5 research outputs found

    Jahn–Teller Effect in the B<sub>12</sub>F<sub>12</sub> Radical Anion and Energetic Preference of an Octahedral B<sub>6</sub>(BF<sub>2</sub>)<sub>6</sub> Cluster Structure over an Icosahedral Structure for the Elusive Neutral B<sub>12</sub>F<sub>12</sub>

    No full text
    The B<sub>12</sub>F<sub>12</sub><sup>–</sup> radical anion was generated by oxidation of [CoCp<sub>2</sub><sup>+</sup>]<sub>2</sub>B<sub>12</sub>F<sub>12</sub><sup>2–</sup> with AsF<sub>5</sub> in SO<sub>2</sub>. In the crystal structure of [CoCp<sub>2</sub><sup>+</sup>]­B<sub>12</sub>F<sub>12</sub><sup>–</sup>, the anion displays a lowered symmetry (<i>D</i><sub>2<i>h</i></sub>) instead of an <i>I<sub>h</sub></i>-symmetric dianion as a result of Jahn–Teller distortion. Moreover, shortening of the B–F bonds and subtle changes of the B–B bonds are observed. DFT calculations show that, for the unknown neutral B<sub>12</sub>F<sub>12</sub>, unprecedented structural isomers [e.g., octahedral B<sub>6</sub>(BF<sub>2</sub>)<sub>6</sub>] are energetically favored instead of an icosahedral structure. The structures and energetics are compared with those of the analogous chlorine compounds

    Jahn–Teller Effect in the B<sub>12</sub>F<sub>12</sub> Radical Anion and Energetic Preference of an Octahedral B<sub>6</sub>(BF<sub>2</sub>)<sub>6</sub> Cluster Structure over an Icosahedral Structure for the Elusive Neutral B<sub>12</sub>F<sub>12</sub>

    No full text
    The B<sub>12</sub>F<sub>12</sub><sup>–</sup> radical anion was generated by oxidation of [CoCp<sub>2</sub><sup>+</sup>]<sub>2</sub>B<sub>12</sub>F<sub>12</sub><sup>2–</sup> with AsF<sub>5</sub> in SO<sub>2</sub>. In the crystal structure of [CoCp<sub>2</sub><sup>+</sup>]­B<sub>12</sub>F<sub>12</sub><sup>–</sup>, the anion displays a lowered symmetry (<i>D</i><sub>2<i>h</i></sub>) instead of an <i>I<sub>h</sub></i>-symmetric dianion as a result of Jahn–Teller distortion. Moreover, shortening of the B–F bonds and subtle changes of the B–B bonds are observed. DFT calculations show that, for the unknown neutral B<sub>12</sub>F<sub>12</sub>, unprecedented structural isomers [e.g., octahedral B<sub>6</sub>(BF<sub>2</sub>)<sub>6</sub>] are energetically favored instead of an icosahedral structure. The structures and energetics are compared with those of the analogous chlorine compounds

    Structures of M<sub>2</sub>(SO<sub>2</sub>)<sub>6</sub>B<sub>12</sub>F<sub>12</sub> (M = Ag or K) and Ag<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub>B<sub>12</sub>F<sub>12</sub>: Comparison of the Coordination of SO<sub>2</sub> versus H<sub>2</sub>O and of B<sub>12</sub>F<sub>12</sub><sup>2–</sup> versus Other Weakly Coordinating Anions to Metal Ions in the Solid State

    No full text
    The structures of three solvated monovalent cation salts of the superweak anion B<sub>12</sub>F<sub>12</sub><sup>2–</sup> (Y<sup>2–</sup>), K<sub>2</sub>(SO<sub>2</sub>)<sub>6</sub>Y, Ag<sub>2</sub>­(SO<sub>2</sub>)<sub>6</sub>Y, and Ag<sub>2</sub>­(H<sub>2</sub>O)<sub>4</sub>Y, are reported and discussed with respect to previously reported structures of Ag<sup>+</sup> and K<sup>+</sup> with other weakly coordinating anions. The structures of K<sub>2</sub>(SO<sub>2</sub>)<sub>6</sub>Y and Ag<sub>2</sub>­(SO<sub>2</sub>)<sub>6</sub>Y are isomorphous and are based on expanded cubic close-packed arrays of Y<sup>2–</sup> anions with M­(OSO)<sub>6</sub><sup>+</sup> complexes centered in the trigonal holes of one expanded close-packed layer of B<sub>12</sub> centroids (⊙). The K<sup>+</sup> and Ag<sup>+</sup> ions have virtually identical bicapped trigonal prism MO<sub>6</sub>F<sub>2</sub> coordination spheres, with M–O distances of 2.735(1)–3.032(2) Å for the potassium salt and 2.526(5)–2.790(5) Å for the silver salt. Each M­(OSO)<sub>6</sub><sup>+</sup> complex is connected to three other cationic complexes through their six μ-SO<sub>2</sub>-κ<sup>1</sup><i>O</i>,κ<sup>2</sup><i>O</i>′ ligands. The structure of Ag<sub>2</sub>­(H<sub>2</sub>O)<sub>4</sub>Y is unique [different from that of K<sub>2</sub>­(H<sub>2</sub>O)<sub>4</sub>Y]. Planes of close-packed arrays of anions are offset from neighboring planes along only one of the linear ⊙···⊙···⊙ directions of the close-packed arrays, with [Ag­(μ-H<sub>2</sub>O)<sub>2</sub>­Ag­(μ-H<sub>2</sub>O)<sub>2</sub>)]<sub>∞</sub> infinite chains between the planes of anions. There are two nearly identical AgO<sub>4</sub>F<sub>2</sub> coordination spheres, with Ag–O distances of 2.371(5)–2.524(5) Å and Ag–F distances of 2.734(4)–2.751(4) Å. This is only the second structurally characterized compound with four H<sub>2</sub>O molecules coordinated to a Ag<sup>+</sup> ion in the solid state. Comparisons with crystalline H<sub>2</sub>O and SO<sub>2</sub> solvates of other Ag<sup>+</sup> and K<sup>+</sup> salts of weakly coordinating anions show that (i) N­[(SO<sub>2</sub>)<sub>2</sub>­(1,2-C<sub>6</sub>H<sub>4</sub>)]<sup>−</sup>, BF<sub>4</sub><sup>–</sup>, SbF<sub>6</sub><sup>–</sup>, and Al­(OC­(CF<sub>3</sub>)<sub>3</sub>)<sub>4</sub><sup>–</sup> coordinate much more strongly to Ag<sup>+</sup> than does Y<sup>2–</sup>, (ii) SnF<sub>6</sub><sup>2–</sup> coordinates somewhat more strongly to K<sup>+</sup> than does Y<sup>2–</sup>, and (iii) B<sub>12</sub>Cl<sub>12</sub><sup>2–</sup> coordinates to K<sup>+</sup> about the same as, if not slightly weaker than, Y<sup>2–</sup>

    Comparison of the Coordination of B<sub>12</sub>F<sub>12</sub><sup>2–</sup>, B<sub>12</sub>Cl<sub>12</sub><sup>2–</sup>, and B<sub>12</sub>H<sub>12</sub><sup>2–</sup> to Na<sup>+</sup> in the Solid State: Crystal Structures and Thermal Behavior of Na<sub>2</sub>(B<sub>12</sub>F<sub>12</sub>), Na<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub>(B<sub>12</sub>F<sub>12</sub>), Na<sub>2</sub>(B<sub>12</sub>Cl<sub>12</sub>), and Na<sub>2</sub>(H<sub>2</sub>O)<sub>6</sub>(B<sub>12</sub>Cl<sub>12</sub>)

    No full text
    The synthesis of high-purity Na<sub>2</sub>B<sub>12</sub>F<sub>12</sub> and the crystal structures of Na<sub>2</sub>(B<sub>12</sub>F<sub>12</sub>) (5 K neutron powder diffraction (NPD)), Na<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub>(B<sub>12</sub>F<sub>12</sub>) (120 K single-crystal X-ray diffraction (SC-XRD)), Na<sub>2</sub>(B<sub>12</sub>Cl<sub>12</sub>) (5 and 295 K NPD), and Na<sub>2</sub>(H<sub>2</sub>O)<sub>6</sub>­(B<sub>12</sub>Cl<sub>12</sub>) (100 K SC-XRD) are reported. The compound Na<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub>(B<sub>12</sub>F<sub>12</sub>) contains {[(Na­(μ-H<sub>2</sub>O)<sub>2</sub>Na­(μ-H<sub>2</sub>O)<sub>2</sub>)]<sup>2+</sup>}<sub>∞</sub> infinite chains; the compound Na<sub>2</sub>(H<sub>2</sub>O)<sub>6</sub>­(B<sub>12</sub>Cl<sub>12</sub>) contains discrete [(H<sub>2</sub>O)<sub>2</sub>Na­(μ-H<sub>2</sub>O)<sub>2</sub>Na­(H<sub>2</sub>O)<sub>2</sub>]<sup>2+</sup> cations with OH···O hydrogen bonds linking the terminal H<sub>2</sub>O ligands. The structures of the two hydrates and the previously published structure of Na<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub>(B<sub>12</sub>H<sub>12</sub>) are analyzed with respect to the relative coordinating ability of B<sub>12</sub>F<sub>12</sub><sup>2–</sup>, B<sub>12</sub>H<sub>12</sub><sup>2–</sup>, and B<sub>12</sub>Cl<sub>12</sub><sup>2–</sup> toward Na<sup>+</sup> ions in the solid state (i.e., the relative ability of these anions to satisfy the valence of Na<sup>+</sup>). All three hydrated structures have distorted octahedral NaX<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub> coordination spheres (X = F, H, Cl). The sums of the four Na–O bond valence contributions are 71, 75, and 89% of the total bond valences for the X = F, H, and Cl hydrated compounds, respectively, demonstrating that the relative coordinating ability by this criterion is B<sub>12</sub>Cl<sub>12</sub><sup>2–</sup> ≪ B<sub>12</sub>H<sub>12</sub><sup>2–</sup> < B<sub>12</sub>F<sub>12</sub><sup>2–</sup>. Differential scanning calorimetry experiments demonstrate that Na<sub>2</sub>(B<sub>12</sub>F<sub>12</sub>) undergoes a reversible, presumably order–disorder, phase transition at ca. 560 K (287 °C), between the 529 and 730 K transition temperatures previously reported for Na<sub>2</sub>(B<sub>12</sub>H<sub>12</sub>) and Na<sub>2</sub>(B<sub>12</sub>Cl<sub>12</sub>), respectively. Thermogravimetric analysis demonstrates that Na<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub>­(B<sub>12</sub>F<sub>12</sub>) and Na<sub>2</sub>(H<sub>2</sub>O)<sub>6</sub>­(B<sub>12</sub>Cl<sub>12</sub>) undergo partial dehydration at 25 °C to Na<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>­(B<sub>12</sub>F<sub>12</sub>) and Na<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>­(B<sub>12</sub>Cl<sub>12</sub>) in ca. 30 min and 2 h, respectively, and essentially complete dehydration to Na<sub>2</sub>(B<sub>12</sub>F<sub>12</sub>) and Na<sub>2</sub>(B<sub>12</sub>Cl<sub>12</sub>) within minutes at 150 and 75 °C, respectively (the remaining trace amounts of H<sub>2</sub>O, if any, were not quantified). The changes in structure upon dehydration and the different vapor pressures of H<sub>2</sub>O needed to fully hydrate the respective Na<sub>2</sub>(B<sub>12</sub>X<sub>12</sub>) compounds provide additional evidence that B<sub>12</sub>Cl<sub>12</sub><sup>2–</sup> is more weakly coordinating than B<sub>12</sub>F<sub>12</sub><sup>2–</sup> to Na<sup>+</sup> in the solid state. Taken together, the results suggest that the anhydrous, halogenated <i>closo</i>-borane compounds Na<sub>2</sub>(B<sub>12</sub>F<sub>12</sub>) and Na<sub>2</sub>(B<sub>12</sub>Cl<sub>12</sub>), in appropriately modified forms, may be viable component materials for fast-ion-conducting solid electrolytes in future energy-storage devices

    Latent Porosity in Alkali-Metal M<sub>2</sub>B<sub>12</sub>F<sub>12</sub> Salts: Structures and Rapid Room-Temperature Hydration/Dehydration Cycles

    No full text
    Structures of the alkali-metal hydrates Li<sub>2</sub>(H<sub>2</sub>O)<sub>4</sub>Z, LiK­(H<sub>2</sub>O)<sub>4</sub>Z, Na<sub>2</sub>(H<sub>2</sub>O)<sub>3</sub>Z, and Rb<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>Z, unit cell parameters for Rb<sub>2</sub>Z and Rb<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>Z, and the density functional theory (DFT)-optimized structures of K<sub>2</sub>Z, K<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>Z, Rb<sub>2</sub>Z, Rb<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>Z, Cs<sub>2</sub>Z, and Cs<sub>2</sub>(H<sub>2</sub>O)­Z are reported (Z<sup>2–</sup> = B<sub>12</sub>F<sub>12</sub><sup>2–</sup>) and compared with previously reported X-ray structures of Na<sub>2</sub>(H<sub>2</sub>O)<sub>0,4</sub>Z, K<sub>2</sub>(H<sub>2</sub>O)<sub>0,2,4</sub>Z, and Cs<sub>2</sub>(H<sub>2</sub>O)­Z. Unusually rapid room-temperature hydration/dehydration cycles of several M<sub>2</sub>Z/M<sub>2</sub>(H<sub>2</sub>O)<sub><i>n</i></sub>Z salt hydrate pairs, which were studied by isothermal gravimetry, are also reported. Finely ground samples of K<sub>2</sub>Z, Rb<sub>2</sub>Z, and Cs<sub>2</sub>Z, which are not microporous, exhibited latent porosity by undergoing hydration at 24–25 °C in the presence of 18 Torr of H<sub>2</sub>O­(g) to K<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>Z, Rb<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>Z, and Cs<sub>2</sub>(H<sub>2</sub>O)­Z in 18, 40, and 16 min, respectively. These hydrates were dehydrated at 24–25 °C in dry N<sub>2</sub> to the original anhydrous M<sub>2</sub>Z compounds in 61, 25, and 76 min, respectively (the exact times varied from sample to sample depending on the particle size). The hydrate Na<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>Z also exhibited latent porosity by undergoing multiple 90 min cycles of hydration to Na<sub>2</sub>(H<sub>2</sub>O)<sub>3</sub>Z and dehydration back to Na<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>Z at 23 °C. For the K<sub>2</sub>Z, Rb<sub>2</sub>Z, and Cs<sub>2</sub>Z transformations, the maximum rate of hydration (rh<sub>max</sub>) decreased, and the absolute value of the maximum rate of dehydration (rd<sub>max</sub>) increased, as <i>T</i> increased. For K<sub>2</sub>Z ↔ K<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>Z hydration/dehydration cycles with the same sample, the ratio rh<sub>max</sub>/rd<sub>max</sub> decreased 26 times over 8.6 °C, from 3.7 at 23.4 °C to 0.14 at 32.0 °C. For Rb<sub>2</sub>Z ↔ Rb<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>Z cycles, rh<sub>max</sub>/rd<sub>max</sub> decreased from 0.88 at 23 °C to 0.23 at 27 °C. For Cs<sub>2</sub>Z ↔ Cs<sub>2</sub>(H<sub>2</sub>O)­Z cycles, rh<sub>max</sub>/rd<sub>max</sub> decreased 20 times over 8 °C, from 6.7 at 24 °C to 0.34 at 32 °C. In addition, the reversible substitution of D<sub>2</sub>O for H<sub>2</sub>O in fully hydrated Rb<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>Z in the presence of N<sub>2</sub>/16 Torr of D<sub>2</sub>O­(g) was complete in only 60 min at 23 °C
    corecore