36 research outputs found

    Composite WO<sub>3</sub>/TiO<sub>2</sub> Nanostructures for High Electrochromic Activity

    No full text
    A composite material consisting of TiO<sub>2</sub> nanotubes (NT) with WO<sub>3</sub> electrodeposited on its surface has been fabricated, detached from its Ti substrate, and attached to a fluorine-doped tin oxide (FTO) film on glass for application to electrochromic (EC) reactions. Several adhesion layers were tested, finding that a paste of TiO<sub>2</sub> made from commercially available TiO<sub>2</sub> nanoparticles creates an interface for the TiO<sub>2</sub> NT film to attach to the FTO glass, which is conductive and does not cause solution-phase ions in an electrolyte to bind irreversibly with the material. The effect of NT length and WO<sub>3</sub> concentration on the EC performance were studied. The composite WO<sub>3</sub>/TiO<sub>2</sub> nanostructures showed higher ion storage capacity, better stability, enhanced EC contrast, and longer memory time compared with the pure WO<sub>3</sub> and TiO<sub>2</sub> materials

    Reversible Hydrogen Storage by NaAlH<sub>4</sub> Confined within a Titanium-Functionalized MOF-74(Mg) Nanoreactor

    No full text
    We demonstrate that NaAlH<sub>4</sub> confined within the nanopores of a titanium-functionalized metalā€“organic framework (MOF) template MOF-74(Mg) can reversibly store hydrogen with minimal loss of capacity. Hydride-infiltrated samples were synthesized by melt infiltration, achieving loadings up to 21 wt %. MOF-74(Mg) possesses one-dimensional, 12 ƅ channels lined with Mg atoms having open coordination sites, which can serve as sites for Ti catalyst stabilization. MOF-74(Mg) is stable under repeated hydrogen desorption and hydride regeneration cycles, allowing it to serve as a ā€œnanoreactorā€. Confining NaAlH<sub>4</sub> within these pores alters the decomposition pathway by eliminating the stable intermediate Na<sub>3</sub>AlH<sub>6</sub> phase observed during bulk decomposition and proceeding directly to NaH, Al, and H<sub>2</sub>, in agreement with theory. The onset of hydrogen desorption for both Ti-doped and undoped nano-NaAlH<sub>4</sub>@MOF-74(Mg) is āˆ¼50 Ā°C, nearly 100 Ā°C lower than bulk NaAlH<sub>4</sub>. However, the presence of titanium is not necessary for this increase in desorption kinetics but enables rehydriding to be almost fully reversible. Isothermal kinetic studies indicate that the activation energy for H<sub>2</sub> desorption is reduced from 79.5 kJ mol<sup>ā€“1</sup> in bulk Ti-doped NaAlH<sub>4</sub> to 57.4 kJ mol<sup>ā€“1</sup> for nanoconfined NaAlH<sub>4</sub>. The structural properties of nano-NaAlH<sub>4</sub>@MOF-74(Mg) were probed using <sup>23</sup>Na and <sup>27</sup>Al solid-state MAS NMR, which indicates that the hydride is not decomposed during infiltration and that Al is present as tetrahedral AlH<sub>4</sub><sup>ĀÆ</sup> anions prior to desorption and as Al metal after desorption. Because of the highly ordered MOF structure and monodisperse pore dimensions, our results allow key template features to be identified to ensure reversible, low-temperature hydrogen storage

    Understanding and Mitigating the Effects of Stable Dodecahydro-<i>closo</i>-dodecaborate Intermediates on Hydrogen-Storage Reactions

    No full text
    Alkali metal borohydrides can reversibly store hydrogen; however, the materials display poor cyclability, oftentimes linked to the occurrence of stable <i>closo</i>-polyborate intermediate species. In an effort to understand the role of such intermediates on the hydrogen storage properties of metal borohydrides, several alkali metal dodecahydro-<i>closo</i>-dodecaborate salts were isolated in anhydrous form and characterized by diffraction and spectroscopic techniques. Mixtures of Li<sub>2</sub>B<sub>12</sub>H<sub>12</sub>, Na<sub>2</sub>B<sub>12</sub>H<sub>12</sub>, and K<sub>2</sub>B<sub>12</sub>H<sub>12</sub> with the corresponding alkali metal hydrides were subjected to hydrogenation conditions known to favor partial or full reversibility in metal borohydrides. The stoichiometric mixtures of MH and M<sub>2</sub>B<sub>12</sub>H<sub>12</sub> salts form the corresponding metal borohydrides MBH<sub>4</sub> (M = Li, Na, K) in almost quantitative yield at 100 MPa H<sub>2</sub> and 500 Ā°C. In addition, stoichiometric mixtures of Li<sub>2</sub>B<sub>12</sub>H<sub>12</sub> and MgH<sub>2</sub> were found to form MgB<sub>2</sub> at 500 Ā°C and above upon desorption in vacuum. The two destabilization strategies outlined above suggest that metal polyhydro-<i>closo</i>-polyborate species can be converted into the corresponding metal borohydrides or borides, albeit under rather harsh conditions of hydrogen pressure and temperature

    Structural Behavior of Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub>

    No full text
    On the basis of X-ray and neutron powder diffraction, first-principles calculations, and neutron vibrational spectroscopy, Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub> was found to exhibit atypical hexagonal symmetry to best stabilize the ionic packing of the relatively small Li<sup>+</sup> cations and large ellipsoidal B<sub>10</sub>H<sub>10</sub><sup>2ā€“</sup> anions. Moreover, differential scanning calorimetry and neutron-elastic-scattering fixed-window scans suggested that Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub>, similar to its polyhedral cousin Li<sub>2</sub>B<sub>12</sub>H<sub>12</sub>, undergoes an orderā€“disorder phase transition near 640 K. These results provide valuable structural information pertinent to understanding the potential role that Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub> plays during LiBH<sub>4</sub> dehydrogenationā€“rehydrogenation as well as its prospects as a superionic Li<sup>+</sup> cation conductor

    Structural Behavior of Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub>

    No full text
    On the basis of X-ray and neutron powder diffraction, first-principles calculations, and neutron vibrational spectroscopy, Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub> was found to exhibit atypical hexagonal symmetry to best stabilize the ionic packing of the relatively small Li<sup>+</sup> cations and large ellipsoidal B<sub>10</sub>H<sub>10</sub><sup>2ā€“</sup> anions. Moreover, differential scanning calorimetry and neutron-elastic-scattering fixed-window scans suggested that Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub>, similar to its polyhedral cousin Li<sub>2</sub>B<sub>12</sub>H<sub>12</sub>, undergoes an orderā€“disorder phase transition near 640 K. These results provide valuable structural information pertinent to understanding the potential role that Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub> plays during LiBH<sub>4</sub> dehydrogenationā€“rehydrogenation as well as its prospects as a superionic Li<sup>+</sup> cation conductor

    Structural Behavior of Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub>

    No full text
    On the basis of X-ray and neutron powder diffraction, first-principles calculations, and neutron vibrational spectroscopy, Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub> was found to exhibit atypical hexagonal symmetry to best stabilize the ionic packing of the relatively small Li<sup>+</sup> cations and large ellipsoidal B<sub>10</sub>H<sub>10</sub><sup>2ā€“</sup> anions. Moreover, differential scanning calorimetry and neutron-elastic-scattering fixed-window scans suggested that Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub>, similar to its polyhedral cousin Li<sub>2</sub>B<sub>12</sub>H<sub>12</sub>, undergoes an orderā€“disorder phase transition near 640 K. These results provide valuable structural information pertinent to understanding the potential role that Li<sub>2</sub>B<sub>10</sub>H<sub>10</sub> plays during LiBH<sub>4</sub> dehydrogenationā€“rehydrogenation as well as its prospects as a superionic Li<sup>+</sup> cation conductor

    Anion Reorientations in the Superionic Conducting Phase of Na<sub>2</sub>B<sub>12</sub>H<sub>12</sub>

    No full text
    Quasielastic neutron scattering (QENS) methods were used to characterize the reorientational dynamics of the dodecahydro-<i>closo</i>-dodecaborate (B<sub>12</sub>H<sub>12</sub><sup>2ā€“</sup>) anions in the high-temperature, superionic conducting phase of Na<sub>2</sub>B<sub>12</sub>H<sub>12</sub>. The icosahedral anions in this disordered cubic phase were found to undergo rapid reorientational motions, on the order of 10<sup>11</sup> jumps s<sup>ā€“1</sup> above 530 K, consistent with previous NMR measurements and neutron elastic-scattering fixed-window scans. QENS measurements as a function of the neutron momentum transfer suggest a reorientational mechanism dominated by small-angle jumps around a single axis. The results show a relatively low activation energy for reorientation of 259 meV (25 kJ mol<sup>ā€“1</sup>)

    Stable Interface Formation between TiS<sub>2</sub> and LiBH<sub>4</sub> in Bulk-Type All-Solid-State Lithium Batteries

    No full text
    In this study, we assembled a bulk-type all-solid-state battery comprised of a TiS<sub>2</sub> positive electrode, LiBH<sub>4</sub> electrolyte, and Li negative electrode. Our battery retained high capacity over 300 dischargeā€“charge cycles when operated at 393 K and 0.2 C. The second discharge capacity was as high as 205 mAh g<sup>ā€“1</sup>, corresponding to a TiS<sub>2</sub> utilization ratio of 85%. The 300th discharge capacity remained as high as 180 mAh g<sup>ā€“1</sup> with nearly 100% Coulombic efficiency from the second cycle. Negligible impact of the exposure of LiBH<sub>4</sub> to atmospheric-pressure oxygen on battery cycle life was also confirmed. To investigate the origin of the cycle durability for this bulk-type all-solid-state TiS<sub>2</sub>/Li battery, electrochemical measurements, thermogravimetry coupled with gas composition analysis, powder X-ray diffraction measurements, and first-principles molecular dynamics simulations were carried out. Chemical and/or electrochemical oxidation of LiBH<sub>4</sub> occurred at the TiS<sub>2</sub> surface at the battery operating temperature of 393 K and/or during the initial charge. During this oxidation reaction of LiBH<sub>4</sub> with hydrogen (H<sub>2</sub>) release just beneath the TiS<sub>2</sub> surface, a third phase, likely including Li<sub>2</sub>B<sub>12</sub>H<sub>12</sub>, precipitated at the interface between LiBH<sub>4</sub> and TiS<sub>2</sub>. Li<sub>2</sub>B<sub>12</sub>H<sub>12</sub> has a lithium ionic conductivity of logĀ­(Ļƒ / S cm<sup>ā€“1</sup>) = āˆ’4.4, charge transfer reactivity with Li electrodes, and superior oxidative stability to LiBH<sub>4</sub>, and thereby can act as a stable interface that enables numerous dischargeā€“charge cycles. Our results strongly suggest that the creation of such a stable interfacial layer is due to the propensity of forming highly stable, hydrogen-deficient polyhydro-<i>closo</i>-polyborates such as Li<sub>2</sub>B<sub>12</sub>H<sub>12</sub>, which are thermodynamically available in the ternary Liā€“Bā€“H system

    Impact of Ionic Liquid Pretreatment Conditions on Cellulose Crystalline Structure Using 1ā€‘Ethyl-3-methylimidazolium Acetate

    No full text
    Ionic liquids (ILs) have been shown to affect cellulose crystalline structure in lignocellulosic biomass during pretreatment. A systematic investigation of the swelling and dissolution processes associated with IL pretreatment is needed to better understand cellulose structural transformation. In this work, 3ā€“20 wt % microcrystalline cellulose (Avicel) solutions were treated with 1-ethyl-3-methylimidazolium acetate ([C<sub>2</sub>mim]Ā­[OAc]) and a mixture of [C<sub>2</sub>mim]Ā­[OAc] with the nonsolvent dimethyl sulfoxide (DMSO) at different temperatures. The dissolution process was slowed by decreasing the temperature and increasing cellulose loading, and was further retarded by addition of DMSO, enabling in-depth examination of the intermediate stages of dissolution. Results show that the cellulose I lattice expands and distorts prior to full dissolution in [C<sub>2</sub>mim]Ā­[OAc] and that upon precipitation the former structure leads to a less ordered intermediate structure, whereas fully dissolved cellulose leads to a mixture of cellulose II and amorphous cellulose. Enzymatic hydrolysis was more rapid for the intermediate structure (crystallinity = 0.34) than for cellulose II (crystallinity = 0.54)

    Orderā€“Disorder Transitions and Superionic Conductivity in the Sodium <i>nido</i>-Undeca(carba)borates

    No full text
    The salt compounds NaB<sub>11</sub>H<sub>14</sub>, Na-7-CB<sub>10</sub>H<sub>13</sub>, Li-7-CB<sub>10</sub>H<sub>13</sub>, Na-7,8-C<sub>2</sub>B<sub>9</sub>H<sub>12</sub>, and Na-7,9-C<sub>2</sub>B<sub>9</sub>H<sub>12</sub> all contain geometrically similar, monocharged, <i>nido</i>-undecaĀ­(carba)Ā­borate anions (i.e., truncated icosohedral-shaped clusters constructed of only 11 instead of 12 {Bā€“H} + {Cā€“H} vertices and an additional number of compensating bridging and/or terminal H atoms). We used first-principles calculations, X-ray powder diffraction, differential scanning calorimetry, neutron vibrational spectroscopy, neutron elastic-scattering fixed-window scans, quasielastic neutron scattering, and electrochemical impedance measurements to investigate their structures, bonding potentials, phase-transition behaviors, anion orientational mobilities, and ionic conductivities compared to those of their <i>closo</i>-polyĀ­(carba)Ā­borate cousins. All exhibited orderā€“disorder phase transitions somewhere between room temperature and 375 K. All disordered phases appear to possess highly reorientationally mobile anions (> āˆ¼10<sup>10</sup> jumps s<sup>ā€“1</sup> above 300 K) and cation-vacancy-rich, close-packed or body-center-cubic-packed structures [like previously investigated <i>closo</i>-polyĀ­(carba)Ā­borates]. Moreover, all disordered phases display superionic conductivities but with generally somewhat lower values compared to those for the related sodium and lithium salts with similar monocharged 1-CB<sub>9</sub>H<sub>10</sub><sup>ā€“</sup> and CB<sub>11</sub>H<sub>12</sub><sup>ā€“</sup> <i>closo</i>-carbaborate anions. This study significantly expands the known toolkit of solid-state, polyĀ­(carba)Ā­borate-based salts capable of superionic conductivities and provides valuable insights into the effect of crystal lattice, unit cell volume, number of carbon atoms incorporated into the anion, and charge polarization on ionic conductivity
    corecore