O Interactions in the Formation of Templated Vanadium Tellurites

Aug 8, 2011 - in the Formation of Templated Vanadium Tellurites. Matthew D. Smith,. †. Samuel M. Blau,. †. Kelvin B. Chang,. †. Matthias Zeller,...
0 downloads 0 Views 4MB Size
ARTICLE pubs.acs.org/crystal

Beyond Charge Density Matching: The Role of C H 3 3 3 O Interactions in the Formation of Templated Vanadium Tellurites Matthew D. Smith,† Samuel M. Blau,† Kelvin B. Chang,† Matthias Zeller,‡ Joshua Schrier,† and Alexander J. Norquist*,† † ‡

Department of Chemistry, Haverford College, Haverford, Pennsylvania 19041, United States Department of Chemistry, Youngstown State University, Youngstown, Ohio 44555, United States

bS Supporting Information ABSTRACT: Differences in the [V2Te2O10]n2n layer topologies in [C6H16N2][V2Te2O10] and [C5H14N2][V2Te2O10] are the result of deviations in the C H 3 3 3 O hydrogen-bonding networks. Stronger influences, such as reagent concentrations and charge density matching, were minimized through the use of both nearly identical reaction gels and 2,5-dimethylpiperazine and 2-methylpiperaine, which have very similar charge densities. Reactant concentration and charge density matching effects were quantified using composition space analyses, molecular and geometric decomposition surface areas, and Iterative Hirshfeld partial atomic charges.

’ INTRODUCTION Templated metal oxides have attracted sustained interest for many years, owing to their remarkably high degrees of compositional and structural diversity1 and for technologically desirable physical properties.2 However, increasing demand for new materials with enhanced properties has highlighted the lack of design and predictability in the syntheses of such compounds. The greatest limitation lies in our poor understanding of the mechanisms by which these compounds form.3 While mechanisms have been postulated,4 8 true design remains elusive. Significant progress has been made in the identification of reaction parameters that most strongly influence the formation of these materials. The primary influence over the structure of the inorganic component is reagent composition. Differences in reactant concentrations are known to directly affect the identity and availability of the primary building units from which the inorganic components are constructed.9 14 Reactant concentrations are clearly affected by a range of experimental parameters, ranging from the dependence of metal speciation on both pH and temperature to differences associated with source materials and reaction times. The manner in which the inorganic reactants oligomerize and polymerize in the systems described above is thought to be controlled by charge density matching4,5 between the organic cations and the anionic inorganic synthons. This secondary influence allows for crystallization only after the charge densities of the cationic and anionic components match. The importance of charge density matching has been demonstrated in a range of systems, including silicates,15 17 oxovanadium phosphates,18 gallium phosphates,4,5 molybdates,19 and vanadium tellurites.20 r 2011 American Chemical Society

Despite the utility of the charge density matching approach, differences in reactant concentrations and amine pKa alone do not wholly dictate the connectivities and structures of the resulting inorganic architectures. Charge density matching cannot differentiate between markedly different inorganic structures that have nearly identical charge densities. For example, we have reported the formation of both [Mo3O10]n2n and [Mo8O26]n4n chains from reactions in which the respective amines had similar pKa's and were used in nearly identical concentrations.19 We have also observed both [V2Te2O10]n2n chains and [V2TeO8]n2n layers in reactions containing either 1,4-diaminobutane or 1,3diaminopropane, respectively.20 In each system, the differences in charge densities of the inorganic components are small. As such, a series of tertiary influences has been proposed, including amine symmetry and hydrogen-bonding preferences.19 21 This report contains an elucidation and observation of tertiary influences in the formation of new organically templated vanadium tellurites. The NaVO3/Na2TeO3/2,5-dimethylpiperazine and NaVO3/Na2TeO3/2-methylpiperazine systems were explored using composition space analysis, resulting in the synthesis of two new compounds, [C6H16N2][V2Te2O10] (1) and [C5H14N2][V2Te2O10] (2). Charge density matching was quantified using Iterative Hirshfeld partial atomic charges,22,23 electron localization functions (ELFs), and both molecular24,25 and geometric decomposition19,20 surface areas. These systems were designed to minimize differences associated with composition Received: June 30, 2011 Revised: August 3, 2011 Published: August 08, 2011 4213

dx.doi.org/10.1021/cg2008232 | Cryst. Growth Des. 2011, 11, 4213–4219

Crystal Growth & Design and charge density matching in order to observe tertiary influences more directly.

ARTICLE

Table 1. Crystallographic Data for [C6H16N2][V2Te2O10] (1) and [C5H14N2][V2Te2O10] (2) compound

’ EXPERIMENTAL SECTION Materials. NaVO3 (99.5%), Na2TeO3 (99.5%), 2,5-dimethylpiperazine (2,5-dmpip, 98%), and 2-methylpiperazine (2-mpip, 95%) were purchased from Aldrich and used as received. Deionized water was used in these syntheses. Synthesis. All reactions were conducted in 23 mL poly(fluoroethylene-propylene)-lined pressure vessels. The pH of each reaction gel was adjusted to 8 using 4 M HCl and stirred for 10 min before heating at 90 °C for 84 h. The reactions were then cooled to room temperature at a rate of 6° h 1 to promote the growth of large single crystals. The pressure vessels were opened in air, and products were recovered through vacuum filtration. [C6H16N2][V2Te2O10] (1) was synthesized through the reaction of 0.0751 g (6.16  10 4 mol) of NaVO3, 0.1368 g (6.47  10 4 mol) of Na2TeO3, 0.1448 g (1.27  10 3 mol) of 2,5-dmpip, and 6.008 g (0.334 mol) of deionized water. Yellow blocks were produced in 80% yield, based upon tellurium. No other reactions products were observed from this reaction gel. Elemental microanalysis for 1 obsd (calc): C 11.47(11.37); H 2.59(2.53); N 4.45(4.42). IR data: N H 1428, 1484, 1608 cm 1, C H 2974 cm 1, Te O 715 cm 1, Te O Te 470, 610 cm 1, V = O 801 cm 1. [C5H14N2][V2Te2O10] (2) was synthesized through the reaction 0.0767 g (6.29  10 4 mol) of NaVO3, 0.1377 g (6.51  10 4 mol) of Na2TeO3, 0.1412 g (1.41  10 3 mol) of 2-mpip, and 6.033 g (0.335 mol) of deionized water. Yellow rods were produced in 40% yield, based upon tellurium. No other reactions products were observed from this reaction gel. Elemental microanalysis for 2 obsd (calc): C 9.59(9.69); H 2.21(2.26); N 4.42(4.52). IR data: N H 1432, 1488, 1613 cm 1, C H 2967 cm 1, Te O 738 cm 1, Te O Te 444, 658 cm 1, V = O 804 cm 1. Single Crystal X-ray Diffraction. Data were collected using a Bruker AXS Smart Apex CCD diffractometer with Mo KR radiation (λ = 0.71073 Å). A single crystal was mounted on a Mitegen micromesh mount using a trace of mineral oil and cooled in situ to 100(2) K for data collection. Frames were collected, indexed, and processed, and the files were scaled and corrected for absorption using APEX2.26 The heavy atom positions were determined using SIR92.27 All other non-hydrogen sites were located from Fourier difference maps. All non-hydrogen sites were refined using anisotropic thermal parameters using full matrix leastsquares procedures on Fo2 with I > 3σ(I). Hydrogen atoms were placed in geometrically idealized positions. All calculations were performed using Crystals.28 Relevant crystallographic data are listed in Table 1. Powder X-ray Diffraction. Powder diffraction patterns were recorded on a GBC-Difftech MMA powder diffractometer. Samples were mounted on aluminum plates. Calculated powder patterns were generated from single crystal data using ATOMS v. 6.0.29 Infrared Spectroscopy. Infrared measurements were obtained using a Perkin-Elmer FT-IR Spectrum 1000 spectrophotometer. Samples were diluted with spectroscopic grade KBr and pressed into pellets. Scans were run over the range of 400 4000 cm 1. Thermogravimetic Analyses. Thermogravimetric analyses (TGA) were conducted using a Q500 thermogravimetric analyzer from TA Instruments. Samples were contained within a platinum crucible and heated in nitrogen at 10 °C min 1 to 950 °C. TGA traces are available in the Supporting Information. Electronic Structure Calculations. Solid-state electronic structure calculations were performed using ABINIT v6.4.1,30,31 using the Perdew Burke Ernzerhof generalized gradient approximation (PBEGGA) exchange-correlation functional, norm-conserving TrollierMartins pseudopotentials, a planewave basis set with energy cutoff of

[C6H16N2][V2Te2O10] (1)

[C5H14N2][V2Te2O10] (2)

formula

C6H16N2O10Te2V2

C5H14N2O10Te2V2

fw space group

633.28 P1 (No. 2)

619.26 C2/c (No. 15)

a/Å

6.7486(17)

19.356(3)

b/Å

7.1572(18)

7.0591(12)

c/Å

8.687(2)

13.582(2)

R/°

69.818(3)

90

β/°

80.019(3)

125.4851(17)

γ/°

85.524(3)

90

V/Å3 Z

387.79(17) 1

1511.1(4) 4

Fcalc/g cm

3

2.712

2.722

0.71073

0.71073

100(2)

100(2)

4.942

5.069

R1a

0.0192

0.0161

wR2b

0.0429

λ/Å T/K μ/mm

a

1

R1 = Σ||Fo|

|Fc||/Σ|Fo|. b wR2 = [Σw(Fo2

0.0407

Fc2)2/[Σw(Fo2)2]1/2.

25 hartree, and utilizing the experimental crystal structures. Sampling of the Brillouin zone was performed with a 6  6  6 Monkhorst-Pack grid. Te pseudopotentials were calculated both with and without 4d-electrons in the valence. Electron Localization Functions (ELF). ELFs were computed from the self-consistent valence electron density using Te pseudopotentials without 4d-electrons in the valence. The valence-electrondensity ELF is quantitatively different from the all-electron result, but the only qualitative difference is the lack of core basins.32 The ELFs for [C6H16N2][V2Te2O10] (1) and [C5H14N2][V2Te2O10] (2) were calculated on 60  64  75 and 180  64  120 grids, respectively. ELFs were visualized using Vesta v1.1.033 with an isosurface value of 0.96. Surface Area Calculations. Surface areas for the inorganic components in [C6H16N2][V2Te2O10] (1) and [C5H14N2][V2Te2O10] (2) were calculated using the DMS program,34 defined by Richards24 as the molecular surface, and a geometric decomposition method that we reported earlier.19,20 Lone pair positions were defined as local maxima in the ELF isosurfaces, with radii of 1.5 Å, based upon Galy’s work.35,36 Calculated surface areas are listed in Table 2. Values are reported for a single [V2Te2O10]2 formula unit. Iterative Hirshfeld Charges. Iterative Hirshfeld (Hirshfeld-I)22,23 atomic partial charge determinations were performed on the self-consistent valence electron density in conjunction with all-electron atomic charge densities generated using the HF96 atomic Hartree Fock code,37 as described in our previous work,38 in which the Te pseudopotentials include 4d-electrons in the valence. The exclusion of 4d-electrons results in the same qualitative results, with small differences for the partial atomic charges on the order of 9% being observed. Full tables of partial atomic charges are provided in the Supporting Information, while [V2Te2O10]n2n charges are listed in Table 2.

’ RESULTS AND DISCUSSION Compound 1 is constructed from [V2Te2O10]n2n layers and [2,5-dmpipH2]2+ cations, as shown in Figure 1. The [V2Te2O10]n2n layers contain isolated [VO4] tetrahedra and [Te2O8] dimers. The [2,5-dmpipH2]2+ cations reside between [V2Te2O10]n2n layers and participate in an extensive hydrogen-bonding network. 4214

dx.doi.org/10.1021/cg2008232 |Cryst. Growth Des. 2011, 11, 4213–4219

Crystal Growth & Design

ARTICLE

Table 2. Surface Area and Charge Density Results for Compounds 1 and 2 compound

anion

[C6H16N2][V2Te2O10] (1)

[V2Te2O10]n2n

[C5H14N2][V2Te2O10] (2)

[V2Te2O10]n2n

surface area (Å2)

method molecular surface

107.81

geometric decomposition

106.39

molecular surface

119.23

geometric decomposition

111.48

Compound 2 also contains [V2Te2O10]n2n layers built from [VO4] tetrahedra and [Te2O8] dimers (see Figure 2). [2mpipH2]2+ cations lie between [V2Te2O10]n2n layers and form the basis of a three-dimensional hydrogen-bonding network. Strong similarities are observed between the inorganic components in 1 and 2. The V Oterminal bonds are shorter than V Obridging bonds, with observed ranges of 1.6125(18) 1.620(2) Å, and 1.6553(16) 1.8242(15) Å, respectively. Similarly, the Te Oterminal bonds are also shorter than Te Obridging bonds, with ranges of 1.8156(16) 1.8174(17) Å and 1.9197(15) 2.710(6) Å, respectively. The local vanadium coordination environments are both tetrahedral, while the tellurium centers in 1 and 2 exhibit distorted square pyramidal geometries. While neighboring [TeO5] polyhedra share two common oxides to form [Te2O8] dimers, no V O V bridges are observed. Connection of the [Te2O8] dimers and [VO4] tetrahedra results in [V2Te2O10]n2n layers in both 1 and 2, as shown in Figures 1a c and 2a c. Within each layer, the [Te2O8] dimers contain stereoactive lone pairs that are antialigned with respect to one another. An extensive three-dimensional hydrogen bonding network exists between the [2,5-dmpipH2]2+ and [2-mpipH2]2+ cations and adjacent [V2Te2O10]n2n layers. Three-dimensional packing figures for 1 and 2 are shown in Figures 1d and 2d, respectively. The bonding networks present in 1 and 2 were analyzed using bond valence sums.39,40 The calculated ΣSi values for each cation correspond to their formal oxidation states, with 5.04 and 5.06 sums for V5+ and 4.11 and 4.16 sums for Te4+. Full tables of bond valence sums are available in the Supporting Information. Despite the similarities between the structures of 1 and 2, notable differences in both the organic and inorganic components exist. First, the [Te2O8] dimers are arranged in a regular tiling in 1, while a herringbone pattern is observed in 2 (see Figures 1a and 2a). In addition, all [VO4] tetrahedra are rotated in the same direction by 3.5° about the a-axis in 1. The rotation of [VO4] tetrahedra in 2 is significantly larger, at 12.1° about the c-axis, with adjacent tetrahedra alternating between clockwise and counter-clockwise rotations. Second, the orientations of the [2,5-dmpipH2]2+ cations mirror the tessellation of the [Te2O8] dimers in compound 1. The [2-mpipH2]2+ cation orientations mirror the inorganic layers in 2 and exhibit a herringbone pattern (see Figure 3). The [(R)-2-mpipH2]2+ and [(S)-2-mpipH2]2+ cations in 2 are disordered over the same site with the imposition of a center of inversion; a partial ordering scheme present along the c-axis is discussed in the Supporting Information. It is interesting to note that while the crystallographic models of the [2,5-dmpipH2]2+ and [2-mpipH2]2+ cations are nearly identical in 1 and 2, the [V2Te2O10]n2n layer topologies differ. Understanding why the topologies of the [V2Te2O10]n2n layers in compounds 1 and 2 differ requires an analysis of the influences involved in their formations. Reactant concentrations and charge density matching are thought to be primary and secondary influences, respectively. As such, an investigation of these influences was performed.

anion charge 1.1554

charge density (e Å 2) 0.01086 0.01072

1.2308

0.01104 0.01032

Reactant concentrations are known to directly affect product composition in hydrothermal or solvothermal syntheses of templated metal oxides,9 14 which can impact charge density matching between the organic and inorganic components. Composition space analysis9 13,41,42 is an effective way to probe such effects. Specifically, composition space diagram were constructed for the NaVO3/Na2TeO3/2,5-dimethylpiperazine and NaVO3/ Na2TeO3/2-methylpiperazine systems, as shown in Figure 4, to allow for the direct observation of effects associated with V5+, Te4+ and amine concentrations. The total number of moles of the amine, vanadium and tellurium are held constant across each system, and the compositions of the crystalline products are plotted as functions of the reactant mole fractions. One templated vanadium tellurite is observed in each system. The crystallization fields of these two compounds are found in approximately the same positions. The TeO2 and NaVTeO5 crystallization fields at higher Te4+ mole fractions preclude the formation of other more tellurium rich templated vanadium tellurites. The identical vanadium tellurite layer compositions reflect the nearly identical positions of the compound 1 and 2 crystallization fields. As 1 and 2 were both synthesized from reaction gels with equivalent compositions, the differences in product structure should be aminedirected. The formation of organically templated metal oxides is thought to involve charge density matching between the cationic organic amines and anionic inorganic structures.4,5 The formation of neutral ammonium secondary building unit pairs allows for infinite condensation and crystallization. The charge densities of the organic amines are generally fixed by pKa, while oligomerization of the inorganic components allows for variable charge densities. The role of charge density matching in the formation of 1 and 2 is evident in the similarities observed between both their organic and inorganic components. The pKa's of 2,5-dmpip and 2-mpip are nearly identical, with values of 9.66 and 5.20 for 2,5dmpip and 9.57 and 5.24 for 2-mpip. The surface areas of the [V2Te2O10]n2n layers in 1 and 2 reflect these small differences (see Table 2). The charge densities of the [V2Te2O10]n2n layers in 1 and 2 are close to one another, with ranges of 0.01072 to 0.01086 and 0.01032 to 0.1104 e Å 2, respectively. These differences are small with respect to other systems in which the organic amine acidities show wider variations,19,20 suggesting that charge density matching does not dictate the difference between [V2Te2O10]n2n layer topologies in compounds 1 and 2. While charge density matching is useful for understanding the macroscopic properties of metal oxide anions, specific connectivities and topologies can be more directly influenced by other tertiary interactions. In the context of compounds 1 and 2, amine volumes and packing efficiencies appear to dictate the differences in [V2Te2O10]n2n topologies that are described above. The change in amine packing arrangements between 1 and 2 is a result of differences in amine volume. A shift from the larger [2,5dmpipH2]2+ to the smaller [2-mpipH2]2+ cations forces a change to the herringbone tessellation described above, which allows the 4215

dx.doi.org/10.1021/cg2008232 |Cryst. Growth Des. 2011, 11, 4213–4219

Crystal Growth & Design

Figure 1. (a c) Three views of the [V2Te2O10]n2n layers, with ELF isosurfaces shown with a boundary condition of 0.96, and (d) the threedimensional packing in [C6H16N2][V2Te2O10] (1). Green and red spheres represent tellurium and oxygen sites, while orange tetrahedra represent [VO4]. Black arrows indicate the direction of [VO4] rotation.

[2-mpipH2]2+ cations to move closer to one another in 2 with respect to the [2,5-dmpipH2]2+ cations in 1. The ab face of the compound 1 unit cell contains one complete [2,5-dmpipH2]2+ cation, with an area of 48.15 Å2 per cation. The bc face of the compound 2 unit cell contains two complete [2-mpipH2]2+ cations, with an area of 47.93 Å2 per cation (see Figure 3). In addition, the orientations of the [2,5-dmpipH2]2+ and [2-mpipH2]2+ cations

ARTICLE

Figure 2. (a c) Three views of the [V2Te2O10]n2n layers, with ELF isosurfaces shown with a boundary condition of 0.96, and (d) the threedimensional packing in [C5H14N2][V2Te2O10] (2). Green and red spheres represent tellurium and oxygen sites, while orange tetrahedra represent [VO4]. Black arrows indicate the direction of [VO4] rotation.

vary. The [2,5-dmpipH2]2+ cations are canted by 8.34°, with respect to the [V2Te2O10]n2n layers. In contrast, the [2-mpipH2]2+ cations in 2 are aligned more directly with the adjacent [V2Te2O10]n2n layers and are canted by only 1.65° (see Figure 5). These changes allow for the [2-mpipH2]2+ to pack more efficiently, reflecting their lower volumes, with respect to [2,5-dmpipH2]2+. 4216

dx.doi.org/10.1021/cg2008232 |Cryst. Growth Des. 2011, 11, 4213–4219

Crystal Growth & Design

ARTICLE

Figure 3. (a) [2,5-dmpipH2]2+ and (b) [2-mpipH2]2+ packing arrangements in 1 and 2, respectively. Blue and white spheres represent nitrogen and carbon. The disorder in the [2-mpipH2]2+ cations is shown in (b). Hydrogen atoms have been removed for clarity.

Figure 5. Hydrogen-bonding interactions between (a) [2,5-dmpipH2]2+ and (b) [2-mpipH2]2+ cations and adjacent [V2Te2O10]n2n layers in 1 and 2, respectively. Selected C H 3 3 3 O bond angles are shown (°).

Figure 4. Composition space diagrams for the (a) NaVO3/Na2TeO3/ 2,5-dmpip and (b) NaVO3/Na2TeO3/2-mpip systems.

The changes in the cation packing arrangements and orientations directly affect the [V2Te2O10]n2n layer topologies in two main ways involving C H 3 3 3 O hydrogen bonding. While

shorter, stronger N H 3 3 3 O hydrogen bonds are also present in 1 and 2, their orientations preclude direct influence over the orientations of the [VO4] tetrahedra. Instead, the C H 3 3 3 O hydrogen bonds play a pivotal role in determining the [VO4] orientations, as summarized in Table 3. The primary C H 3 3 3 O hydrogen bond acceptors in compound 1 are O1 and O2, while in compound 2, O3 also accepts a hydrogen bond (see Figure 5). This difference is caused by rotation of [2-mpipH2]2+, as discussed above, and results in a shift in the O3 position toward the [2-mpipH2]2+ cation. It should be noted that the C3 H9 3 3 3 O1 interaction in 2, denoted by a black line in Figure 5b, is short, with a C3 3 3 3 O1 distance of 3.081(6) Å. However, its angle is far too acute at 108° to involve any significant bonding, as described in a recent C H 3 3 3 X bonding review.43 Also, the hydrogen-bonding interactions between [2,5dmpipH2]2+ or [2-mpipH2]2+ and O1 are distinctly different (see Figure 6). The four interactions observed in O1 in 1 are relatively isotropic in their directions. In contrast, the directions 4217

dx.doi.org/10.1021/cg2008232 |Cryst. Growth Des. 2011, 11, 4213–4219

Crystal Growth & Design

ARTICLE

Table 3. Selected Hydrogen-Bonding Interactions in Compounds 1 and 2 [C6H16N2][V2Te2O10] (1)

D 3 3 3 A distance (Å)

D H 3 3 3 A angle (°)

N1 H1 3 3 3 O5 N1 H2 3 3 3 O5 C2 H4 3 3 3 O1 C3 H8 3 3 3 O1

2.859(4)

151

2.683(4) 3.166(4)

175 139

3.350(4)

134

C2* H5 3 3 3 O1 C3* H7 3 3 3 O1

3.195(4)

129

3.340(4)

125

[C5H14N2][V2Te2O10] (2)

D 3 3 3 A distance (Å)

D H 3 3 3 A angle (°)

N1 H1 3 3 3 O5 N1 H2 3 3 3 O5 C2 H6 3 3 3 O1 C3 H8 3 3 3 O1

2.694(6)

166

2.770(6)

152

3.182(6) 3.492(6)

123 121

C2* H5 3 3 3 O1 C3* H9 3 3 3 O1

3.105(6)

102

3.081(6)

108

two compounds with similar stoichiometries from nearly identical reaction gels and the use of two amines whose charge densities are nearly equal prohibits differences associated with reactant concentration or charge density matching, respectively, between systems and allows for the direct observation of the tertiary influence of hydrogen bonding.

’ CONCLUSIONS Differences in the C H 3 3 3 O hydrogen-bonding networks in [C6H16N2][V2Te2O10] (1) and [C5H14N2][V2Te2O10] (2) are the source of deviations in the respective [V2Te2O10]n2n layer topologies. Observation of these tertiary influences is only possible because deviations in the effects associated with reactant concentrations and charge density matching are small in the NaVO3/Na2TeO3/2,5-dimethylpiperazine and NaVO 3/Na2TeO3/2-methylpiperazine systems. ’ ASSOCIATED CONTENT

bS

Supporting Information. Tables of bond valence sums, calculated Iterative Hirshfeld partial atomic charges, [2-mpipH2]2+ disorder mechanism in 2, figures of the compound 2 model used in all electronic structure calculations, geometric decomposition models and thermogravimetric traces for 1 and 2. An X-ray crystallographic information file (CIF) is available for [C6H16N2][V2Te2O10] (1) and [C5H14N2][V2Te2O10] (2). This material is available free of charge via the Internet at http://pubs.acs.org.

’ AUTHOR INFORMATION Corresponding Author

*Address: 370 Lancaster Avenue, Haverford, PA 19041. Tel (610) 896 2949. Fax (610) 896 4963. E-mail: anorquis@ haverford.edu. Web: http://www.haverford.edu/chem/Norquist/.

Figure 6. Hydrogen-bonding interactions in (a) 1 and (b) 2.

of the two C H 3 3 3 O1 hydrogen-bonding interactions in 2 are markedly anisotropic. Much like the C3* H9 3 3 3 O1 angle described above, the C2* H5 3 3 3 O1 angle of 102° is too acute to be meaningful. The resulting asymmetry in these interactions, coupled with the inclusion of the C3 H9 3 3 3 O3 interaction, results in a distinct rotation of the [VO4] in the direction of the C2 H6 3 3 3 O1 and C3 H8 3 3 3 O1 bonds. It is the rotation of the [VO4] tetrahedra, coupled with the herringbone arrangement of the [2-mpipH2]2+ cations, which causes the change in layer topology between 1 and 2. The determination that hydrogen-bonding alone is responsible for the differences in [V2Te2O10]n2n layer topologies between 1 and 2 is only possible because stronger effects have been equalized between systems. Specifically, the synthesis of

’ ACKNOWLEDGMENT The authors acknowledge support from the NSF (Award No. CHE-0911121), the Henry Dreyfus Teacher-Scholar Awards Program, and grants to Haverford College from the HHMI Undergraduate Science Education Program. M.Z. acknowledges support for the purchase of a diffractometer from the NSF Grant 0087210, the Ohio Board of Regents Grant CAP-491 and from Youngstown State University. This research used resources of the National Energy Research Scientific Computing Center, which is supported by the Office of Science of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231. ’ REFERENCES (1) Cheetham, A. K.; Ferey, G.; Loiseau, T. Angew. Chem., Int. Ed. 1999, 38, 3268–3292. (2) Haag, W. O. Stud. Surf. Sci. Catal. 1994, 84, 1375–1394. (3) Yu, J.; Xu, R. Acc. Chem. Res. 2010, 43, 1195–1204. (4) Ferey, G. J. Fluorine Chem. 1995, 72, 187–193. (5) Ferey, G. Chem. Mater. 2001, 13, 3084–3098. (6) Rao, C. N. R.; Natarajan, S.; Choudhury, A.; Neeraj, S.; Ayi, A. A. Acc. Chem. Res. 2001, 34, 80–87. (7) Rao, C. N. R.; Dan, M.; Behera, J. N. Pure Appl. Chem. 2005, 77, 1655–1674. (8) Murugavel, R.; Walawalkar, M. G.; Dan, M.; Roesky, H. W.; Rao, C. N. R. Acc. Chem. Res. 2004, 37, 763–774. 4218

dx.doi.org/10.1021/cg2008232 |Cryst. Growth Des. 2011, 11, 4213–4219

Crystal Growth & Design (9) Halasyamani, P.; Willis, M. J.; Stern, C. L.; Lundquist, P. M.; Wong, G. K.; Poeppelmeier, K. R. Inorg. Chem. 1996, 35, 1367–1371. (10) Norquist, A. J.; Heier, K. R.; Stern, C. L.; Poeppelmeier, K. R. Inorg. Chem. 1998, 37, 6495–6501. (11) Thomas, P. M.; Norquist, A. J.; Doran, M. B.; O’Hare, D. J. Mater. Chem. 2003, 13, 88–92. (12) Veltman, T. R.; Stover, A. K.; Narducci Sarjeant, A.; Ok, K. M.; Halasyamani, P. S.; Norquist, A. J. Inorg. Chem. 2006, 45, 5529–5537. (13) Nelson, J. H.; Johnston, A. R.; Narducci Sarjeant, A.; Norquist, A. J. Solid State Sci. 2007, 9, 472–484. (14) Norquist, A. J.; Doran, M. B.; Thomas, P. M.; O’Hare, D. Dalton Trans. 2003, 1168–1175. (15) Monnier, A.; Schuth, F.; Huo, Q.; Kumar, D.; Margolese, D.; Maxwell, R. S.; Stucky, G. D.; Krishnamurty, M.; Petroff, P.; Firouzi, A.; Janicke, M.; Chmelka, B. F. Science 1993, 261, 1299–1303. (16) Huo, Q.; Margolese, D. I.; Ciesla, U.; Feng, P.; Gier, T. E.; Sieger, P.; Leon, R.; Petroff, P. M.; Schueth, F.; Stucky, G. D. Nature 1994, 368, 317–321. (17) Tolbert, S. H.; Landry, C. C.; Stucky, G. D.; Chmelka, B. F.; Norby, P.; Hanson, J. C.; Monnier, A. Chem. Mater. 2001, 13, 2247–2256. (18) El Haskouri, J.; Roca, M.; Cabrera, S.; Alamo, J.; Beltran-Porter, A.; Beltran-Porter, D.; Marcos, M. D.; Amoros, P. Chem. Mater. 1999, 11, 1446–1454. (19) Casalongue, H. S.; Choyke, S. J.; Narducci Sarjeant, A.; Schrier, J.; Norquist, A. J. J. Solid State Chem. 2009, 182, 1297–1303. (20) Chang, K. B.; Hubbard, D. J.; Zeller, M.; Schrier, J.; Norquist, A. J. Inorg. Chem. 2010, 49, 5167–5172. (21) Norquist, A. J.; Thomas, P. M.; Doran, M. B.; O’Hare, D. Chem. Mater. 2002, 14, 5179–5184. (22) Bultinck, P.; Van Alsenoy, C.; Ayers, P. W.; Carbo-Dorca, R. J. Chem. Phys. 2007, 126, 144111. (23) Bultinck, P.; Ayers, P. W.; Fias, S.; Tiels, K.; Van Alsenoy, C. Chem. Phys. Lett. 2007, 444, 205–208. (24) Richards, F. M. Annu. Rev. Biophys. Bioeng. 1977, 6, 151–176. (25) Greer, J.; Bush, B. L. Proc. Natl. Acad. Sci. U. S. A. 1978, 75, 303–307. (26) Apex2 v2009.7-0; Bruker AXS Inc.: Madison (WI), USA, 2009. (27) Altomare, A.; Cascarano, G.; Giacovazzo, C.; Guagliardi, A. J. Appl. Crystallogr. 1993, 26, 343–350. (28) Betteridge, P. W.; Carruthers, J. R.; Cooper, R. I.; Prout, K.; Watkin, D. J. J. Appl. Crystallogr. 2003, 36, 1487. (29) Dowty, E. ATOMS v. 6.0; Shape Software: Kingsport, TN, 2002. (30) Gonze, X.; Beuken, J. M.; Caracas, R.; Detraux, F.; Fuchs, M.; Rignanese, G. M.; Sindic, L.; Verstraete, M.; Zerah, G.; Jollet, F.; Torrent, M.; Roy, A.; Mikami, M.; Ghosez, P.; Raty, J. Y.; Allan, D. C. Comput. Mater. Sci. 2002, 25, 478–492. (31) Gonze, X.; Amadon, B.; Anglade, P. M.; Beuken, J. M.; Bottin, F.; Boulanger, P.; Bruneval, F.; Caliste, D.; Caracas, R.; Cote, M.; Deutsch, T.; Genovese, L.; Ghosez, P.; Giantomassi, M.; Goedecker, S.; Hamann, D. R.; Hermet, P.; Jollet, F.; Jomard, G.; Leroux, S.; Mancini, M.; Mazevet, S.; Oliveira, M. J. T.; Onida, G.; Pouillon, Y.; Rangel, T.; Rignanese, G. M.; Sangalli, D.; Shaltaf, R.; Torrent, M.; Verstraete, M. J.; Zerah, G.; Zwanziger, J. W. Comput. Phys. Commun. 2009, 180, 2582–2615. (32) Kohout, M.; Savin, A. J. Comput. Chem. 1997, 18, 1431–1439. (33) Momma, K.; Izumi, F. J. Appl. Crystallogr. 2008, 41, 653–658. (34) Huang, C. DMS, Computer Graphics Lab; University of California: San Francisco, CA, USA, 2002. (35) Andersson, S.; Astrom, A.; Galy, J.; Meunier, G. J. Solid State Chem. 1973, 6, 187–190. (36) Galy, J.; Meunier, G.; Andersson, S.; Astrom, A. J. Solid State Chem. 1975, 13, 142–159. (37) Gaigalas, G.; Froese Fischer, C. Comput. Phys. Commun. 1996, 98, 255–264. (38) Glor, E. C.; Blau, S. M.; Yeon, J.; Zeller, M.; Shiv Halasyamani, P.; Schrier, J.; Norquist, A. J. J. Solid State Chem. 2011, 184, 1445–1450. (39) Brown, I. D.; Altermatt, D. Acta Crystallogr. Sect. B 1985, 41, 244–247.

ARTICLE

(40) Brese, N. E.; O’Keeffe, M. Acta Crystallogr. Sect. B 1991, 47, 192–197. (41) Harrison, W. T. A.; Dussack, L. L.; Jacobson, A. J. J. Solid State Chem. 1996, 125, 234–242. (42) Doran, M. B.; Norquist, A. J.; O’Hare, D. Inorg. Chem. 2003, 42, 6989–6995. (43) Desiraju, G. R. Acc. Chem. Res. 2002, 35, 565–573.

4219

dx.doi.org/10.1021/cg2008232 |Cryst. Growth Des. 2011, 11, 4213–4219