Publications

Advanced Filters

2021

A method for the electrohydrodynamic deposition of carbonaceous materials utilizing an electrohydrodynamic cell comprising two electrodes comprised of a conductive material, by first combining a solid phase comprising a carbonaceous material and a suspension medium, placing the suspension between the electrodes, applying an electric field in a first direction, varying the intensity of the electric field sufficiently to drive lateral movement, increasing the electrical field to stop the lateral transport and fix the layers in place, then removing the applied field and removing the electrodes. Among the many different possibilities contemplated, the method may advantageously utilize: varying the spacing between the electrodes; removing the buildup from one or both electrodes; placing the electrodes into different suspensions; adjusting the concentration, pH, or temperature of the suspension(s); and varying the direction, intensity or duration of the electric fields.

2020

SubSim, Hyung et al. “Multifunctional Graphene-Based Additives for Enhanced Combustion of Cracked Hydrocarbon Fuels under Supercritical Conditions.” Combustion Science and Technology 192.7 (2020): 1420–1435.
As a particulate fuel additive, functionalized graphene sheets (FGS), with and without the decoration of nanoparticles, provide a means to form stable colloids with liquid hydrocarbons, act as in situ catalysts, and prevent attached nanoparticles from agglomerating or sintering during heating. Recent pyrolysis experiments and simulations have shown the synergetic effect of Pt and FGS structures accelerating fuel conversion and hydrogen formation. In this paper, the role of graphene-based additives (with and without platinum (Pt) nanoparticles) on enhancing the ignition and combustion characteristics of cracked n-dodecane hydrocarbon fuels under supercritical conditions is examined. Supercritical ignition and combustion experiments were conducted using a high pressure and high temperature windowed combustion chamber coupled to a supercritical fuel pyrolysis reactor. The combustion experiments indicated that the presence of small amounts of particulate additives (100 ppmw) to the fuel can reduce ignition times and increase subsequent combustion rates. In particular, the addition of FGS-supported Pt nanoparticles reduced ignition delay times by nearly a factor of 3 (12.4 to 4.2 ms), increased spray spreading angles by approximately 32.0% (15.4 to 20.3°), reduced the flame liftoff length by 54.0% (1.74 to 0.8 mm), and demonstrated an increase in fuel conversion for a fixed reaction time of 35.0% relative to the pure fuel baseline. These results support the notion that nanostructured fuel additives have the potential to enhance both the heat sink capacity and combustion performance of liquid hydrocarbon fuels for their use in advanced high speed propulsion systems.

Experiments and simulations were used to demonstrate that decorating functionalized graphene sheets (FGSs) with platinum nanoparticles (Pt@FGS) stabilized these particles. Addition of these particles to liquid hydrocarbon fuels was observed to significantly affect decomposition under supercritical conditions at a pressure of 4.75 MPa and temperatures from 753 to 803 K. The suspension of only 50 ppmw Pt@FGS in the fuel (equivalent to adding 10 ppmw Pt) enhanced fuel conversion rates (by up to 24%) with a major effect on specific product yields. The production of low-molecular-weight species increased in the pyrolysis products (with the hydrogen yield increasing by a factor of 12.5). ReaxFF molecular dynamics (MD) simulations supported a mechanism in which synergy between Pt and FGS catalyzed dehydrogenation during n-C12H26 pyrolysis. The highest conversion rates and greatest yields of hydrogen and low-molecular-weight species were observed for fuels containing Pt@FGS particles rather than those containing either FGSs or Pt-clusters alone. Analysis of the platinum decorated FGSs post reaction indicated no deterioration of the composite particles.

SubSim, Hyung et al. “Functionalized Graphene Sheet As a Dispersible Fuel Additive for Catalytic Decomposition of Methylcyclohexane.” Combustion and Flame 217 (2020): 212–221.
The decomposition of methylcyclohexane (MCH) containing suspended functionalized graphene sheets (FGS) was studied using a high-pressure flow reactor, under supercritical conditions over the temperature range of 780 – 825 K at a constant pressure of 4.72 MPa. Experiments showed both fuel conversion rates and C1-C2 product yields were increased by 43.3% and 62.1%, respectively, with the addition of 50 ppmw of FGS at 820 K. The reaction mechanisms between hydrocarbon and FGS were computationally investigated using reactive molecular dynamics (MD) with ReaxFF force fields at the temperatures of 1700, 1800, and 1900 K. The MD simulations revealed that oxygen-containing functional groups attached to the FGS plays an important role in catalyzing the decomposition of the fuel. Heterogeneous dehydrogenation of MCH into a C7H13 radical intermediate led to formation of secondary and tertiary radicals, such as H, CH3, and C2H5, during early extents of reaction promoting additional fuel-consuming steps, including H-abstraction and hydrogenation.

2019

Embodiments of the present invention relate to battery electrodes incorporating composites of graphene and selenium-sulfur compounds for improved rechargeable batteries. In one embodiment, a conductive composition comprises a conductive composition having a Se—S compound, a conductive additive. The Se—S compound is present as SexS8-x, wherein 0<x<8.

2016

The present invention relates generally to carbon structures and specifically to electrohydrodynamically formed structures of carbonaceous material.  Coatings, such as physically coherent films, coatings, membranes, or tapes made from high carbon content materials, such as graphene sheets, can be assembled using electrophoretic deposition, tape casting, spin casting, drop casting, or filtration.  Cast or filtered structures typically have to be at least 400 nm thick to provide continuity and mechanical stability.  Such structures contain flaws created by removing the liquid through drying or filtration.  In addition, such structures can have a reduced flexibility and compliance, which can result in an increase in susceptibility to damage during transfer and/or fitting to the item to be covered.  Similar to electrohydrodynamic deposition, electrophoretic deposition uses an applied electric field to attract particles or sheets to a surface having an overall charge opposite to the charge intrinsic to or induced on the particles or sheets, thereby coating the surface, as described in United States patent No. 2,894,888 to Shyne, et al., and United States patent No. 3,932,231 to Hara, et al., and many others.  However, in electrophoretic deposition the particles or sheets adhere at the point of initial contact to the substrate or previously deposited layers and remain fixed in position, which leaves defects or gaps between the particles or sheets comprising the layers that constitute the coating, membrane, or film.  A fully dense covering requires several layers, resulting in increased thickness of the coating, membrane, or film which limits its applications.

2015

We have covalently grafted tetrazine derivatives to graphene oxide through nucleophilic substitution. Since the tetrazine unit is electroactive and nitrogen-rich, with a reduction potential sensitive to the type of substituent and degree of substitution, we used electrochemistry and X-ray photoelectron spectroscopy to demonstrate clear evidence for grafting through covalent bonding. Chemical modification was supported by Fourier transform infrared spectroscopy and thermal analysis. Tetrazines grafted onto graphene oxide displayed different mass losses compared to unmodified graphene and were more stable than the molecular precursors. Finally, a bridging tetrazine derivative was grafted between sheets of graphene oxide to demonstrate that the separation distance between sheets can be maintained while designing new graphene-based materials, including chemically bound, redox structures.

Zhang, Cui et al. “Combined Effects of Functional Groups, Lattice Defects, and Edges in the Infrared Spectra of Graphene Oxide.” The Journal of Physical Chemistry C 119 (2015): 18167–18176. Print.

Infrared spectroscopy in combination with density functional theory calculations has been widely used to characterize the structure of graphene oxide and its reduced forms. Yet, the synergistic effects of different functional groups, lattice defects, and edges on the vibrational spectra are not well understood. Here, we report first-principles calculations of the infrared spectra of graphene oxide performed on realistic, thermally equilibrated, structural models that incorporate lattice vacancies and edges along with various oxygen-containing functional groups. Models including adsorbed water are examined as well. Our results show that lattice vacancies lead to important blue and red shifts in the OH stretching and bending bands, respectively, whereas the presence of adsorbed water leaves these shifts largely unaffected. We also find unique infrared features for edge carboxyls resulting from interactions with both nearby functional groups and the graphene lattice. Comparison of the computed vibrational properties to our experiments clarifies the origin of several observed features and provides evidence that defects and edges are essential for characterizing and interpreting the infrared spectrum of graphene oxide.

2012

Dabbs, and Aksay. “Multifunctional and Low-Density Inorganic Nanocomposites.” JOM 64 (2012): 226–233. Print.

We summarize our recent studies on the use of low-density nanoporous silica structures prepared through templating of a self-assembling disordered liquid-crystalline L (3) phase, as a matrix for use in numerous applications, including sensing, optical data storage, drug release, and structural. The silica matrix exhibits low density (0.5 g cm(-3) to 0.8 g cm(-3) for monoliths, 0.6 g cm(-3) to 0.99 g cm(-3) for fibers) coupled with high surface areas (up 1400 m(2) g(-1)) and void volumes (65% or higher). High-surface-area coatings are used to increase the sensitivity of mass-detecting quartz crystal microbalances to over 4000 times that of uncoated crystals. Monoliths, films, and fibers are produced using the templated silica gel. Once dried and converted to silica, the nanostructured material exhibits high fracture strength (up to 35 MPa in fibers) and Young's modulus (30 GPa to 40 GPa in fibers). These values are, respectively, two orders of magnitude and twice those of nanostructured silicas having comparable densities.

2011

Korkut et al. “High Surface Area Tapes Produced With Functionalized Graphene.” ACS Nano 5 (2011): 5214–5222. Print.

We describe a scalable method for producing continuous graphene networks by tape casting surfactant-stabilized aqueous suspensions of functionalized graphene sheets. Similar to all other highly connected graphene-containing networks, the degree of overlap between the sheets controls the tapes' electrical and mechanical properties. However, unlike other graphene-containing networks, the specific surface area of the cast tapes remains high (>400 m(2).g(-1)). Exhibiting apparent densities between 0.15 and 0.51 g.cm(-3), with electrical conductivities up to 24 kS.m(-1) and tensile strengths over 10 MPa, these tapes exhibit the best combination of properties with respect to density heretofore observed for carbon-based papers, membranes, or films.

2009

Sabourin et al. “Functionalized Graphene Sheet Colloids for Enhanced Fuel Propellant Combustion.” ACS Nano 3 (2009): 3945–3954. Print.

We have compared the combustion of the monopropellant nitromethane with that of nitromethane containing colloidal particles of functionalized graphene sheets or metal hydroxides. The linear steady-state burning rates of the monopropellant and colloidal suspensions were determined at room temperature, under a range of pressures (3.35-14.4 MPa) using argon as a pressurizing fluid. The ignition temperatures were lowered and burning rates increased for the colloidal suspensions compared to those of the liquid monopropellant alone, with the graphene sheet suspension having significantly greater burning rates (i.e., greater than 175%). The relative change in burning rate from neat nitromethane increased with increasing concentrations of fuel additives and decreased with increasing pressure until at high pressures no enhancement was found.

2008

Sai, Dabbs, and Aksay. “Mechanical Properties of L3-Templated Nanostructured Silica.” ACS Polymer Preprints 49.2 (2008): 546–547. Print.

2007

Adamson et al. “Non-Peptide Polymeric Silicatein Alpha Mimic for Neutral PH Catalysis in the Formation of Silica.” Macromolecules 40 (2007): 5710–5717. Print.

We have synthesized a catalytically active polymer inspired by the naturally occurring protein silicatein alpha and have shown it to catalyze the formation of silica from tetraethoxysilane under near-neutral pH and ambient temperatures. We based the composition of the polymer on the functionalities found in silicatein alpha, specifically those essential components of the catalytically active site for the hydrolysis of silicon alkoxides. Our bioinspired polymer is a block copolymer of poly(2-vinylpyridine-b-1,2-butadiene), functionalized by the addition of hydroxyl groups via hydroboration chemistry. The catalytic action of our polymer on tetraethoxysilane at neutral pH and ambient temperature conditions has been confirmed using a modified molybdic acid assay method, thermogravimetric analysis, and Fourier transform infrared spectroscopy. The structure of the resulting gel is investigated by scanning electron microscopy and solid-state nuclear magnetic resonance. The microscopic features of the material formed resemble that of gels formed by the acid-catalyzed hydrolysis of tetraethoxysilane.

2006

Bhansali et al. “The Stability of L3 Sponge Phase in Acidic Solutions.” Langmuir 22 (2006): 4060–4064. Print.

In the synthesis of the disordered lyotropic liquid crystalline L-3 sponge phase prepared with the cosurfactants cetylpyridinium chloride and hexanol, aqueous NaCl solution is used as the solvent. When this sponge phase is used as the template for L3 silica-phase processing, we replace NaCl with HCl to facilitate the acid catalysis of tetramethoxysilane in forming a templated silica gel, assuming that changing the solvent from NaCl(aq) to HCl(aq) of equivalent ionic strength does not affect the stability range of the L3 phase. In this work, we confirm that changing the pH of the solvent from neutral to acidic (with HQ has negligible effect on the L3 phase region. Equivalent ionic strength is provided by either NaCl(aq) or HCl(aq) solvent; therefore, a similar phase behavior is observed regardless of which aqueous solvent is used.

Dabbs, Mulders, and Aksay. “Solvothermal Removal of the Organic Template from L3 (‘sponge’) Templated Silica Monoliths.” Journal of Nanoparticle Research 8 (2006): 603–614. Print.

We compare the methods of continuous solvent (Soxhlet) and supercritical solvent extractions for the removal of the organic template from nanostructured silica monoliths. Our monoliths are formed by templating the L-3 liquid crystal phase of cetylpyridinium chloride in aqueous solutions with tetramethoxy silane. The monoliths that result from both Soxhlet and supercritical extraction methods are mechanically robust, optically clear, and free of cracks. The Soxhlet method compares favorably with supercritical solvent extraction in that equivalent L-3-templated silica can be synthesized without the use of specialized reactor hardware or higher temperatures and high pressures, while avoiding noxious byproducts. The comparative effectiveness of various solvents in the Soxhlet process is related to the Hildebrand solubility parameter, determined by the effective surface area of the extracted silica.

Malik et al. “Silica Monoliths Templated on L3 Liquid Crystal.” Langmuir 22 (2006): 325–331. Print.

Dimensionally stable, optically clear, highly porous (similar to 65% of the apparent volume), and high surface area (up to 1400 m(2)/g) silica monoliths were fabricated as thick disks (0.5 cm) by templating the isotropic liquid crystalline L-3 phase with silica through the hydrolysis and condensation of a silicon alkoxide and then removing the organic constituents by supercritical ethanol extraction. The L3 liquid crystal is a stable phase formed by the cosurfactants cetylpyridinium chloride monohydrate and hexanol in HCl(aq) solvent. Extracted 0.5 cm thick disks exhibited a low ratio of scattered to transmitted visible light (1.5 x 10(-6) at 22 from the surface normal). The degree of silica condensation in the monoliths was high, as determined by Si-29 NMR measurements of Q(3) and Q(4) peak intensities (0.53 and 0.47, respectively). As a result, the extracted and dried monoliths were mechanically robust and did not fracture when infiltrated by organic solvents. Photoactive liquid monomers were infiltrated into extracted silica monoliths and polymerized in situ, demonstrating the possible application of templated silica to optical storage technology.

2005

Dabbs et al. “Inhibition of Aluminum Oxyhydroxide Precipitation With Citric Acid.” Langmuir 21 (2005): 11690–11695. Print.

Citric acid has been shown to act as an agent for increasing the solubility of aluminum oxyhydroxides in aqueous solutions of high (>2.47 mol/mol) hydroxide-to-aluminum ratios. Conversely, citric acid also colloidally stabilizes particles in aqueous suspensions of aluminum-containing particles. Solutions of aluminum chloride, with and without citric acid added, were titrated with NaOH(aq), The presence and size of particles were determined using quasi-elastic light scattering. In solutions that contained no citric acid, particles formed instantaneously when NaOH(aq) was added but these were observed to rapidly diminish in size, disappearing at OH/Al ratios below 2.5 mol/mol. When the OH/Al ratio was raised beyond 2.5 by adding more NaOH(aq), suspensions of colloidally stable particles formed. Large polycations containing 13 aluminum atoms were detected by Al-27 solution NMR in citric-acid-free solutions with OH/Al ratios slightly lower than 2.5. In comparison, adding citric acid to solutions of aluminum chloride inhibited the formation of large aluminum-containing polycations. The absence of the polycations prevents or retards the subsequent formation of particles, indicating that the polycations, when present, act as seeds to the formation of new particles. Particles did not form in solutions with a citric acid/aluminum ratio of 0.8 until sufficient NaOH(aq) was added to raise the OH/Al ratio to 3.29. By comparison, lower amounts of citric acid did not prevent particles from forming but did retard the rate of growth.

2004

Adamson et al. “Non-Peptide, Silicatein-α Inspired Silica Condensation Catalyst.” ACS Abstracts 227 (2004): 145. Print.

2003

2001

Malik et al. “Disordered Mesoporous Silicates Formed by Templation of a Liquid Crystal (L3).” Solid-State Chemistry of Inorganic Materials III 2001: n. pag. Print.

2000

Dabbs, and Aksay. “Self-Assembled Ceramics Produced by Complex-Fluid Templation.” Annual Review of Physical Chemistry 51 (2000): 601–622. Print.

This review examines the use of self-assembly in the fabrication of ceramic mesostructures, emphasizing the use of amphiphilic surfactants and block copolymers. The association between this area of research and biomimetics is discussed, linking developments in synthetic self-assembly with biomineralization. The fabrication of hierarchical structures through the use of simultaneous processing is shown to be a necessary condition for applications of this new technology.

Hsu et al. “Sol-Gel Coated Glass Cells for Spin-Exchange Polarized 3He.” Applied Physics Letters 77 (2000): 2069–2071. Print.

We have developed a high-purity sol-gel coating for the interior surface of glass cells used for polarizing He-3 by spin-exchange optical pumping. The coating is designed to minimize spin relaxation due to wall collisions. A longitudinal spin-relaxation time T-1 in a sol-gel coated Pyrex cell of 344 +/- 8 h was achieved, the longest T-1 we have ever recorded for a gaseous sample. Repeated trials indicated that the coating was quite robust. Results using an uncoated Pyrex cell were also quite good, although inferior to the performance of the coated cell. (C) 2000 American Institute of Physics. [S0003-6951(00)02839-4].

McGrath et al. “Silica Gels With Tunable Nanopores through Templating of the L3 Phase.” Langmuir 16 (2000): 398–406. Print.

We describe a detailed synthesis of a silicified inorganic/organic nanoporous monolithic composite conforming to the lyotropic liquid crystalline L-3 phase. The pore dimensions of the silicified LQ phase scale with the solvent volume fraction in the synthesis reaction mixture. Changing the solvent fraction in the initial solution changes the ultimate pore diameter in the silicate, providing a simple method for tuning the diameter of the pores in the matrix. The resulting monolith is optically isotropic and transparent with a nonperiodic network. Accessible pores (which permeate the entire structure) in the silicified materials correlate with the solvent domain of the original liquid crystalline phase and therefore negate the need to remove the surfactant in order to gain access to the pore network. Measured characteristic dimensions are from 6 to well over 35 nm. X-ray scattering studies indicate a low polydispersity in the pore diameters for a given solvent fraction. Transmission electron and atomic force microscope images are consistent with a random morphology and measured surface areas exceed 960 m(2) g(-1) in extracted materials.

1999

Dabbs, Yao, and Aksay. “Nanocomposite Mullite Mullite Powders by Spray Pyrolysis.” Journal of Nanoparticle Research 1 (1999): 127–130. Print.

A mullite/mullite nanocomposite powder has been synthesized, composed of nanometer-size 3Al(2)O(3)center dot 2SiO(2) ('3 : 2') mullite precipitates within a matrix of the high alumina 2Al(2)O(3)center dot SiO2 ('2 : 1') mullite. Historically, the transition from the metastable high-alumina phase to the thermodynamically stable '3 : 2' phase of mullite has been thought to be a continuous process, involving a continuous solid solution between the two forms of mullite. In contradiction to this widely held view, our high resolution transmission electron microscopic characterization confirms that a first order phase transition between two distinct mullites occurs. The high degree of interface coherence between the precipitates and the matrix allows us to speculate that the mechanical properties of the matrix could be enhanced by a process similar to the precipitation hardening of metals.

1998

McGrath et al. “Formation of a Silicate L3 Phase With Continuously Adjustable Pore Sizes.” Science 279 (1998): 1289–1289. Print.

1997

Lahiri et al. “Porphyrin Amphiphiles As Templates for the Nucleation of Calcium Carbonate.” Journal of the American Chemical Society 119 (1997): 5449–5450. Print.
McGrath et al. “Formation of a Silicate L3 Phase With Continuously Adjustable Pore Sizes.” Science 277 (1997): 552–556. Print.

The lyotropic L-3 phase was used as a template to form nanoporous monolithic silicates with continuously adjustable pore sizes. The monolith was optically isotropic and transparent with a nonperiodic network. The pore size was adjusted by a change in the solvent volume fraction rather than by a change of the surfactant. Unlike other silicates, the bicontinuous pores were water-filled; removal of surfactant was not necessary to access the pores. Measured characteristic dimensions were from six to more than 35 nanometers. For a given solvent fraction, x-ray scattering indicated little variation of pore widths, in marked contrast to the polydisperse pores of aerogels.

McHale et al. “Metastability of Spinel-Type Solid Solutions in the SiO2-Al2O3 System.” Chemistry of Materials 9 (1997): 3096–3100. Print.

The addition of small amounts (2-10 wt %) of SiO2 to gamma-Al2O3 increases the temperature of heat treatment necessary for transformation to alpha-Al2O3 by similar to 100 K. We have studied this system using high-temperature solution calorimetry in molten 2PbO . B2O3 at 1043 K, Our results indicate that the spinel-type Al2O3-SiO2 solid solutions with 2-10 wt % SiO2 are always energetically metastable by 30-35 kJ.mol(-1) (on a 4 O2- per mole basis) with respect to alpha-Al2O3 and quartz. Calculation of the maximum configurational entropy of the solid solutions allowed determination of the likely most negative value of the Gibbs free energy of the materials, The solid solutions are somewhat entropy stabilized, but still thermodynamically metastable by > 10 kJ.mol(-1) at 1400 K, Therefore, SiO2 addition appears to provide mainly a kinetic hindrance to alpha-Al2O3 formation.

1996

Lahiri et al. Biomimetic Fabrication of Materials - The Minimalist Approach. Vol. 2716. N.p., 1996. Print.
Rintoul et al. “Structure and Transport Properties of a Porous Magnetic Gel via X-Ray Microtomography.” Physical Review E 54 (1996): 2663–2669. Print.

A three-dimensional digitized image of a porous magnetic gel is determined by x-ray microtomographic techniques. The complex connected pore-space topology is quantitatively characterized by measuring a variety of statistical correlation functions, including the chord-length distribution function, the pore-size distribution function, and the Lineal-path function. This structural information is then employed to estimate transport properties, such as the fluid permeability and trapping rate, of the gel.

Treadwell, Dabbs, and Aksay. “Mullite (3Al2O3-2SiO2) Synthesis With Aluminosiloxanes.” Chemistry of Materials 8 (1996): 2056–2060. Print.

High-yield mullite (3A1(2)O(3)-2SiO(2)) precursors consist of aluminosiloxanes :synthesized from mixtures of aluminum and silicon alkoxides. Atomic level mixing of the aluminum and silicon oxides is demonstrated by the low-temperature conversion (<1000 degrees C) of the aluminosiloxanes to phase-pure mullite. The proper selection of monomeric side groups serves several functions: (i) controlling reactivity of the silicon and aluminum monomers, thereby favoring atomic-level mixing; (ii) maintaining the tractability of the resulting aluminosiloxane; (iii) improving the yield during mullitization of the aluminosiloxane through easy thermolytic removal. The tractability of the aluminosiloxane compounds permits these materials to be used, in fiber spinning, the casting of thin films and monoliths, and as impregnants to powder compacts.

1995

Aksay et al. “Bioinspired Processing of Ceramic-Matrix Composites.” International Workshop on Ceramic Science and Technology 1995: 51–64. Print.
Aksay et al. “Synthesis and Processing of Nanostructured Ceramics.” Nano-Sized Powder and Related Composite Materials, NEPTIS-IV. Osaka: Nisshin Engineering Co. Ltd., 1995. Print.
Aksay, and Dabbs. “Bioinspired Processing of Nanostructured Materials.” 8th International Metallurgy and Materials Congress 1995: 775–794. Print.
Yao et al. “The Breakup of the Intermediate Gold Aggregates.” Microscopy and Microanalysis 1995: 196–197. Print.
Srinivasan et al. “Nanocomposite Processing via Infiltration of Mesoporous Silica.” Microscopy and Microanalysis 1995: 210–211. Print.

1994

Keckler et al. “Silification of Wood-Cell Walls.” Proceedings, 52nd Annual Meeting - Microscopy Society of America (1994): 428–429. Print.
Dabbs, Yao, and Aksay. “SURFACTANT MEDIATED DEFORMATION OF GOLD PARTICLES.” Proceedings, 52nd Annual Meeting - Microscopy Society of America (1994): 446–447. Print.
Jun et al. “Processing of Monolithic Magnetic Gels for Magnetophoresis.” Langmuir 10 (1994): 3377–3379. Print.

Magnetite particles and monosized polystyrene beads were trapped in a silica-gel, which was then dried by using supercritical fluid extraction. When the monolithic dried gel is sintered, the polystyrene beads are pyrolyzed, leaving a porous magnetized piece of ceramic with controlled pore sizes. These ''magnetic gel'' ceramics provide a novel class of materials for use in gel magnetophoresis and other biophysical applications.

1993

Aksay et al. “Bioinspired Processing of Ceramic-Matrix Composites.” Third Euro-Ceramics 1993: 405–418. Print.

1992

Hama, Dabbs, and Aksay. “Low Temperature Sintering of Ceramic Materials.” 1992: n. pag. Print.

1991

Aksay, Dabbs, and Sarikaya. “MULLITE FOR STRUCTURAL, ELECTRONIC, AND OPTICAL APPLICATIONS.” Journal of the American Ceramic Society 74 (1991): 2343–2358. Print.

Mullite (3Al2O3.2SiO2) is becoming increasingly important in electronic, optical, and high-temperature structural applications. This paper reviews the current state of mullite-related research at a fundamental level, within the framework of phase equilibria, crystal structure, synthesis, processing, and properties. Phase equilibria are discussed in terms of the problems associated with the nucleation kinetics of mullite and the large variations observed in the solid-solution range. The incongruent melting behavior of mullite is now widely accepted. Large variations in the solid solubility from 58 to 76 mol% alumina are related to the ordering/disordering of oxygen vacancies and are strongly coupled with the method of synthesis used to form mullite. Similarly, reaction sequences which lead to the formation of mullite upon heating depend on the spatial scale at which the components are mixed. Mixing at the atomic level is useful for low-temperature (< 1000-degrees-C) synthesis of mullite but not for low-temperature sintering. In contrast, precursors that are segregated are better suited for low-temperature (1250-degrees to 1500-degrees-C) densification through viscous deformation. Flexural strength and creep resistance at elevated temperatures are significantly affected by the presence of glassy boundary inclusions; in the absence of glassy inclusions, polycrystalline mullite retains > 90% of its room-temperature strength to 1500-degrees-C and displays very high creep resistance. Because of its low dielectric constant, mullite has now emerged as a substrate material in high-performance packaging applications. Interest in optical applications mainly centers on its applicability as a window material within the mid-infrared range.

1986

Shinohara, Dabbs, and Aksay. “Infrared Transparent Mullite through Densification of Monolithic Gels at 1250°C.” Infrared and Optical Transmitting Materials 1986: 12–18. Print.

1984

Marschman et al. “Preliminary Comments on the Influence of Slag Additives on the Solubility of Arsenic, Bismuth, and Selenium in Slag.” Metallurgical Slags and Fluxes 1984: 417–425. Print.
Dabbs et al. “INFLUENCE OF AL2O3 AND CAO ON THE SOLUBILITY OF AS, BI AND SE IN SILICATE SLAG.” Journal of Metals 36 (1984): 89–89. Print.

1983

Dabbs, and Lynch. “Analysis of the Molecular Form of Arsenic in Silicate Slag.” Advances in Sulfide Smelting. New York: The Metallurgical Society, AIME, 1983. Print.