Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where J.D. Eick is active.

Publication


Featured researches published by J.D. Eick.


Journal of Dental Research | 1998

A TEM Study of Two Water-based Adhesive Systems Bonded to Dry and Wet Dentin

B. Van Meerbeek; Yasuhiro Yoshida; Paul Lambrechts; Guido Vanherle; E.S. Duke; J.D. Eick; S.J. Robinson

To keep the exposed collagen scaffold penetrable to resin, it has been recommended that the conditioned dentin surface be maintained in a visibly moist condition, a clinical technique commonly referred to as wet bonding. In this study, resin-dentin interfaces produced with two water-based adhesive systems-OptiBond (OPTI, Kerr) and Scotchbond Multi-Purpose (SBMP, 3M)-were compared by transmission electron microscopy, following the application of either a dry- or a wet-bonding technique. The hypothesis advanced was that the ultramorphology of the hybrid layer would differ depending on which bonding method was applied. A morphologically well-organized hybrid layer of collagen fibrils intermingled with resin in tiny interfibrillar channels was consistently formed by the OPTI system. The SBMP system was found to produce a hybrid layer with a more variable ultrastructure, less distinctly outlined collagen fibrils, and a characteristic electron-dense phase located at its surface. No major differences in hybrid layer ultrastructure were observed when the two adhesive systems investigated were bonded to either dry or wet dentin. When the adhesives were dry-bonded, no ultrastructural evidence of collapsed demineralized collagen, incompletely or not at all infiltrated by resin, could be detected. In addition, when the two adhesives were bonded to wet dentin, no signs of overwetting phenomena, that would have indicated that water was ineffectively removed, were apparent. It has been hypothesized that the amount of water provided with the hydrophilic primer solution of either of the two adhesive systems investigated suffices to re-hydrate and re-expand the gently air-dried and collapsed collagen network. Further research should be directed to determine whether this hypothesized self-rewetting effect can be extrapolated to other adhesive systems that provide water-based primers.


Journal of Dental Research | 1996

Correlative Transmission Electron Microscopy Examination of Nondemineralized and Demineralized Resin-Dentin Interfaces Formed by Two Dentin Adhesive Systems

B. Van Meerbeek; L.J. Conn; E.S. Duke; J.D. Eick; S.J. Robinson; D. Guerrero

The resin-dentin interface formed by two dentin adhesives, Optibond (OPTI, Kerr) and Scotchbond MultiPurpose (SBMP, 3M), was ultramorphologically examined by transmission electron microscopy (TEM). Ultrastructural information from nondemineralized and demineralized sections was correlated. It was hypothesized that the different chemical formulations of the two adhesives would result in a different morphological appearance of the hybrid layer. Ultrastructural TEM examination proved that each of the two dentin adhesive systems was able to establish a micromechanical bond between dentin and resin with the formation of a hybrid layer. However, the interfacial hybridization process that took place to produce this resin-dentin bond appeared to be specifically related to the chemical composition and application modes of both systems. OPTI consistently presented with a hybrid layer with a relatively uniform ultrastructure, electron density, and acid resistance. These three parameters were found to be more variable for the hybrid layer formed by SBMP. Characteristic of SBMP was the identification of an amorphous phase deposited at the outer surface of the hybrid layer. Although both adhesive systems investigated follow a total-etch concept, their specific chemical formulations result in different interfacial ultrastructures that are probably related to different underlying bonding mechanisms. The clinical significance of these morphological findings, however, is still unknown.


Journal of Biomedical Materials Research | 2001

The stability of methacrylate biomaterials when enzyme challenged: Kinetic and systematic evaluations

David M. Yourtee; R. E. Smith; K. A. Russo; S. Burmaster; J. M. Cannon; J.D. Eick; E. L. Kostoryz

This study addressed whether methacrylate monomers and polymers used in dentistry might degrade from enzymolysis by acetylcholinesterase (ACHE), cholesterol esterase (CHE), porcine liver esterase (PRLE), and a pancreatic lipase (PNL). Short (hour) and long-term (day) exposures were performed. Product ratios were used to determine surface hydrolysis of the polymeric materials. Enzyme kinetics were studied for the monomers when challenged by ACHE, CHE, and PRLE. In the case of PRLE, the V(max) for the dimethacrylate substrates varied slightly, but amounted to as much as 10% of that of p-nitrophenylacetate. The K(m) for triethylene glycol dimethacrylate (TEGDMA) was 197 microM for ACHE and 1107 microM for CHE. The V(max) was 2.7 nmol/min for ACHE and 3.5 nmol/min for CHE. TEGDMA was converted by CHE at 2% the rate of cholesteryl oleate. Long-term incubations of monomers with CHE and ACHE produced degrees of hydrolysis that evidenced structure dependency in the ability of the enzymes to effect hydrolysis. Particularly resistant were aromativ derivatives and those with branching in methacrylate linkages. Overall, the study confirms the ability of physiologically important esterases to catalyze the hydrolysis of biomaterial methacrylates.


Journal of Biomedical Materials Research | 1998

Application of solubility parameter theory to dentin-bonding systems and adhesive strength correlations

R.G. Miller; C.Q. Bowles; Cecil C. Chappelow; J.D. Eick

The principal aim of this study was to investigate the relationships between the solubility parameters of ectched dentin, and adhesive primer solutions and adhesive bond strength. Solubility parameters characterize the molecular interactions which determine physical properties such as wetting, and thus can serve as tools to aid development of polymeric adhesives and interpenetrating polymer networks. If an adhesive monomer has a solubility parameter close to that of a polymer substrate, then the monomer may act as a solvent for the polymer and penetrate below the surface. Subsequent polymerization of the monomer may then produce an interpenetrating network, thus adhering without necessarily forming primary chemical bonds to the substrate. The dentin substrate considered in this study was abraded dentin treated with ethylenediaminetetraaceitc acid. Solubility parameters delta pr, delta h, and delta d calculated for the etched dentin substrate were 20.3, 23.6, and 16.0 (J/cm3)1/2, respectively. Solubility parameters of the primers were expressed using Hansens three-dimensional scheme. The data indicate a correlation between the calculated solubility parameters of the etched dentin, and dentin primers and the resulting bond strengths. The results corroborate the significance of solubility parameter considerations for adhesive bonding to dentin.


Journal of Dental Research | 1998

Elements of Light-cured Epoxy-based Dental Polymer Systems

Frank Millich; L. Jeang; J.D. Eick; Cecil C. Chappelow; Charles S. Pinzino

The greatest problem with current dental composite systems is their polymerization shrinkage. Extensive work is being done by many investigators to alleviate this problem. Our approach has been to examine epoxy- and spiro-orthocarbonate (SOC)-based resins. The hypothesis to be tested in this study was that the cure characteristics of experimental visible-light-cured epoxy resin systems are governed by the types and concentrations of co-reactants and activators. Resin samples containing onium salt initiators and a thiozanthone sensitizer were successfully cured by means of either an experimental visible-light irradiation system or a commercially available dental lamp. Test resins consisted of di-epoxies alone or in combination, epoxy mixtures in combination with an SOC, or an epoxy in combination with a caprolactone-derived polyol. Significant findings were as follows: (a) Resins containing the SOC had longer cure times than their counterparts; (b) the optimum ratios of epoxy to polyol for most rapid cure were 50:50 or 60:40 under conditions tested; (c) resins containing TONE 305 polyol generally were faster to cure than those containing no polyol, or TONES 201 or 310; and (d) a resin mixture was found that had a cure time of 1 to 3 min when irradiated with a commercial dental lamp. Based on this exploratory study, it should be possible for clinically relevant cure times to be achieved for visible-light-cured epoxy-based resins by careful manipulation and optimization of key elements.


Journal of Dental Research | 1995

Scanning Transmission Electron Microscopy/Energy-dispersive Spectroscopy Analysis of the Dentin Adhesive Interface Using a Labeled 2-Hydroxyethylmethacrylate Analogue:

J.D. Eick; S.J. Robinson; T.J. Byerley; Chappell Rp; Paulette Spencer; Cecil C. Chappelow

In an attempt to compare the morphology of the dentin adhesive interface and the wetting and penetration of the adhesive in relation to the dentin surface, we studied four dentin adhesive systems using scanning transmission electron microscopy (STEM) and energy-dispersive spectroscopy (EDS). 2-Hydroxyethylmethacrylate (HEMA), a monomer common to many commercial dentin adhesive systems, was altered to produce a thiolated analogue (HETMA). Sulfur, traceable by EDS and STEM, was substituted for the oxygen atom in the backbone of the HEMA molecule. The resulting analogue, with solubility parameters and other wetting and physical properties very similar to those of HEMA, was applied to four sets of tooth specimens, each pre-treated with a different primer or etchant. Three separate pre-treatments-nitric acid, maleic acid, and citric acid/ferric chloride—created a demineralized zone approximately 1 to 3 μm thick at the dentin surface. The HETMA was found to permeate freely into this zone when either of the latter two pre-treatments was used. However, the band of dentin that was demineralized by the nitric acid pre-treatment appeared impermeable to the HETMA. The fourth pre-treatment, an alcohol-based solution including the phosphorus acid ester PENTA and HEMA, modified the smear layer of the tooth slightly and did not appear to demineralize the dentin. HETMA applied to the specimens pre-treated with PENTA and HEMA was clearly in intimate contact with the dentin or modified smear layer; however, it did not penetrate or diffuse into these areas. It did flow into the dentinal tubules, as was also evident with each of the other systems. It was concluded that the acid pre-treatment of the dentin greatly influenced the wetting behavior of the dentin adhesive and thus could substantially affect the resultant bond strength of the dentin adhesive systems.


Journal of Dental Research | 1996

Quantitative Analysis of the Dentin Adhesive Interface by Auger Spectroscopy

J.D. Eick; R.G. Miller; S.J. Robinson; C.Q. Bowles; P.L. Gutshall; Cecil C. Chappelow

The ultimate success of a dentin adhesive bond is dependent in large part on specific conditions at the interface between the tooth and the adhesive. Most current dentin adhesive systems use some sort of pre-treatment to demineralize the first few microns of the dentin surface, leaving a meshwork of collagen into which the adhesive resin can penetrate, infiltrate, and polymerize. The general hypothesis tested in this experiment was that the penetration and distribution of adhesive resin into the demineralized zone are a function of the conditioner used as a pre-treatment for the adhesive application. Four commercially available adhesive systems were modified to incorporate hydroxyethylthiomethacrylate (HETMA), a sulfur-substituted, traceable analogue of 2-hydroxyethylmethacrylate (HEMA), thereby allowing for a qualitative measurement of the amount and distribution of monomer in the treated dentin substrate by energy-dispersive x-ray spectroscopy (EDS) and a quantitative measurement by Auger electron spectroscopy (AES). The dentin pre-treatments investigated were: (1) 10% citric acid/3% ferric chloride, (2) 10% maleic acid, (3) 2.5% nitric acid, and (4) an alcoholic solution of HEMA with a phosphorus acid ester. These pre-treatments were applied to freshly extracted teeth that had been sectioned to expose the dentin and ground to simulate the smeared layer. After the appropriate pre-treatment was applied, a 10% (v/v) solution of HETMA in acetone was applied to the surface, followed by the corresponding adhesive resin, which was then polymerized. The samples were then processed for observation by scanning transmission electron microscopy (STEM), AES, and STEM/EDS analysis. The results indicated significant differences in the ability of HETMA to penetrate the dentin surface conditioned by the four pre-treatments investigated here. This study also demonstrated that AES and STEM/EDS could be used in a correlative fashion to determine the distribution of HETMA within or adjacent to the treated dentin surface.


Journal of Dental Research | 2001

Effects of Dental Resins on TNF-α-induced ICAM-1 Expression in Endothelial Cells

Elisabet L. Kostoryz; P.Y. Tong; A.F. Strautman; Alan G. Glaros; J.D. Eick; David M. Yourtee

Many reports have demonstrated inflammation after the placement of dental restorations. To explain this side-effect, we studied a biomarker in the inflammatory response. The intercellular adhesion molecule-1 (ICAM-1) is a key mediator for recruitment of leukocytes to the site of inflammation. Therefore, we investigated whether methacrylates (a BISGMA-based dental resin, BISGMA, and MAA) and Cyracure™ UVR 6105, an epoxy monomer, could alter ICAM-1 expression in unstimulated and TNF-a-stimulated endothelial cells. Six-well plates with monolayers of human umbilical vein cells, ECV 304 (ATCC CRL 1998), were exposed to TNF-a (1 ng/mL) in the presence and absence of subtoxic and TC50 doses of chemicals for 24 hrs at 37°C/5% CO2. Several doses of TNF-a (0.5-2 ng/mL) were co-incubated with 100 μL of undiluted aqueous dental resin extracts. Cells were harvested and stained with mAB FITC-conjugated anti-human ICAM-1 (CD54). ICAM-1 expression was measured by flow cytometry. Cells expressed basal levels of ICAM-1, which was up-regulated by TNF-a but was not changed by all samples studied. Except for UVR 6105, the methacrylates significantly decreased ICAM-1 expression in TNF-α-stimulated cells. These findings suggest that methacrylates may decrease the recruitment of leukocytes to sites of inflammation.


Journal of Dental Research | 1996

Design and Development of Isocyanatoacrylates as Dental Adhesives

Cecil C. Chappelow; Thomas J. Byerley; Charles S. Pinzino; Frank Millich; J.D. Eick

During the last 12 years, significant progress has been made in the development of dental adhesive systems. Some of the more promising systems are based on multifunctional structures that contain polymerizable vinyl double bonds and reactive isocyanate groups. The utility of compounds with such structures as adhesives arises in part because their isocyanate functionality is available for reaction independently, without compromising the reactivity of the vinyl groups. The hypotheses tested in this investigation were: (1) that the monomer reactivity ratios (r1, r2) for the free-radical-initiated copolymerization of ethyl a-isocyanatoacrylate (a-EIA) and 2-isocyanatoethyl methacrylate (IEM) with selected vinyl monomers can be determined; (2) that these reactivity ratios can be used to establish Q (reactivity) and e (polarity) values for a-EIA and IEM; and (3) that these reactivity parameters can be useful in designing copolymers with controlled compositions for dental adhesive applications. The free-radical copolymerization characteristics of a-EIA and IEM were studied. The isocyanate monomers were copolymerized at seven comonomer ratios with n-butyl acrylate (NBA), methyl methacrylate (MMA), and styrene (STY). Reactivity ratios, r1 and r2, were calculated for each of the copolymer systems, giving: IEM (r1) = 0.38 and STY (r2) = 0.44; IEM (r1) = 1.19 and MMA (r2) = 0.84; IEM (r1) = 2.50 and NBA (r2) = 0.40; a-EIA (r1) = 2.20 and STY (r2) = 0.06; α-EIA (r1) = 7.00 and MMA (r2) = 0.10; and a-EIA (r1) = 23.50 and NBA (r2) = 0.04. The Q (reactivity) and e (polarity) values for IEM and a-EIA were calculated from r1 and r2 with use of the Alfrey-Price equations, giving, for IEM, Q = 0.89 and e = 0.60, and, for a-EIA, Q = 7.64 and e = 0.74. These reactivity parameters are useful for tailoring copolymers with controlled compositions and properties. Based on these calculated reactivity parameters, several copolymers of IEM [for example, IEM/2-hydroxyethyl methacrylate (HEMA)] are currently being prepared and evaluated as adhesives.


Journal of Applied Polymer Science | 2000

Photoreactivity of expanding monomers and epoxy‐based matrix resin systems

Cecil C. Chappelow; Charles S. Pinzino; L. Jeang; C.D. Harris; Andrew J. Holder; J.D. Eick

The relative photoreactivity of five expanding monomers (EMs) in homopolymerization, and as comonomers in a candidate low shrinkage dental matrix resin, were evaluated. The EMs were 2,8-dimethyl-1,5,7,11-tetraoxaspiro[5.5]undecane (DM-TOSU); 3,9-diethyl-3,9-dipropionyloxymethyl-1,5,7,11-tetraoxaspiro[5.5]undecane (DEDPM-TOSU); 1,3-dioxane-2-one (DOO); 4-methyl-1,3-dioxane-2-one (M-DOO); and 5,5-diethyl-1,3-dioxane-2-thione (DE-DOT). The candidate low shrinkage resin system was an 80/20 mixture of UVR-6105 epoxide/polytetrahydrofuran (Mn ≈ 250). All reaction mixtures contained a diaryliodonium salt as a photoinitiator and camphorquinone as photosensitizer. Reactivities were evaluated using photodifferential scanning calorimetry. For homopolymerizations, the reactivity ranking (based on time to exotherm peak and total enthalpy) was DE-DOT ≫ DM-TOSU > DOO > M-DOO ≥ DEDPM-TOSU. In the comonomer system, the reactivity ranking was M-DOO > DEDPM- TOSU > DM-TOSU > DOO ≥ DE-DOT. This experimental work was substantiated and extended by molecular modeling studies employing the AM1 semiempirical method. Heats of formation of protonated EM structures, and heats of formation and potential energies of possible polymerization pathways were estimated. The relative reactivities of EM-based polymerization systems are related to chemical structure and the dominance of the most favored reaction mechanism.

Collaboration


Dive into the J.D. Eick's collaboration.

Top Co-Authors

Avatar

S.J. Robinson

University of Missouri–Kansas City

View shared research outputs
Top Co-Authors

Avatar

C.Q. Bowles

University of Missouri

View shared research outputs
Top Co-Authors

Avatar

R.G. Miller

University of Missouri

View shared research outputs
Top Co-Authors

Avatar

Andrew J. Holder

University of Missouri–Kansas City

View shared research outputs
Top Co-Authors

Avatar

David M. Yourtee

University of Missouri–Kansas City

View shared research outputs
Top Co-Authors

Avatar

L. Jeang

University of Missouri–Kansas City

View shared research outputs
Top Co-Authors

Avatar

B. Van Meerbeek

Katholieke Universiteit Leuven

View shared research outputs
Top Co-Authors

Avatar

C.D. Harris

University of Missouri–Kansas City

View shared research outputs
Top Co-Authors

Avatar

E.S. Duke

University of Texas Health Science Center at San Antonio

View shared research outputs
Top Co-Authors

Avatar

Frank Millich

University of Missouri–Kansas City

View shared research outputs
Researchain Logo
Decentralizing Knowledge