum oxide in a N2 environment. AlON on AlN and AlN on Al2O3 are formed by the Ar1 irradiation in O2 gas and N2 gas environments, respectively, and the formation of new surface layers is confirmed on the basis of Al2p near core levels and O1s, N1s core levels XPS depth profile analysis. Cu(1000 Å) films were deposited by ion-beam sputtering on Ar+ irradiated/unirradiated AlN surfaces, and the change of the adhesion strength was investigated by a scratch test. Cu films deposited on the irradiated AlN under an O2 environment showed higher bond strength than that on the unirradiated AlN. The improvement of bond strength of Cu films on the AlN surface resulted from the interface bonds between Cu and the surface layers. The bending strength of polycrystalline Al2O3 irradiated by Ar+ ions in N2 environment was also increased and the formation of nitride layer on the alumina was confirmed. A possible new surface layer formation mechanism on ceramics by the ion assisted reaction has been discussed in terms of surface analysis, chemical bond, and mechanical strength.
APA, Harvard, Vancouver, ISO, and other styles
17

Gao, Hong-Ying, Philipp Alexander Held, Saeed Amirjalayer, Lacheng Liu, Alexander Timmer, Birgitta Schirmer, Oscar Díaz Arado, et al. "Intermolecular On-Surface σ-Bond Metathesis." Journal of the American Chemical Society 139, no. 20 (May 10, 2017): 7012–19. http://dx.doi.org/10.1021/jacs.7b02430.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Korneta, W., and Z. Pytel. "The Bond Percolation near the Surface." physica status solidi (b) 154, no. 1 (July 1, 1989): K1—K4. http://dx.doi.org/10.1002/pssb.2221540143.

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

Xu, Yao, Ramachandran Gnanasekaran, and David M. Leitner. "Analysis of Water and Hydrogen Bond Dynamics at the Surface of an Antifreeze Protein." Journal of Atomic, Molecular, and Optical Physics 2012 (May 14, 2012): 1–6. http://dx.doi.org/10.1155/2012/125071.

Full text
Abstract:
We examine dynamics of water molecules and hydrogen bonds at the water-protein interface of the wild-type antifreeze protein from spruce budworm Choristoneura fumiferana and a mutant that is not antifreeze active by all-atom molecular dynamics simulations. Water dynamics in the hydration layer around the protein is analyzed by calculation of velocity autocorrelation functions and their power spectra, and hydrogen bond time correlation functions are calculated for hydrogen bonds between water molecules and the protein. Both water and hydrogen bond dynamics from subpicosecond to hundred picosecond time scales are sensitive to location on the protein surface and appear correlated with protein function. In particular, hydrogen bond lifetimes are longest for water molecules hydrogen bonded to the ice-binding plane of the wild type, whereas hydrogen bond lifetimes between water and protein atoms on all three planes are similar for the mutant.
APA, Harvard, Vancouver, ISO, and other styles
20

Wardell, James L., Mukesh M. Jotani, and Edward R. T. Tiekink. "Hydrazinium 2-amino-4-nitrobenzoate dihydrate: crystal structure and Hirshfeld surface analysis." Acta Crystallographica Section E Crystallographic Communications 73, no. 4 (March 24, 2017): 579–85. http://dx.doi.org/10.1107/s2056989017004352.

Full text
Abstract:
In the anion of the title salt hydrate, H5N2+·C7H5N2O4−·2H2O, the carboxylate and nitro groups lie out of the plane of the benzene ring to which they are bound [dihedral angles = 18.80 (10) and 8.04 (9)°, respectively], and as these groups are conrotatory, the dihedral angle between them is 26.73 (15)°. An intramolecular amino-N—H...O(carboxylate) hydrogen bond is noted. The main feature of the crystal packing is the formation of a supramolecular chain along thebaxis, with a zigzag topology, sustained by charge-assisted water-O—H...O(carboxylate) hydrogen bonds and comprising alternating twelve-membered {...OCO...HOH}2and eight-membered {...O...HOH}2synthons. Each ammonium-N—H atom forms a charge-assisted hydrogen bond to a water molecule and, in addition, one of these forms a hydrogen bond with a nitro-O atom. The amine-N—H atoms form hydrogen bonds to carboxylate-O and water-O atoms, and the amine N atom accepts a hydrogen bond from an amino-H atom. The hydrogen bonds lead to a three-dimensional architecture. An analysis of the Hirshfeld surface highlights the major contribution of O...H/H...O hydrogen bonding to the overall surface,i.e. 46.8%, compared with H...H contacts (32.4%).
APA, Harvard, Vancouver, ISO, and other styles
21

Lin, Xiao Shan, and Yi Xia Zhang. "Finite Element Analysis of FRP-Reinforced Concrete Beams with Bond-Slip Effect." Applied Mechanics and Materials 553 (May 2014): 661–66. http://dx.doi.org/10.4028/www.scientific.net/amm.553.661.

Full text
Abstract:
In this paper, several FRP-reinforced concrete beams are modelled using a recently developed one-dimensional two-node layered composite beam element which accounts for the bond-slip between reinforcing bars and the surrounding concrete. Effect of different surfaces of FRP reinforcing bars on the structural response of FRP-reinforced concrete beams with bond-slip effect is also investigated. It is found that the type of rebar surface has a significant influence on bond strength and structural behaviour, and that the grain-covered surface provides the best bond between the FRP rebars and concrete, the smooth surface the worst, and FRP rebars with ribbed and braided surfaces perform similarly. Keywords: Bond-slip; Composite beam element; FRP-reinforced concrete beam; Nonlinear finite element analysis
APA, Harvard, Vancouver, ISO, and other styles
22

Sauli, Z., V. Retnasamy, S. Taniselass, A. H. M. Shapri, and R. Vairavan. "Wire Bond Shear Test Simulation on Sharp Groove Surface Bond Pad." Advanced Materials Research 622-623 (December 2012): 647–51. http://dx.doi.org/10.4028/www.scientific.net/amr.622-623.647.

Full text
Abstract:
Wire bonding process is first level interconnection technology used in the semiconductor packaging industry. The wire bond shear tests are used in the industry to examine the bond strength and reliability of the bonded wires. Hence, in this study thesimulation on wire bond shear test is performed on a sharp groove surface bond pad. ANSYS ver 11 was used to perform the simulation. The stress response of the bonded wires are investigated.The effects of three wire materials gold(Au), aluminum(Al) and copper(Cu) on the stress response during shear test were examined. The simulation results showed that copper wire bond induces highest stress and gold wire exhibits the least stress response.
APA, Harvard, Vancouver, ISO, and other styles
23

Ghaffari, Alireza, and Amirreza Ghaffari. "Retrofitting Bond of Damaged Concrete Surface in Harbour Structures." Advanced Materials Research 684 (April 2013): 177–81. http://dx.doi.org/10.4028/www.scientific.net/amr.684.177.

Full text
Abstract:
An experimental research of the bonds amongst retrofitted materials on concrete structures under seashore climate has conducted. The strength of a bond between repair materials and concrete substratum has been assessed on the base of slant shear test experiment. In this research 52 samples has prepared and used for experiment .The fresh concrete with fine aggregate mix only (less than 9mm particle size without coarse aggregate) and 14% silica fume by weight of cement (normally 8 to 10% but 14% because of high amount of fine aggregate ) added to the mix which improve the properties of concrete such as bond strength as well as compressive strength and reduces permeability of sea water in corrosion or deterioration of steel bars by protecting reinforcing steel from erosion (pivotal aim of research) and even reduces abrasion resistance .Therefore silica fume was rolled an essential repairer materials on seashore structures. The retrofitted specimens were cured on water pool and kept on the seashore simulation climate upto required curing times . The strength of the improved samples has studied in three aspects as bond strength according surface roughness ,curing periods and concrete additives like silica fumes and fibres .
APA, Harvard, Vancouver, ISO, and other styles
24

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
25

Lapinska, Barbara, Jacek Rogowski, Joanna Nowak, Joseph Nissan, Jerzy Sokolowski, and Monika Lukomska-Szymanska. "Effect of Surface Cleaning Regimen on Glass Ceramic Bond Strength." Molecules 24, no. 3 (January 22, 2019): 389. http://dx.doi.org/10.3390/molecules24030389.

Full text
Abstract:
This study investigated the effect of saliva contamination on chemical changes of ceramic surface as well as the influence of saliva cleaning methods on ceramic-resin bond strength. Saliva was used to contaminate leucite (LGC) and lithium disilicate (LDGC) glass ceramic surfaces. The following cleaning methods were tested: water spray, cleaning with orthophosphoric acid, universal cleaning paste, ultrasonic cleaning with water, re-etching with hydrofluoric acid. Non-contaminated ceramic sample served as control. Chemical analysis of ceramic surfaces was performed using time-of-flight secondary ion mass spectrometry (TOF-SIMS). Shear bond strength (SBS) of ceramics to resin material was tested after 24-hour water storage and after thermocycling. The most effective cleaning method of saliva-contaminated ceramic surface was cleaning LGC surface with orthophosphoric acid or re-etching the LDGC surface with hydrofluoric acid. The application of the following methods resulted in obtaining reliable bond strength.
APA, Harvard, Vancouver, ISO, and other styles
26

YAMAMOTO, KEIZO, HIROYUKI YOSHINAGA, and SASUKE MIYAZIMA. "SITE AND BOND PERCOLATION PROBLEM FOR CONSTRUCTION OF MACROSCOPIC SURFACE IN A CUBIC LATTICE." Fractals 17, no. 01 (March 2009): 131–35. http://dx.doi.org/10.1142/s0218348x09004235.

Full text
Abstract:
When we distribute particles and bonds onto a cubic lattice, we assume that a plaquette is formed, if four corner sites of a square are occupied or four bonds of a square are occupied. When these plaquettes are produced in a cubic lattice and occupy a certain number of square surfaces in the cubic lattice, then a macroscopic surface is formed following a chain construction. This is the second phase transition added to usual site and bond percolation problems where a spreading chain is formed.
APA, Harvard, Vancouver, ISO, and other styles
27

Machmud, Edy, Moh Dharmautama, and Cencen Tjandi Yanto. "Pengaruh ukuran mesh permukaan logam terhadap kuat rekat geser semen resin adesif pada restorasi gigitiruan jembatan adesif Effect of metal surface mesh size to shear bond strength of adhesive resin cement to denture adhesive bridge." Journal of Dentomaxillofacial Science 11, no. 3 (October 30, 2012): 149. http://dx.doi.org/10.15562/jdmfs.v11i3.329.

Full text
Abstract:
To overcome the failure of cement adhesive bond between tooth and metal surfaces, a variety of alternative treatmentsis applied to the metal and tooth surfaces. In order to overcome this disadvantage, this study deals with the treatmenton the metal surface in mesh form. This research examined different mesh sizes of 40, 60 and 80 mesh to the metalsurface treatment in shear bond strength of adhesive resin cement. 40 pairs of upper central incisors samples wereprepared in palatal surfaces with a thickness of 0.3 mm and made a pattern of blue wax. The samples were divided intofour groups; three mesh treatment groups (40, 60 and 80) and a control group (no mesh). In the treatment groups, amesh was positioned to the opposing surface of wax pattern to the tooth preparation surface, was casted with nickelchromiumalloy, attached to the palatal surface of the teeth with adhesive resin cement, immersed in artificial salivafor 24 hours at 37°C. Moreover, shear bond strength test and residual resin cement on the metal surface were tested inLaboratory of the Department of Metallurgical Engineering. Data was analyzed using one-way ANOVA (p=0.05).Shear bond strength adherence of mesh plates 60 was found greater than 40, 80 and control group. Stereomicroscopeinspection of adhesive resin cement remaining on mesh 60 was similar than 40, but more than mesh 80 or controlgroup. It was concluded shear bond strength of adhesive resin cement to metal surfaces depends on the size of meshused, and shear bond strength of adhesive resin cement that still intact on the mesh surface.
APA, Harvard, Vancouver, ISO, and other styles
28

Mitchell, K. A. R. "Analysis of surface bond lengths reported for chemisorption on metal surfaces." Surface Science 149, no. 1 (January 1985): 93–104. http://dx.doi.org/10.1016/s0039-6028(85)80015-7.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Mitchell, K. A. R. "Analysis of surface bond lengths reported for chemisorption on metal surfaces." Surface Science Letters 149, no. 1 (January 1985): A4. http://dx.doi.org/10.1016/0167-2584(85)90796-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Levine, Lee. "Wire Bonding: The Ultrasonic Bonding Mechanism." International Symposium on Microelectronics 2020, no. 1 (September 1, 2020): 000230–34. http://dx.doi.org/10.4071/2380-4505-2020.1.000230.

Full text
Abstract:
Abstract Wire bonding is a welding process. During both ball and wedge bonding, wire and bond pad are massively deformed between the bond tool and the anvil of the bond pad or substrate. The dominant variables affecting deformation are ultrasonic energy, temperature, bond force and bond time. Deformation exposes new surface material that is clean and has not been exposed to atmospheric contamination and oxidation. As the new wire and bond pad surfaces mix, they form diffusion couples that grow and transform into the intermetallic weld nugget. The initial mixing is not at equilibrium in that it does not initially form the compounds described by the equilibrium phase diagram, but temperature and time very quickly allows diffusion to relax the initial mixture into the equilibrium phase diagram compounds. This paper will discuss the mechanisms behind the formation of ball and wedge bonds.
APA, Harvard, Vancouver, ISO, and other styles
31

Mitchell, K. A. R., S. A. Schlatter, and R. N. S. Sodhi. "Further analysis of surface bond lengths measured for chemisorption on metal surfaces." Canadian Journal of Chemistry 64, no. 7 (July 1, 1986): 1385–89. http://dx.doi.org/10.1139/v86-237.

Full text
Abstract:
This paper compares bond lengths deduced from the methods of surface crystallography with predictions from the Pauling–Schomaker–Stevenson approach and from a new alternative approach suggested by recent work of Brown and Altermatt. Examples considered are specifically for X—M surface bond lengths where atoms X from groups 16 or 17 are adsorbed on well-defined surfaces of a metal M. The alternative approach introduced here is parametrised with reference to structural data from solid compounds of formula MX. The two predictive approaches considered, when used together, appear to be quite adequate for guiding choices of trial model structures to be included in surface crystallographic analyses with low-energy electron diffraction (LEED); also they seem reasonable for checking the general reliability (or otherwise) of surface bond length data. Two further features introduced by this work are (i) evidence that the Cl—Ag distance reported by LEED for Cl adsorbed on the Ag(100) surface is broadly consistent with the structure of solid AgCl; (ii) evidence for S adsorbed on the Fe(110) surface that these analyses can guide investigations of lateral relaxations of surface metal atoms. As more reliable structural data become available, extensions of these analyses should help to identify the finer details in X—M bond lengths which result from the special coordination arrangements occurring at surfaces.
APA, Harvard, Vancouver, ISO, and other styles
32

Ding, Wanyu, Yuanyuan Guo, Dongying Ju, Susumu Sato, and Teruo Tsunoda. "The effect of CH4/H2 ratio on the surface properties of HDPE treated by CHx ion beam bombardment." Modern Physics Letters B 30, no. 17 (June 30, 2016): 1650214. http://dx.doi.org/10.1142/s0217984916502146.

Full text
Abstract:
The surface of high density polyethylene (HDPE) substrate was bombarded by the CH[Formula: see text] group ion beam, which was generated by the mixture of CH4/H2. Varying the CH4/H2 ratio, HDPE surfaces with different chemical bond structures and properties were obtained. Raman and XPS results show that [Formula: see text] and [Formula: see text] bond structures are formed at HDPE surface bombarded by CH[Formula: see text] group ions. The [Formula: see text] bond fraction at bombarded HDPE surface depends on the H2 ratio in CH4/H2 mixture, because the H ion/atom/molecule can improve the growth of [Formula: see text] bond structure. For HDPE surface bombarded by CH4/H2 = 50/50, [Formula: see text] bond fraction reaches the maximum of 30.5%, the surface roughness decreases to 17.04 nm, and the static contact angle of polar H2O molecule increased to 140.2[Formula: see text].
APA, Harvard, Vancouver, ISO, and other styles
33

Burrer, Phoebe, Amanda Costermani, Matej Par, Thomas Attin, and Tobias T. Tauböck. "Effect of Varying Working Distances between Sandblasting Device and Composite Substrate Surface on the Repair Bond Strength." Materials 14, no. 7 (March 26, 2021): 1621. http://dx.doi.org/10.3390/ma14071621.

Full text
Abstract:
This study investigates the effect of defined working distances between the tip of a sandblasting device and a resin composite surface on the composite–composite repair bond strength. Resin composite specimens (Ceram.x Spectra ST (HV); Dentsply Sirona, Konstanz, Germany) were aged by thermal cycling (5000 cycles, 5–55 °C) and one week of water storage. Mechanical surface conditioning of the substrate surfaces was performed by sandblasting with aluminum oxide particles (50 µm, 3 bar, 10 s) from varying working distances of 1, 5, 10, and 15 mm. Specimens were then silanized and restored by application of an adhesive system and repair composite material (Ceram.x Spectra ST (HV)). In the negative control group, no mechanical surface pretreatment or silanization was performed. Directly applied inherent increments served as the positive control group (n = 8). After thermal cycling of all groups, microtensile repair bond strength was assessed, and surfaces were additionally characterized using scanning electron microscopy (SEM) and energy dispersive X-ray spectroscopy (EDX). The negative control group reached the significantly lowest microtensile bond strength of all groups. No significant differences in repair bond strength were observed within the groups with varying sandblasting distances. Composite surfaces sandblasted from a distance of 1 mm or 5 mm showed no difference in repair bond strength compared to the positive control group, whereas distances of 10 or 15 mm revealed significantly higher repair bond strengths than the inherent incremental bond strength (positive control group). In conclusion, all sandblasted test groups achieved similar or higher repair bond strength than the inherent incremental bond strength, indicating that irrespective of the employed working distance between the sandblasting device and the composite substrate surface, repair restorations can be successfully performed.
APA, Harvard, Vancouver, ISO, and other styles
34

Peng, Hui Yuen, Mutharasu Devarajan, Teik Toon Lee, and David Lacey. "Investigation of oxygen followed by argon plasma treatment on LED chip bond pad for wire bond application." Soldering & Surface Mount Technology 27, no. 4 (September 7, 2015): 129–36. http://dx.doi.org/10.1108/ssmt-03-2015-0006.

Full text
Abstract:
Purpose – The purpose of this paper is to investigate the efficiencies of argon (Ar), oxygen (O2) and O2 followed by Ar (O2→Ar) plasma treatments in terms of contaminant removal and wire bond interfacial adhesion improvement. The aim of this study is to resolve the “lifted ball bond” issue, which is one of the critical reliability checkpoints for light emitting diodes (LEDs) in automotive applications. Design/methodology/approach – Ar, O2 and O2→Ar plasma treatments were applied to LED chip bond pad prior to wire bonding process with different treatment durations. Various surface characterization methods and contact angle measurement were then used to characterize the surface properties of these chip bond pads. To validate the improvements of Ar, O2 and O2→Ar plasma treatments to the wire bond interfacial adhesion, the chip bond pads were wire bonded and examined with a ball shear test. Moreover, the contact resistance of the wire bond interfaces was also measured by using four-point probe electrical measurements to complement the interfacial adhesion validation. Findings – Surface characterization results show that O2→Ar plasma treatment was able to remove the contaminant while maintaining relatively low oxygen impurity content on the bond pad surface after the treatment and was more effective as compared with the O2 and Ar plasma treatments. However, O2→Ar plasma treatment also simultaneously reduced high-polarity bonds on the chip bond pad, leading to a lower surface free energy than that with the O2 plasma treatment. Ball shear test and contact resistance results showed that wire bond interfacial adhesion improvement after the O2→Ar plasma treatment is lower than that with the O2 plasma treatment, although it has the highest efficiency in surface contaminant removal. Originality/value – To resolve “lifted ball bond” issue, optimization of plasma gas composition ratios and parameters for respective Ar and O2 plasma treatments has been widely reported in many literatures; however, the O2→Ar plasma treatment is still rarely focused. Moreover, the observation that wire bond interfacial adhesion improvement after O2→Ar plasma treatment is lower than that with the O2 plasma treatment although it has the highest efficiency in surface contaminant removal also has not been reported on similar studies elsewhere.
APA, Harvard, Vancouver, ISO, and other styles
35

KURIHARA, Norihiko, Takuya YOSHIDA, Taichi ENOMOTO, and Satoru KOGA. "BOND AND SHEAR PROPERTIES OF BOND SURFACE JOINTED BY CHEMICAL ROUGHENING METHOD." Cement Science and Concrete Technology 73, no. 1 (March 31, 2020): 340–47. http://dx.doi.org/10.14250/cement.73.340.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Liu, Y., J. P. Li, E. B. Hunziker, and K. de Groot. "Incorporation of growth factors into medical devices via biomimetic coatings." Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 364, no. 1838 (November 30, 2005): 233–48. http://dx.doi.org/10.1098/rsta.2005.1685.

Full text
Abstract:
Orthopaedic and dental surgeons are fully aware of the need for implants to bond well with the surrounding living bone if long-lasting clinical success is to be achieved. For example, well-bonded hip implants have a 10 year failure rate, which is lowered fivefold if bonding is poor or absent. The techniques that are currently available to impart implant surfaces with the desired osteoconductive properties are essentially limited. To overcome the inherent difficulties, we have developed a ‘biomimetic’ coating process. By this means, implants with complex surface geometries, such as porous spinal implants, can be furnished with a bone-bonding surface. Furthermore, these coatings can be rendered osteoinductive as well as osteoconductive (by incorporating osteogenic agents). Using this facility, we have induced bone formation at an ectopic site in rats, and have accelerated osseointegration (bone bonding) at an orthotopic dental site in adult miniature pigs. Our preliminary results indicated that these osteoinductive dental implants bond with surrounding bone within one week instead of the usual three weeks. We believe that surfaces coated with biomimetic coatings into which osteogenic growth factors are incorporated hold great potential for use in clinical orthopaedics and dentistry.
APA, Harvard, Vancouver, ISO, and other styles
37

Kobayashi, Katsuyoshi, and Emiko Ishikawa. "Surface-state conduction through dangling-bond states." Surface Science 540, no. 2-3 (August 2003): 431–40. http://dx.doi.org/10.1016/s0039-6028(03)00880-x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Knapp, James K., and Thomas A. Taylor. "Waterjet roughened surface analysis and bond strength." Surface and Coatings Technology 86-87 (December 1996): 22–27. http://dx.doi.org/10.1016/s0257-8972(96)02995-7.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Kebede, Getachew G., Pavlin D. Mitev, Peter Broqvist, Jolla Kullgren, and Kersti Hermansson. "Hydrogen-Bond Relations for Surface OH Species." Journal of Physical Chemistry C 122, no. 9 (January 11, 2018): 4849–58. http://dx.doi.org/10.1021/acs.jpcc.7b10981.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Hoffmann, Regina, Clemens Barth, Adam S. Foster, Alexander L. Shluger, Hans J. Hug, Hans-Joachim Güntherodt, Risto M. Nieminen, and Michael Reichling. "Measuring Site-Specific Cluster−Surface Bond Formation." Journal of the American Chemical Society 127, no. 50 (December 2005): 17863–66. http://dx.doi.org/10.1021/ja055267i.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Held, Philipp Alexander, Harald Fuchs, and Armido Studer. "Covalent-Bond Formation via On-Surface Chemistry." Chemistry - A European Journal 23, no. 25 (January 18, 2017): 5874–92. http://dx.doi.org/10.1002/chem.201604047.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Orłowski, B. A., B. J. Kowalski, J. Bonnet, C. Hricovini, and R. Pinchaux. "Dangling bond states on HgSe(110) surface." Vacuum 45, no. 2-3 (February 1994): 199–201. http://dx.doi.org/10.1016/0042-207x(94)90170-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Ramesh, Ganesh, T. V. Padmanabhan, Padma Ariga, Shalini Joshi, S. Bhuminathan, and Vasantha Vijayaraghavan. "Surface Bond Strength in Nickel Based Alloys." Journal of Indian Prosthodontic Society 13, no. 4 (October 6, 2012): 551–54. http://dx.doi.org/10.1007/s13191-012-0186-x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Haneman, D., N. S. McAlpine, E. Busch, and C. Kaalund. "Semiconductor bond rupture phenomena and surface properties." Applied Surface Science 92 (February 1996): 484–90. http://dx.doi.org/10.1016/0169-4332(95)00282-0.

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

Jackson, David. "CO2 SPRAY CLEANING AND OSEE NON-CONTACT INSPECTION FOR WIRE BOND PAD PREPARATION." International Symposium on Microelectronics 2014, no. 1 (October 1, 2014): 000307–12. http://dx.doi.org/10.4071/isom-tp47.

Full text
Abstract:
Surface pad contamination is a major cause of poor performance for wire bonding operations. Examples of the wide range of contaminants that can degrade wire bond pull strength include, for example:Halogens and hydrocarbons: plasma etching, epoxy outgassing (dry processing), photoresist strippers, cleaning solvents.Contaminants from plating operations: thallium, brighteners, lead, iron, chromium, copper, nickel, hydrogen.Sulfur compounds: packing containers, ambient air, cardboard and paper, rubber bands.Miscellaneous organic contaminants: epoxy outgassing, photoresist, general ambient air (poor storage).Miscellaneous inorganic compounds: sodium, chromium, phosphorous, bismuth, cadmium, moisture, glass, vapor, nitride, carbon, silver, copper, tin.Human sources of contamination: skin particles, hair, sweat, spittle, mucus, cosmetics, hand lotions, facial make-up and fibers from clothing. As can be seen, there are many types of surface contaminations that may challenge a wire bonding operation, all of which must be removed to insure reliable and strong bonds. In this regard, conventional precision cleaning processes for high reliability surface pad preparation typically involve multiple steps, chemistries, and equipment to accomplish complete decontamination. Moreover, conventional cleaning methods are sometimes non-selective for the surface contaminants and substrates. For example, conventional vacuum plasma using Ar/O2 is typically used to clean bond pads. Vacuum plasmas are usually performed off-line, taking up to 30 minutes to complete, and are non-selective for the organic contamination. The entire organic substrate (i.e., PCB) is etched away during the plasma cleaning process to remove the bond pad contamination. During treatment, secondary organic surface contaminations (plasma treatment by-products from reacted substrate) are produced which can re-contaminate bonding surfaces. Advanced carbon dioxide (CO2) spray cleaning technology provides various methods for consistently preparing bond pads for critical wire bonding operations. A patented hybrid CO2 particle-plasma spray is presented in this paper that has demonstrated efficacy for selectively treating bond pad surfaces to remove a wide range of challenging surface contaminations. Moreover, a novel non-contact surface inspection technology called Optically Stimulated Electron Emission (OSEE) - developed to address surface cleaning and inspection issues that led to the 1986 Challenger Spacecraft explosion - is used to measure the effectiveness of the new CO2 surface cleaning processes. Statistically significant studies have been performed to determine the effectiveness of the selective CO2 particle-plasma surface treatment process for preparing bond pads for gold ribbon bonding operations. One such study compared and contrasted the performance of this new single-step CO2 surface treatment method with that of a conventional multi-step solvent-plasma method. The two treatment methods were used to prepare the surface of a metalized ceramic wafer that simulated bond pad surfaces and treatment areas representative of an actual high-reliability electronic board. The test results of this evaluation demonstrated that the CO2 particle-plasma surface treatment process is statistically similar to or sometimes better than a solvent-plasma hybrid cleaning process. CO2 spray cleaning was determined to be better for some types of contaminants as well – and in particular more relevant mixed-contaminant challenge tests. The CO2 cleaning process demonstrates a lower defect-per-million (DPM) level and an improved CpK. Finally, in this study OSEE surface quality analysis was performed before and after surface cleaning. OSEE analysis provided a reliable non-contact means of determining the proper level of surface pad preparation.
APA, Harvard, Vancouver, ISO, and other styles
46

Pazinatto, Flavia Bittencourt, Fabrício Luscino Alves de Castro, and Maria Teresa Atta. "Influence of Surface Inclination on the Spreading Velocity of Simplified Adhesive Systems." International Journal of Experimental Dental Science 2, no. 2 (2013): 92–97. http://dx.doi.org/10.5005/jp-journals-10029-1048.

Full text
Abstract:
ABSTRACT Aim Adhesion can be influenced by adhesives spreading. Both slow and fast spreading can be deleterious as they can respectively lead adhesive to partially cover the demineralized substrate or to accumulate on line angles of cavity preparations. Since brands of dentin bonding systems present distinct compositions and thus different behaviors its important to know how fast they spread over dental substrates. The purpose is to determine the influence of surface inclinations on the spreading velocity of simplified adhesive systems. Materials and methods Spreading velocities of adhesive systems (Adper Single Bond, Adper Single Bond Plus, Adper Prompt, Prime and Bond 2.1, Prime and Bond NT, One-Up Bond F) were measured on glass slide surfaces inclined at 45° and 90°. The spreading of each drop was observed at a 30 seconds interval. Data was recorded and the values obtained at 30 seconds were analyzed by two-way ANOVA and Student-Newman-Keuls (SNK) tests (α = 0.05). Results The type of adhesive system and the angle of inclination influenced spreading velocity (p < 0.05). Adper Single Bond Plus and One-Up Bond F exhibited the lowest spreading velocities of all materials tested (p < 0.01). Adhesives spreading were similar on surfaces inclined 45° and 90°, except for Adper Single Bond Plus which spread faster and Prime and Bond NT that spread slower on 90° angled surfaces (p < 0.01). Conclusion The materials tested showed complex spreading patterns since the spreading velocities changed only when some specific material/inclination combinations where tested. How to cite this article Pazinatto FB, de Castro FLA, Saini R, Atta MT. Influence of Surface Inclination on the Spreading Velocity of Simplified Adhesive Systems. Int J Experiment Dent Sci 2013;2(2):92-97.
APA, Harvard, Vancouver, ISO, and other styles
47

Trzcionka, Agata, Ruta Narbutaite, Alma Pranckeviciene, Rytis Maskeliūnas, Robertas Damaševičius, Gintautas Narvydas, Dawid Połap, Katarzyna Mocny-Pachońska, Marcin Wozniak, and Marta Tanasiewicz. "In Vitro Analysis of Quality of Dental Adhesive Bond Systems Applied in Various Conditions." Coatings 10, no. 9 (September 17, 2020): 891. http://dx.doi.org/10.3390/coatings10090891.

Full text
Abstract:
Introduction: There are several methods of reducing a microleakage, and one of them is choosing appropriate adhesive material. The aim of the work was the in vitro analysis of 4 bonds: 3M ESPE “Single bond”, Dentsply “Prime and Bond Active”, Coltene “One Coat 7 Universal”, and Kuraray “Clearfil Universal Bond Quick”. Material and methods: 136 healthy molar teeth were collected and randomly split into 4 groups and Vth Class cavities were prepared. Chosen adhesives were used in four groups of teeth with the same composite. Teeth were the thermocycled, sealed, covered with lacquer, and submerged in 1% methylene blue solution for 24 h. After the thermocycling, the vertices of each tooth were sealed using dental wax. Each tooth was then fully covered with lacquer. All teeth were then submerged into 1% methylene blue solution for 24 h in room temperature. In the next step they were transversely cut through a center of restoration. The Olympus BX43 microscope was used to photograph each cut tooth. With the usage of Olympus stream software, measurement of the dye’s leakage was performed. Results. The statistical analysis proved that the most effective material when applied to ideally prepared cavity surface was Dentsply “Prime and Bond Active”. The second material was 3M ESPE “Single Bond”, third—Coltene “One Coat 7 Universal” and fourth—Kuraray “Clearfil Universal Bond Quick”. The most effective material applied to a too-dry surface was Dentsply “Prime and Bond Active”, second—3M ESPE “Single Bond”, third—Coltene “One Coat 7 Universal” and fourth—Kuraray “Clearfil Universal Bond Quick”. When it comes to too damp surfaces the best results were obtained with Dentsply “Prime and Bond Active” then Coltene “One Coat 7 Universal”, 3M ESPE “Single Bond” and Kuraray “Clearfil Universal Bond Quick”. Conclusion: The level of cavity dampness influences the quality of adhesives. Better results are obtained with over-dried surfaces than over-damp, which is connected with the dilution of the material.
APA, Harvard, Vancouver, ISO, and other styles
48

de Oliveira, Marcelo Tavares, Patrícia Moreira de Freitas, Carlos de Paula Eduardo, Glaucia Maria Bovi Ambrosano, and Marcelo Giannini. "Influence of Diamond Sono-Abrasion, Air-Abrasion and Er:YAG Laser Irradiation on Bonding of Different Adhesive Systems to Dentin." European Journal of Dentistry 01, no. 03 (July 2007): 158–66. http://dx.doi.org/10.1055/s-0039-1698332.

Full text
Abstract:
ABSTRACTObjectives: Different surface treatments may affect bonding performance of adhesive systems to dentin. This study evaluated the influence of different methods of surface treatment on adhesion of bonding agents to dentin.Methods: TDentin surfaces abraded with #600-grit SiC paper were used as control. Three methods of surface treatment (sono-abrasion, air-abrasion and Er:YAG laser irradiation) were used under specific parameters. Four adhesive systems (Tyrian, Clearfil SE Bond, Unifil Bond and Single Bond) were applied to treated surfaces, according to the manufacturers’ instructions. Composite blocks were built on bonded surfaces, then restored teeth were vertically and serially sectioned to obtain bonded slices for interfacial micromorphologic analysis or to produce beam specimens for μ-TBS bond test. Data were analyzed with two-way ANOVA and Tukey test at a significance level of 5%.Results: The results indicated that the preparation of dentin with sono-abrasion or laser did not affect the bond strength, while the preparation of dentin with SiC paper and air-abrasion influenced the bond strength for some systems. A clear difference of the preparation of dentin surfaces and formation of hybrid layer and resin tags were noted.Conclusions: Bonding effectiveness of both the etch-and-rinse and the self-etch adhesives can be influenced by different methods of dentin preparation. (Eur J Dent 2007;1:158-166)
APA, Harvard, Vancouver, ISO, and other styles
49

Zhou, Li, Yuetong Qian, Kang Gan, Hong Liu, Xiuju Liu, and Deli Niu. "Effect of different surface treatments and thermocycling on shear bond strength to polyetheretherketone." High Performance Polymers 29, no. 1 (July 28, 2016): 87–93. http://dx.doi.org/10.1177/0954008316628966.

Full text
Abstract:
This study was designed to evaluate the shear bond strength of an adhesive/composite system subjected to different pretreated polyetheretherketone (PEEK) surfaces using different thermocycling conditioning methods. A total of 128 specimens were equally divided into four main groups ( n = 32/group): control (no pretreatment), air abrasion, argon plasma pretreatment, and femtosecond laser groups. The surface topographies and surface roughness were observed by atomic force microscopy after different pretreatments. The specimens were bonded with SE Bond/Clearfil AP-X™. All bonded specimens were stored in distilled water at 37°C for 24 h. Afterward, each group was divided into three subgroups ( n = 8/group) as follows: (a) stored in water for 56 h (37°C); (b) thermal aging for 5000 cycles (5°C/55°C); and (c) thermal aging for 10,000 cycles (5°C/55°C). The shear bond strengths were measured. Air abrasion, argon plasma pretreatment, and femtosecond laser significantly strengthened the bond of SE Bond/Clearfil AP-X™ to PEEK composite compared with that without additional pretreatment. In the same surface pretreatment, the shear bond strengths of specimens conditioned using water storage were higher than that using thermocycles (TCs). Additionally, the specimens with 5000 TC showed significantly higher shear bond strength than that with 10000 TC.
APA, Harvard, Vancouver, ISO, and other styles
50

Nelson, C. B., and H. Fang. "Tight binding model of induced band shift in CoO nanoparticles." Canadian Journal of Physics 98, no. 1 (January 2020): 39–44. http://dx.doi.org/10.1139/cjp-2018-0974.

Full text
Abstract:
Recently researchers have shown that water splitting occurs in 6 nm particles of Co-O, an effect not seen in the pure crystal, indicating that the band shift in these particles is probably due in part to surface modification. Others have shown that band edges in metal oxide crystals can be shifted up by induced dipoles on the surface. These are created by attaching polar molecules or passive ligands. Here we present a tight binding model that predicts that the observed band shift in Co-O is caused by the formation of local dipole moments, which result from dangling bonds on a distorted surface coupled to attached water molecules. Holding the bond distances fixed, we show that this effect occurs at a variety of bond angles. We show that the most probable angle is [Formula: see text], implying that this technique can be applied to the study of amorphous surfaces.
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