MAGNETIC RESONANCE IN CHEMISTRY Magn. Reson. Chem. 2007; 45: S116–S128 Published online in Wiley InterScience (www.interscience.wiley.com) DOI: 10.1002/mrc.2124

Investigation of structure and dynamics in the sodium metallocenes CpNa and CpNa·THF via solid-state NMR, X-ray diffraction and computational modelling Cory M. Widdifield, Joel A. Tang, Charles L. B. Macdonald and Robert W. Schurko∗ Department of Chemistry and Biochemistry, University of Windsor, Windsor, Ontario, Canada N9B 3P4 Received 16 July 2007; Revised 18 September 2007; Accepted 24 September 2007

Solid-state 23 Na NMR spectra of two organometallic complexes, cyclopentadienylsodium (CpNa) and the tetrahydrofuran (THF) solvate of CpNa (CpNa·THF), are presented. Analytical simulations of experimental spectra and calculated 23 Na electric-field gradient (EFG) tensors confirm that both complexes are present in microcrystalline samples of CpNa recrystallized from THF. For the solvate, 23 Na NMR experiments at 9.4 T and 11.7 T elucidate sodium chemical shielding (CS) tensor parameters, and establish that the EFG and CS tensor frames are non-coincident. Single-crystal X-ray diffraction (XRD) experiments are used to ˚ b = 20.611.3/ A, ˚ c = 9.8236.14/ A, ˚ determine the crystal structure of CpNa·THF: Cmca (a = 9.3242.15/ A, ˚ 3 , Z = 8). For CpNa, 23 Na NMR data acquired at multiple field strengths a = b = g = 90°, V = 1887.9.5/A establish sodium CS tensor parameters more precisely than in previous reports. Variable-temperature (VT) powder XRD (pXRD) experiments determine the temperature dependence of the CpNa unit cell parameters. The combination of 23 Na quadrupolar NMR parameters, pXRD data and calculations of 23 Na EFG tensors is used to examine various models of dynamic motion in the solid state. It is proposed that the sodium atom in CpNa undergoes an anisotropic, temperature-dependent, low frequency motion within the ab crystallographic plane, in contrast with previous models. Copyright  2007 John Wiley & Sons, Ltd. Supplementary electronic material for this paper is available in Wiley InterScience at http://www.interscience.wiley.com/ jpages/0749-1581/suppmat/

KEYWORDS: solid-state NMR; 13 C NMR; 23 Na NMR; metallocenes; ab initio; chemical shielding; CSA; EFG; XRD; CP/MAS

INTRODUCTION Approximately 50 years prior to the initial synthesis1 and characterization2,3 of ferrocene (Cp2 Fe, Cp D C5 H5 , cyclopentadienyl) in the early 1950s, the first alkalimetal metallocenes, cyclopentadienylsodium (CpNa)4 and cyclopentadienylpotassium (CpK),5 were prepared. Numerous studies have dealt with alkali-metal containing metallocenes due to their utility as starting reagents in organometallic syntheses.6 – 10 These metallocenes also possess a wide range of structural motifs, including solvated11 – 14 and unsolvated15 – 17 polymeric chains, solvated oligomers18,19 and monomers,20,21 and may also be constituents in mixedmetal metallocenes22 – 24 and anionic sandwich complexes.25,26 Although a consensus was reached early on regarding the ionic nature of alkali-metal metallocenes (i.e. a positively charged metal centre interacting with negatively charged Cp rings), it is only recently that their actual molecular structures have been elucidated.27 – 32 Owing to substantial advances in solid-state NMR hardware and characterization

Ł Correspondence to: Robert W. Schurko, Department of Chemistry and Biochemistry, University of Windsor, Windsor, Ontario, Canada N9B 3P4. E-mail: [email protected]

Copyright  2007 John Wiley & Sons, Ltd.

techniques,33 – 37 we have been able to characterize both structure and dynamics in several alkali-metal metallocenes.38,39 For a series of Cp0 Na complexes (Cp0 D C5 H5y Ry ; R D Me, iPr, etc. y D 0–5), we previously used solid-state 23 Na and 13 C NMR spectroscopy and ab initio calculations to investigate (i) metallocene structure, (ii) Cp0 ring dynamics, (iii) temperature-dependent structural behaviour and (iv) sodium chemical shielding (CS) and electric-field gradient (EFG) tensors.38 In addition, a base-substituted species, CpNaÐTHF (THF D tetrahydrofuran), was identified and its structure was proposed on the basis of available solid-state 13 C and 23 Na NMR data and preliminary computational modelling. This is noteworthy as single crystals for many polymeric alkali-metal metallocenes are extremely difficult to isolate and their structures have largely been deduced on the basis of synchrotron powder X-ray diffraction (pXRD) data and subsequent Rietveld analyses.40 Herein, we extend the discussion on CpNa and CpNaÐTHF, primarily using solid-state 23 Na NMR, for three reasons. First, CpNa, which is air- and moisture-sensitive, is an integral component in many organometallic syntheses and has recently been reported to act as an initiator in a living photopolymerization process.41,42 Second, solid-state 23 Na NMR plays an increasingly important role in understanding

Investigation of structure and dynamics in sodium metallocenes

cation- interactions, which occur in a plenitude of biological systems, and inorganic and organometallic materials. The Cp0 Na series of complexes are excellent model systems to help one understand the relationships between molecular structure, dynamic motion and 23 Na chemical shifts and quadrupolar parameters. Lastly, CpNa and CpNaÐTHF are often found as an inhomogeneous microcrystalline mixture (vide infra); hence, it is of great interest to have rapid and accurate methods for the identification of these species as well as for any impurities which result from decomposition. To this end, we report a combination of solid-state NMR spectroscopy, single-crystal X-ray diffraction (XRD), variable-temperature (VT) pXRD and ab initio calculations in an attempt to characterize structure and dynamics in CpNa and CpNaÐTHF. A single-crystal XRD structure for CpNaÐTHF is reported. The sodium EFG and CS tensors at the sodium atom in CpNaÐTHF are determined experimentally, and the previously observed38 sodium chemical shielding anisotropy (CSA) in CpNa is defined more precisely using solid-state 23 Na NMR data acquired at several magnetic field strengths. Finally, VT solid-state 23 Na NMR data and VT pXRD experiments in combination with detailed computational modelling are used to propose a model for temperature-dependent structural dynamics in CpNa.

EXPERIMENTAL CpNa·THF crystal isolation A 2.0 M solution of CpNa dissolved in THF was purchased from Aldrich. While under nitrogen, a portion of this solution was decanted into a vial. In order to effect the slow, incomplete, evaporation of THF, this vial was placed inside a more voluminous container containing liquid THF that was capped and left to equilibrate for 2 weeks. A crystal of CpNaÐTHF, suitable for single-crystal XRD experiments, was isolated from the vial of mixed CpNa and CpNaÐTHF. Suitable crystals for CpNa could not be isolated.

Single-crystal XRD A colourless crystal with dimensions of 0.50 ð 0.40 ð 0.40 mm was selected, packed into a capillary tube while under a nitrogen atmosphere and then flame-sealed. The raw XRD data were collected at T D 1732 K, using a ˚ graphite-monochromated radiation Mo-K˛ ( D 0.71073 A) source with a Bruker APEX CCD area detector running the SMART data collection software.43 The crystal produced the following unit cell parameters: C9 H13 NaO; Fw D 160.18; ˚ b D 20.6113 A, ˚ orthorhombic; Cmca (a D 9.324215 A, ˚ ˛ D ˇ D  D 90° , V D 1887.95 A ˚ 3, c D 9.823614 A, Z D 8); Dc D 1.127 g cm3 . 1826 total reflections, collected over the range 2.86 <  < 27.47° , were used to determine the unit cell parameters. For formal data collection, a total of 8587 partial and complete reflections were collected over the range 1.98 <  < 27.57° , covering the indices h D 12 to 12, k D 26 to 26 and l D 12 to 12. A collection strategy of both ω and  scans was employed for both cell determination and data collection steps. Of the reflections collected, 1162 were symmetry independent. Data reduction was done using

Copyright  2007 John Wiley & Sons, Ltd.

SAINT.44 The structure was solved for using direct methods with SHELXTL45 and refined (full-matrix least squares on F2 ) using SHELXL-97.46 Overall quality factors (of the 659 reflections with I > 2 I) of the refined structure are: R1 D 0.0639 and wR2 D 0.1332 with a GooF of 1.079.

Powder XRD Microcrystalline samples of CpNa and mixed CpNa/ CpNaÐTHF were prepared under nitrogen atmosphere and were each flame-sealed in 1.0 mm capillary tubes. Data were collected at several temperature points ranging from T D 1231 K to 323(1) K using a D8 DISCOVER powder X-ray diffractometer equipped with an Oxford Cryosystems 700 series Cryostream Plus cooler. This diffractometer uses ˚ radiation source with a Bruker a Cu-K˛  D 1.54056 A AXS HI-STAR area detector running under the General Area Detector Diffractions System. XRD powder patterns were simulated using PowderCell 2.4.47

Solid-state NMR Samples of CpNa and CpNa/CpNaÐTHF, the latter from the same vial that produced the single crystal used in XRD experiments, were ground into microcrystalline powders under nitrogen atmosphere, packed tightly into 4 mm o.d. zirconia rotors, and sealed with airtight caps. Spectra were obtained using 9.4 T [v0 1 H ³ 400 MHz] wide bore Varian InfinityC and 11.7 T [v0 1 H ³ 500 MHz] standard bore Bruker AMX-500 NMR spectrometers at the University of Windsor, and a 21.1 T [v0 1 H ³ 900 MHz] standard bore Bruker AVANCE II NMR spectrometer at the Canadian National Ultrahigh-field NMR Facility for Solids in Ottawa. Varian/Chemagnetics double- and triple-resonance 4 mm magic-angle spinning (MAS) probes were used for all experiments conducted at B0 D 9.4 T, while Bruker double-resonance 4 mm probes were used for experiments conducted at higher fields. The central-transition (CT) selective pulse widths used for 23 Na experiments are equal to the non-selective solution pulse widths after scaling by a factor of I C 1/21 . 13

C CP/MAS NMR

Carbon chemical shifts were referenced to tetramethylsilane [TMS; υiso 13 C D 0.0 ppm], using the high-frequency signal of adamantane [υiso 13 C D 38.57 ppm] as an external secondary reference. Experiments were conducted at several MAS frequencies: 1.37, 2.41 and 4.4 kHz, and employed proton saturation, cross-polarization (CP) and decoupling field strengths, v1 1 H, of 82–104 kHz, ca 45–55 kHz, respectively. Optimized parameters include a CP contact time ranging from 9 to 12 ms and recycle delays between 9 and 15 s. A spectral width of 70.0 kHz was used and 110–328 transients were collected. 23

Na NMR

Sodium chemical shifts were referenced to 1.0 M NaCl [υiso 23 Na D 0.0 ppm]. A Hahn echo pulse sequence was used for all static and MAS (vrot D 10–11 kHz) experiments using rotor-synchronized data collection wherever applicable. To reduce the appearance of nutation distortions in the

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

S117

S118

C. M. Widdifield et al.

spectra,38,48 soft, CT-selective 23 Na /2 pulse widths ranging from 9.0 to 17.5 µs were used. Experiments typically employed a spectral width of 50 kHz, CW 1 H decoupling fields of 50 kHz, recycle delays of 4–6 s, and a collection of 400–2000 and 29 transients at standard and ultrahigh magnetic fields, respectively.

Spectral simulations 23

Na EFG and CS tensor parameters were determined using the analytical simulation program WSolids.49 In select cases, numerical simulations were carried out using SIMPSON.50 All SIMPSON simulations employed the direct method of powder averaging using the zcw4180 crystal file, with the start and detect operators set equal to I1z and I1c , respectively. The number of  angles was set to 20 for MAS simulations. All error bounds were determined by the bidirectional variation of individual NMR parameters and best-fit spectra were obtained by minimizing the RMS difference between simulated and experimental spectra.

The point charges selected for use in embedded cluster molecular orbital (ECMO) calculations were based on a Mulliken population analysis with the charges of the hydrogen atoms incorporated into the heavy atoms [qNa D C0.75 e ˚ and qC D 0.15 e]. A sphere of 170 point charges (¾20 A in diameter) was used and the individual point charges were translated relative to a central Cp2 Na3 C cluster in accordance with the observed changes in the unit cell (checked for consistency using PowderCell 2.4). When modelling temperature-dependent motions of the sodium atom in CpNa, a single Cp2 Na cluster was used. The sodium ˚ increments towards the centroid atom was moved in 0.03 A of one of the Cp rings to mimic motion along the c-axis of the unit cell. For displacements in the ab-plane, the sodium ˚ increments towards several difatom was moved in 0.02 A ferent fixed points. The points chosen correspond to the ab-coordinates of the Cp ring carbons and the C–C bond midpoints.

Theoretical calculations Atomic coordinates for CpNa and CpNaÐTHF were taken from the available crystal structures,17 and the hydrogen positions of CpNaÐTHF were optimized at the B3LYP/631GŁŁ level of theory. Calculations were restricted to small polymeric chains of CpNa ([Cp2 Na] /[Cp2 Na3 ]C ) and CpNaÐTHF ([Cp2 NaÐTHF] /[Cp2 Na3 Ð3THF]C / Cp3 Na3 Ð3THF/[Cp4 Na3 Ð3THF] ) (see Supporting Information, Tables S1 and S2 for optimized CpNaÐTHF atomic positions and SCF energies) and were performed using Dell Precision workstations running Red Hat Linux. CpNaÐTHF calculations were carried out on two conformers; however, as calculated SCF energies and NMR parameters did not differ significantly between conformers, results are reported for one conformer only. Gaussian 0351 calculations used either restricted Hartree-Fock (RHF) or hybrid density functional theory (DFT) methods using Becke’s three-parameter hybrid functional with the correlation functional proposed by Lee, Yang and Parr (B3LYP).52,53 All Gaussian calculations employed the standard polarized double- and triple- basis sets provided with the software (6-31GŁŁ /6-311GŁŁ /6-311CGŁŁ /6-311CCGŁŁ ) on all non-sodium atoms. Most calculations involved the usage of standard basis sets on the sodium atoms, although select calculations used a well-tempered basis set (WTBS).54,55 In order to convert the 23 Na quadrupolar coupling constant (CQ ) into MHz from atomic units,56,57 the formula CQ D eV33 Q/h ð 9.7177 ð 1021 V m2 was used, where V33 is the largest component of the EFG tensor in its own principal axis system and Q23 Na D 0.1054 ð 1028 m2 . This median value of Q23 Na was chosen so as to not favour any methodology for the observation of Q23 Na.58 – 62 Sodium and carbon chemical shielding tensors were calculated using the gauge-including atomic orbitals (GIAO) method63,64 and were referenced against the sodium magnetic shielding constant for infinitely dilute NaC ions in D2 O at 297 K [ iso 23 Na D 576.6 ppm ³ υiso 23 Na D 0.0 ppm], as described by Mason and Bryce et al.65,66 and the theoretical isotropic shift of CO [υiso 13 C D 187.1 ppm], respectively.

Copyright  2007 John Wiley & Sons, Ltd.

RESULTS AND DISCUSSION Crystal structure of CpNa·THF To date, only a handful of polymeric sodocene solvates have been structurally characterized using single-crystal XRD techniques due to difficulties in isolating single crystals of suitable size and/or quality. A number of these solvates possess Cp rings that contain large substituents as observed in (benzyl)CpNaÐTHF and (cyclopentyl)CpNaÐTHF,67 but only one example can be found where the Cp ring is unsubstituted: CpNaÐTMEDA (tetramethylethylenediamine).31,32 Our previous NMR studies on linear Cp0 Na complexes revealed the presence of a linear, zig-zagging, solvated species that consistently occurs in samples of CpNa recrystallized from THF.41 13 C NMR data identify the THF solvate and 23 Na MAS NMR spectra indicate a sodium atom coordinated by two Cp rings in a zig-zagging fashion. The refined crystal structure of CpNaÐTHF confirms our earlier hypothesis (Fig. 1), although the THF carbon atoms belong to two energetically identical conformers, each having an assigned occupancy of 50%. The sodium atom is 5 -coordinated to two Cp rings, forming a zig-zagging, chainlike polymeric structure [6 Cpcent –Na–Cpcent  D 143.4° , where Cpcent is the geometric centre of a Cp ring], similar to the structure reported for CpNaÐTMEDA. Both solvates have increased Cpcent –Na distances relative to CpNa: in CpNa, ˚ 17 while in CpNaÐTMEDA the Cpcent –Na distance is 2.357 A, ˚ respec˚ 31,32 and 2.455 A, and CpNaÐTHF they are 2.667 A tively. The increase in Cpcent –Na bond distance is due to the presence of the THF ligand that coordinates to the metal, which serves to increase the electron density at the sodium atom and electrostatically repulse the Cp ligands. This increase is modest compared to that for CpNaÐTMEDA, as the sodium–THF interaction is relatively weak31 and does not highly perturb the electronic configuration of the sodium valence.

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

Investigation of structure and dynamics in sodium metallocenes

Figure 1. (a) A rendering of CpNaÐTHF using SHELXTL.45 Thermal ellipsoids are shown at the 50% probability level. Both THF conformers are displayed: one is drawn using solid lines and the other using dotted lines. (b) View of the CpNaÐTHF unit cell along the a-crystallographic axis. (c) A view along the b-axis highlights the zig-zagging linear polymeric nature of the molecule. Hydrogen atoms have been omitted in (a–c) and only one THF conformer is shown in (c) to improve clarity.

Solid-state 23 Na NMR spectroscopy CpNa/CpNaÐTHF As we were unable to isolate a pure sample of microcrystalline CpNaÐTHF large enough for NMR experimentation, solid-state 23 Na MAS NMR experiments were conducted on a mixed CpNa/CpNaÐTHF sample. 23 Na MAS NMR spectra recorded at B0 D 9.4 T and 11.7 T depict two sodium powder patterns, each clearly dominated by the second-order quadrupolar interaction (Fig. 2). Simulations of the low- and high-frequency powder patterns reveal 23 Na EFG tensor and chemical shift parameters in accordance with previously published data for CpNa and CpNaÐTHF, respectively.38 Sodium CSA in CpNaÐTHF was not previously measured; hence, static 23 Na NMR experiments were also conducted on the mixed sample at two external magnetic fields (B0 D 9.4 and 11.7 T) (Fig. 3). The sodium chemical shift range is relatively small and there are few examples of sodium CSA in the current literature (the largest observed value for is ca 20 ppm).38,66,68 – 72 On the basis of analytical simulations of the powder patterns, it is established that for CpNaÐTHF is ca 20 ppm, with values of 18(3) ppm and 21(3) ppm taken

Copyright  2007 John Wiley & Sons, Ltd.

Figure 2. Solid-state 23 Na MAS NMR spectra of CpNa/CpNaÐTHF. Both numerical SIMPSON, (a), which accounts for finite pulse widths and sample rotation, and analytical WSolids, (b and e), simulations, which account for neither factor, closely match the experimental spectra acquired at B0 D 9.4 T (c) and B0 D 11.7 T (f). A de-convolution of the two sodium sites is provided in (d). Impurities are marked with asterisks. Inset: magnification of the high-frequency region of the experimental spectrum in (f), highlighting the impurity. Spinning sidebands are denoted with †.

from simulations at B0 D 9.4 and 11.7 T, respectively. In CpNaÐTHF, the CS tensor (B0 D 9.4 T/B0 D 11.7 T) deviates from axial symmetry [ D 0.32/0.22], and the EFG and CS tensor frames are non-coincident: ˛ D 75° 10/80° 10, ˇ D 80° 10/85° 10,  D 20° 10/27° 5. Further discussion is provided in the Theoretical Calculations section.

CpNa The effects of sodium CSA are very subtle in the 23 Na NMR spectra of CpNa; hence, 23 Na NMR experiments were conducted at 21.1 T to measure the sodium CS tensor parameters with greater accuracy. Previous solid-state 23 Na NMR experiments on CpNa at B0 D 9.4 T resolved a small, axiallysymmetric sodium CS tensor [ D 12.53.0/ D 1.0],38 where axial symmetry was assumed on the basis of molecular geometry. The account is very similar when one moves to a slightly higher magnetic field [B0 D 11.7 T; see Table 1 and Fig. 4(b)]; however, when compared to the values obtained at a substantially higher magnetic field (B0 D 21.1 T) (Table 1 and Fig. 4(d)), the CS tensor parameters now lie outside the error ranges associated with the moderate-field accounts. The CS tensor has a larger span [ D 17.51.5 ppm] and is not axially symmetric [ D 0.81]. The high-field data are

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

S119

S120

C. M. Widdifield et al.

Table 1. Experimental Sodium CS and EFG Tensor Parameters

Experiment CpNa·THF MAS e MAS f MAS e Static Static CpNa MAS e,g MAS f MAS e,g Static f Static g Static Static

B0 (T)

CQ (MHz)

Q a

9.4 1.82(2) 0.37(2) 9.4 1.82(2) 0.39(2) 11.7 1.84(3) 0.39(3) 9.4 – – 11.7 – –

b (ppm)

– – – 18(3) 21(3)

c

υiso d (ppm)

– 45.4(3) – 45.5(2) – 45.1(3) 0.3(2) – 0.2(2) –

9.4 2.97(3) 0.03(3) – – 57.6(3) 9.4 2.97(3) 0.02(2) – – 57.5(3) 11.7 3.00(3) 0.03(2) – – 57.0(3) 9.4 – – 12.5(3.0) 1.0 h – 11.7 – – 17(3) 0.9(1) – 11.7 – – 14(2) 0.9(1) 57.5(3) 21.2 – – 17.5(1.5) 0.8(1) 58.2(4)

a

EFG tensor asymmetry parameter, Q D V11  V22 /V33 . CS tensor span, D υ11  υ33 . c CS tensor skew,  D 3υ22  υiso / . d Isotropic chemical shift, υiso D υ11 C υ22 C υ33 /3. e Vrot D 10 kHz. f Taken from Ref. 38. g Present in a mixture with CpNaÐTHF. h Assumed on the basis of molecular symmetry. b

Figure 3. Static solid-state 23 Na NMR spectra of CpNa/CpNaÐTHF. Spectra acquired at (c) B0 D 9.4 T and (g) B0 D 11.7 T correspond closely to analytical simulations only if the effects of sodium CSA and non-coincident CS and EFG tensors are included (b and f). Simulations which do not include these effects for the CpNaÐTHF site (a and e) clearly do not match the experimental spectra. A deconvolution is provided in (d) and impurities are marked with asterisks.

expected to be more reliable, since a of 17.5 ppm accounts for ca 10 and 37% of the total powder pattern width at 9.4 and 21.1 T, respectively (Figs 3(c) and 4(d)).

Identification of impurity phases The mixed CpNa/CpNaÐTHF sample, which is light brown in colouration, will decompose rapidly upon exposure to air or heat to yield a dark brown decomposition product. In the static 23 Na NMR spectra of CpNa/CpNaÐTHF acquired at 11.7 T, there is a minor powder pattern with a centre of gravity at 9.6 ppm, which is likely the result of partial decomposition from oxidation or exposure to heat. The site cannot be resolved using MQMAS experiments perhaps due to disorder and/or a distribution of large quadrupolar interactions (see Supporting Information, Fig. S1 and Table S3). A CpNa/CpNaÐTHF sample was exposed to air and moisture in order to characterize the by-products of sample degradation. Solid-state 23 Na NMR spectra (Fig. S2) of the decomposed sample reveal a broad powder pattern centred at 1.6 ppm, indicating a sodium environment very similar to hydrated NaC [υiso 23 Na ³ 0 ppm]. Powder XRD data indicate a disordered decomposition product (vide infra), and 23 Na NMR powder patterns corresponding to CpNa and its solvate are absent. Occasionally, commercial samples of CpNa contain a substantial amount of this impurity phase, which is expected

Copyright  2007 John Wiley & Sons, Ltd.

to have a significant impact on synthetic yields for a number of organometallic syntheses. Checks for decomposition via pXRD can be deceiving, as the impurity phase is believed to be largely amorphous. Alternatively, simple 23 Na NMR experiments provide a rapid check of the purity of such samples. Further, an additional peak [υ D 68.55 ppm] is identified in Fig. 2, possibly arising from some minor impurity; however, due to the very low intensity of this resonance, no second-order pattern can be discerned and the chemical nature of this species cannot be ascertained.

Variable-temperature powder XRD CpNa VT powder XRD experiments were conducted upon pure CpNa and mixed CpNa/CpNaÐTHF samples, in order to examine structural changes as a function of temperature. The unit cell dimensions of CpNa change slightly as a function of temperature (Fig. S3) and can be fit to a function which is linear in temperature (Fig. 5). Unit cell changes were determined by observing the variation in the 2 values associated with the (200), (020) and (002) peaks, since the distance between Bragg planes corresponding to the same Miller index is given by: d D [h/a2 C k/b2 C l/c2 ]1/2 . On the basis of previous solid-state 23 Na NMR data and computational modelling, it was concluded that CpNa exhibits temperature-dependent structural changes, whereby the Cpcent –Na distance (and hence the c-axis of the unit cell) increases with decreasing temperature. However, the actual changes in unit cell parameters with temperature are not in line with this earlier proposal. There is no

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

Investigation of structure and dynamics in sodium metallocenes

Figure 4. Static solid-state 23 Na NMR spectra of CpNa at (a and b) B0 D 11.7 T and (c and d) B0 D 21.1 T. Bottom traces (b and d) correspond to experimental spectra with top traces (a and c) representing analytical simulations using the parameters outlined in Table 1. Dotted line traces in (a) and (c) are equivalent to the solid line traces, but neglect CSA. Impurities are marked with asterisks.

correlation between the c unit cell length and temperature (R2 D 0.0651; Fig. S4(a)), while the lengths of the a- and baxes show strong linear correlations with temperature: R2 D 0.9946 and 0.9834, respectively (Fig. 5). Several scenarios for temperature-dependent structural changes are discussed in the Theoretical Calculations section below.

CpNa/CpNaÐTHF The pXRD spectrum (Fig. S5(a)) of the mixed sample was expected to include several peaks in the same positions as the pXRD spectrum of CpNa (Fig. S3), with additional peaks attributed to CpNaÐTHF. However, the pXRD patterns acquired for several recrystallized samples of CpNa/ CpNaÐTHF do not show characteristic peaks associated with either CpNa (based upon its refined pXRD structure) or CpNaÐTHF (according to its single-crystal XRD structure). Solid-state 23 Na NMR spectra indicate an impurity phase, ˚ can be but only the broad peak at 2 D 3.60° d D 24.5 A attributed to decomposition of the sample upon exposure to air and moisture (Fig. S5(b)). Additional low-angle peaks (asterisks in Fig. S5(a)) showed variability in their intensities over a number of samples, and may correspond to sample decomposition prior to hydration. Since there is no 10.50

11.05

10.40

11.00

10.30 10.95

10.20 10.10

10.90

10.00

10.85

9.90 10.80

9.80 9.70 123

153

183 213 243 273 Temperature (K)

298

10.75 323

Figure 5. Plot of CpNa unit cell lengths as a function of temperature. The a  and b  unit cell lengths can both be fit to functions which are linear with respect to temperature: ˚ D 0.0020 TK C 9.5149; bA ˚ D 0.0008 TK C 10.7694. aA

Copyright  2007 John Wiley & Sons, Ltd.

correspondence between this pXRD pattern and those of pure CpNa and CpNaÐTHF, it is not possible to define the unit cell parameters of the mixed CpNa/CpNaÐTHF sample; hence, direct commentary upon solid-state structural changes as a function of temperature cannot be made. A preliminary conclusion is that the solid-state structure of the mixed sample cannot be represented by individual microcrystalline regions of CpNa and CpNaÐTHF. We hypothesize that the material may be composed of one or more polymorphic species (e.g. both constituents of the mixed sample crystallize such that their unit cell parameters in the mixture differ from when they are pure). 13

C CP/MAS NMR

Spectra of the mixed microcrystalline sample were obtained at two MAS frequencies, revealing three isotropic signals (Fig. 6). The isotropic resonance at υiso 13 C D 104.32 ppm corresponds to the Cp ring carbon atoms of both CpNa and CpNaÐTHF (Table 2). The resolution of the 13 C CP/MAS NMR experiments do not allow for their differentiation (Fig. 6, left inset), hinting that coordination of the THF ligand to the sodium atom does not greatly perturb the electronic environment in the Cp rings, in support of the modest structural changes outlined earlier. Resonances at ca 71.0(3)/69.6(3) ppm and 27.6(2) ppm are assigned to the ˛ and ˇ THF ring carbon atoms (Fig. 6, right inset), respectively. Herzfeld–Berger analysis of the powder pattern, which corresponds to the Cp ring carbons, highlights an axiallysymmetric shielding tensor [13 C D C1.0] of appreciable magnitude [ 13 C D 121.1 š 8.0 ppm]. While the span reported here for the Cp ring carbons of the mixed sample is slightly higher than observed in the pure CpNa complex,38 the errors associated with Herzfeld–Berger analysis reduce the significance of this discrepancy.

Theoretical calculations 23 Na EFG tensors Previous ab initio calculations of the 23 Na EFG tensor for CpNaÐTHF were carried out upon gas-phase, geometryoptimized structures,38 whereas the theoretical 23 Na EFG tensor parameters presented here are obtained from calculations

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

S121

S122

C. M. Widdifield et al.

Table 2. Experimental Carbon CS Tensor Parameters

Molecule

Site

υ11 (ppm)

υ22 (ppm)

υ33 (ppm)

υiso (ppm)

(ppm)



CpNaÐTHF

Cp ˛ ˇ Cp

144.7 – – 141.0

144.7 – – 138.8

23.5 – – 33.8

104.3(2) 71.0(3)/69.6(3) 27.6(2) 104.5

121.1(8.0) – – 107.1

1.0 – – 0.96

CpNa a a

Taken from Ref. 38.

Figure 6. 13 C CP/MAS NMR spectra of CpNa/CpNaÐTHF at (a) vrot D 4.4 kHz and (b) vrot D 1.37 kHz. The isotropic peak associated with the Cp ring carbons is denoted with an asterisk, carbon atoms in the THF ring are denoted with symbols, and correspond to the carbon atoms shown in the right inset. Left inset: closeup of the isotropic Cp ring carbon signal prior to the application of line broadening.

on molecular geometries obtained from the CpNaÐTHF crystal structure. The 23 Na EFG tensor orientation of CpNaÐTHF presented herein (Fig. 7(a)–(b)) does not agree with the earlier gas-phase calculations.38 V22 and V33 lie nearly in the bc-crystallographic plane, with V22 pointing near the sodium–oxygen bond axis and V11 nearly along the direction of chain propagation. The tensor orientation does not

Copyright  2007 John Wiley & Sons, Ltd.

vary significantly over the series of calculations, provided that moderately large basis sets (6-311GŁŁ or larger) are applied. Each principal component of the 23 Na EFG tensor is predicted to be oriented near a crystallographic axis, with 6 V11 –Na–a, 6 V22 –Na–c and 6 V33 –Na–b, equal to (standard deviation in parenthesis) 1.6(1.1)° , 3.2(0.7)° and 2.9(0.1)° , respectively, for the calculations in Tables 3, 4 and

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

Investigation of structure and dynamics in sodium metallocenes

Table 3. Experimental and Theoretical 23 Na EFG Tensor Parameters for CpNaÐTHFa

Cluster Experimental CpNaÐTHF CpNaÐTHF CpNaÐTHF RHF [Cp2 NaÐTHF] [Cp2 Na3 Ð3THF]C Cp3 Na3 Ð3THF [Cp4 Na3 Ð3THF] [Cp2 NaÐTHF] [Cp2 Na3 Ð3THF]C Cp3 Na3 Ð3THF [Cp4 Na3 Ð3THF] B3LYP [Cp2 NaÐTHF] [Cp2 Na3 Ð3THF]C Cp3 Na3 Ð3THF [Cp4 Na3 Ð3THF] [Cp2 NaÐTHF] [Cp2 Na3 Ð3THF]C Cp3 Na3 Ð3THF [Cp4 Na3 Ð3THF]

V11 b (au)

V22 (au)

V33 (au)

jCQ jc (MHz)

Q

– – –

– – –

– – –

1.82(2) 1.79 1.78

0.37(2) 0.61 0.70

6-311GŁŁ 6-311GŁŁ 6-311GŁŁ 6-311GŁŁ 6-311CGŁŁ 6-311CGŁŁ 6-311CGŁŁ 6-311CGŁŁ

0.0349 0.0037 0.0125 0.0228 0.0331 0.0052 0.0132 0.0245

0.0789 0.1167 0.1068 0.0961 0.0814 0.1165 0.1069 0.0963

0.1138 0.1204 0.1192 0.1189 0.1144 0.1218 0.1201 0.1207

2.675 2.828 2.802 2.793 2.689 2.861 2.822 2.837

0.387 0.939 0.791 0.617 0.422 0.914 0.780 0.595

6-311GŁŁ 6-311GŁŁ 6-311GŁŁ 6-311GŁŁ 6-311CGŁŁ 6-311CGŁŁ 6-311CGŁŁ 6-311CGŁŁ

0.0302 0.0005 0.0079 0.0176 0.0300 0.0021 0.0094 0.0207

0.0782 0.1165 0.1074 0.0977 0.0795 0.1159 0.1062 0.0954

0.1084 0.1170 0.1153 0.1153 0.1096 0.1179 0.1155 0.1161

2.547 2.750 2.709 2.709 2.575 2.771 2.714 2.728

0.444 0.991 0.863 0.694 0.452 0.965 0.838 0.644

Basis set

MAS-RT MAS-183 K MAS-153 K

a

Using standard (i.e. a WTBS was not employed) basis sets on all atoms. Vii are the principal components of the EFG tensor, where jV33 j ½ jV22 j ½ jV11 j. c Calculated CQ is converted from atomic units into MHz by multiplying V33 by (eQ/h)(9.7177 ð 1021 V m2 ), where Q23 Na D 0.1054 ð 1028 m2 . b

Table 4. Experimental and Theoretical 23 Na EFG Tensor Parameters for CpNaÐTHFa

Cluster Experimental CpNaÐTHF CpNaÐTHF CpNaÐTHF RHF [Cp2 NaÐTHF] [Cp2 Na3 Ð3THF]C Cp3 Na3 Ð3THF [Cp2 NaÐTHF] [Cp2 Na3 Ð3THF]C Cp3 Na3 Ð3THF a

Basis set

MAS-RT MAS-183 K MAS-153 K 6-311GŁŁ 6-311GŁŁ 6-311GŁŁ 6-311CGŁŁ 6-311CGŁŁ 6-311CGŁŁ

V11 (au)

V22 (au)

V33 (au)

jCQ j (MHz)

Q

– – –

– – –

– – –

1.82(2) 1.79 1.78

0.37(2) 0.61 0.70

0.0284 0.0159 0.0188 0.0193 0.0010 0.0031

0.0505 0.0637 0.0607 0.0501 0.0783 0.0726

0.0789 0.0796 0.0795 0.0694 0.0793 0.0757

1.854 1.870 1.867 1.631 1.864 1.778

0.279 0.601 0.527 0.443 0.975 0.918

Using a WTBS on the sodium atom(s).

5. For comparison, the orientation of the EFG tensor in CpNa is shown in Fig. 7(d),38 where the distinct V33 component is along the axis of chain propagation (the c-crystallographic axis), and V11 and V22 are oriented in electronically similar environments (hence, Q ³ 0), nearly parallel to the Cp rings. Calculated CQ values fall into two regimes, with calculations employing DZ and TZ basis sets predicting lower and higher relative values, respectively (Tables 3 and S4(a)). Although experimental CQ values seem to be best reproduced

Copyright  2007 John Wiley & Sons, Ltd.

using a DZ basis, additional considerations reduce one’s confidence in them. First, the 23 Na EFG tensor orientation varies significantly over all DZ calculations. Second, the range of calculated Q values is greater when comparing DZ calculations (0.103 to 0.890) to TZ calculations (0.386 to 0.939). While TZ calculations consistently predict nearly identical EFG tensor orientations, they overestimate experimental CQ values by approximately 50%. The choice of Q23 Na cannot fully account for this difference between

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

S123

S124

C. M. Widdifield et al.

Table 5. Experimental and Theoretical Sodium CS Tensor Parameters for CpNaÐTHF

Cluster Experimental CpNaÐTHF CpNaÐTHF RHF [Cp2 NaÐTHF] [Cp2 Na3 Ð3THF]C [Cp2 NaÐTHF] [Cp2 NaÐTHF] [Cp2 NaÐTHF] B3LYP [Cp2 NaÐTHF] [Cp2 Na3 Ð3THF]C [Cp2 NaÐTHF] [Cp2 Na3 Ð3THF]C [Cp2 NaÐTHF] [Cp2 NaÐTHF]

Basis set

υ11 (ppm)

υ22 (ppm)

υ33 (ppm)

Static9.4 T Static11.7 T

– –

– –

– –

6-31GŁŁ 6-31GŁŁ 6-311GŁŁ 6-311CGŁŁ 6-311CCGŁŁ

50.48 43.17 48.31 48.50 48.46

54.72 53.92 51.12 51.27 51.28

6-31GŁŁ 6-31GŁŁ 6-311GŁŁ 6-311GŁŁ 6-311CGŁŁ 6-311CCGŁŁ

42.83 43.32 36.48 34.76 36.63 36.93

46.66 46.88 41.52 40.33 42.26 42.07

υiso (ppm)

iso (ppm)

(ppm)



45.8 45.1

– –

18(3) 21(3)

0.3(2) 0.2(2)

77.25 76.27 75.94 76.27 76.40

60.81 57.79 58.46 58.68 58.71

637.41 634.39 635.06 635.28 635.31

26.77 33.10 27.63 27.77 27.94

0.68 0.35 0.80 0.80 0.80

71.51 79.66 65.84 67.88 66.67 66.95

53.66 56.62 47.95 47.66 48.52 48.65

630.26 633.22 624.55 624.26 625.12 625.25

28.68 36.34 29.35 33.12 30.04 30.02

0.73 0.80 0.66 0.66 0.63 0.66

Figure 7. Calculated orientations of the 23 Na EFG ((a) D along b-crystallographic axis; (b) D along a-axis) and sodium CS (c) tensors for CpNaÐTHF, (d) as well as the sodium EFG and CS tensor orientations in CpNa. All hydrogen atoms have been removed to enhance clarity. When viewing the 23 Na EFG tensor of CpNaÐTHF along the a-axis in (b), one Cp carbon atom has been removed to better illustrate the tensor components.

experimental and theoretical results. In our previous work on polymeric potassium metallocenes,39 it was shown that inter-chain effects were minimal and that the usage of a WTBS on the alkali-metal atom, in combination with the RHF method, produced better agreement between the theoretical and experimental 39 K EFG tensor parameters. By running a similar set of calculations for CpNaÐTHF, it is seen that the RHF/WTBS combination yields CQ values that are in good agreement with the experimental values (Tables 4 and S4(b)) and produce consistent 23 Na EFG tensor orientations. RHF/WTBS/6-311GŁŁ calculations on neutral

Copyright  2007 John Wiley & Sons, Ltd.

and cationic clusters provide reasonably good agreement between experimental and theoretical Q values, keeping in mind that the crystal structure was determined at 173(2) K, and earlier VT solid-state 23 Na NMR experiments predict an Q value near 0.65 at this temperature, as opposed to the room temperature value of 0.37(2).38

Chemical shielding tensors in CpNaÐTHF All calculations yield similar sodium CS parameters, and are in good agreement with experimental findings (Table 5). The 23 Na nucleus in CpNaÐTHF is typically found to

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

Investigation of structure and dynamics in sodium metallocenes

be deshielded relative to CpNa, confirming experimental observations and earlier conclusions that increased Cpcent -Na distances result in reduced magnetic shielding for s-block nuclei.38 Notable trends over all calculations, with respect to experimental values, include: (i) the 23 Na nucleus is shielded by ca 10 ppm, (ii) CS tensor spans are overestimated by about 50%, and (iii) skews are more positive. The theoretical principal components of the sodium CS tensors over all calculations involving triple- basis sets lie very close to the crystallographic axes, with average values for 6  11 –Na–b, 6  –Na–a, and 6  –Na–c being 4.6(1.1)° , 3.0(1.4)° and 22 33 3.4(0.1)° , respectively. Computations involving larger basis sets (i.e. a WTBS) and cluster sizes were attempted, but did not provide additional insight. Simulations of the CpNaÐTHF site in the static solid-state 23 Na NMR spectra of the mixed sample indicate that the CS and EFG tensors are non-coincident. Provided a triple- basis set is used, the orientations of the EFG and CS tensors vary only slightly, and average Euler Angle values over all computations are in fair (˛ and ) to very good (ˇ) agreement with experimental findings: ˛ D 89.80.1° , ˇ D 89.40.2° and  D 1.70.6° . Additional information and discussion pertaining to calculated carbon chemical shielding tensors in CpNaÐTHF can be found in the Supporting Information (Tables S5 and S6).

Reconciliation of VT NMR/pXRD data for CpNa Previous VT 23 Na NMR measurements showed a linear decrease in CQ with decreasing temperature. To rationalize this, it was concluded that the Cpcent –Na distance of CpNa increases with decreasing temperature.38 The pXRD data presented herein indicate that the length of the unit cell along the c-crystallographic axis is invariant with respect to temperature, while the a and b values decrease with decreasing temperature. As variation in the length of the unit cell c-axis is not present, [and hence, there is no significant change in rCpcent –Na)], additional computational modelling is used to reconcile the 23 Na NMR and pXRD data sets. The following alternative models are considered: (i) medium-range inter-chain interactions, (ii) movement of the sodium atom towards the Cp rings (close to the c unit cell axis) and (iii) movement of the sodium atom in directions perpendicular to the c-axis. As each CpNa chain lies adjacent to six other CpNa chains (Fig. S6), it is possible that medium-range (ca 3.5 ˚ electrostatic interactions might influence the 23 Na to 7.0 A) EFG tensor. The first model shows that as the lengths of the a- and b-crystallographic axes are decreased (modelling a decrease in both system temperature and inter-chain spacing), the value of CQ decreases very slightly, while Q increases slightly (Fig. 8(a)/Table S7). This model is in clear disagreement with the solid-state 23 Na NMR observations and is rejected. The second model looks at the possibility of the sodium atom moving towards one of the two proximate Cp rings (i.e. displacement nearly along the c-crystallographic axis, Fig. S7), a type of motion which would be expected to increase in frequency and/or amplitude as the temperature of the system is increased. The calculations show that as the

Copyright  2007 John Wiley & Sons, Ltd.

(a)

0.16 2.515

0.14

2.512

0.12

2.509

0.10

2.506

0.08

2.503

0.06

2.500

0.04

2.497

0.02

2.494 123

0.00 153

183

213

243

273

303

Temperature (K)

(b)

3.0

0.040

2.8

0.035

2.6

0.030

2.4

0.025

2.2

0.020

2.0

0.015

1.8 1.6

0.010

1.4

0.005

1.2 2.357 2.297 2.237 2.177 2.117 2.057 1.997 1.937

0.000

r(Cpcent-Na) (Å)

Figure 8. (a) ECMO calculations which attempt to incorporate medium-range electrostatic interactions predict that CQ ♦ should increase slightly with increasing temperature (dashed line; CQ MHz D 9.8 ð 105 TK C 2.483; R2 D 0.9443), while Q  should decrease with increasing temperature (solid line; Q D 7.0 ð 104 TK C 0.2310; R2 D 0.9872). (b) Movement of the sodium atom towards one Cp ring results in a decrease in the calculated CQ value (♦; 2 ˚ C 2.8652; CQ MHz D 0.0064r 2 A˚  C 0.0288 rA 2 R D 0.9993), while Q  is essentially invariant until rCpcent -Na > 0.45 A˚ from the crystallographic value.

sodium atom is moved towards one of the Cp rings, the value of CQ decreases in a fashion that can be fitted to a second-order polynomial function (Fig. 8(b)). The Q does ˚ not vary significantly until the displacement is over 0.45 A from the experimentally determined position (Fig. 8(b) and Table S8), which is unlikely in reality. A constant Q value is consistent with the solid-state 23 Na NMR data, however, a CQ that is at a maximum at the sodium crystallographic position, which is also the position at which the SCF system energy is minimized, discounts this model as well. The third model involves the sodium atom undergoing a displacement from its central position in a direction perpendicular to the axis of chain propagation (Fig. S8). According to calculations for an isolated Cp2 Na cluster (Fig. 9(a), Table S9), as the sodium atom moves from its crystallographic position, the CQ value is observed to increase, while the Q value changes only very slightly (below our experimental detection limits). It is expected that as the temperature increases, the thermal energy available to the system would increase the amplitude of such motion, giving rise to an increase in CQ , in agreement ˚ displacement from the with the NMR data. A 0.2 A crystallographic position corresponds to an energy change of ca 0.001 Hartrees or 2.6 kJ mol1 , so displacements from

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

S125

S126

C. M. Widdifield et al.

(a)

-546.4326 -546.4328 a 3.20 -546.4330 a -546.4332 c 3.15 -546.4334 3.10 -546.4336 -546.4338 3.05 -546.4340 3.00 -546.4342 -546.4344 2.95 -546.4346 2.90 -546.4348 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 3.25

b

Displacement from Crystallographic Position in ab-plane (Å)

(b)

b

a a

c

Figure 9. (a) Sodium atom displacements perpendicular to the c-axis of the unit cell (the displacement considered here is illustrated 2 ˚  546.43; R2 D 0.9754) and CQ ♦; in the inset) result in energy (; E(Hartrees) D 1 ð 105 x2 A˚   5 ð 105 xA 2 2 2 ˚ C 2.9223; R D 1.000) value increases, while Q does not vary significantly. (b) A schematic CQ MHz D 0.0043 x A˚   0.0168 xA of the proposed temperature-dependent motion, where the lightly shaded dotted region represents the displacement region of the sodium atom. As the temperature is decreased, this region becomes smaller, as indicated by the solid circle.

the equilibrium position at T D 323 K are proposed to be ˚ to 0.2 A ˚ greater than at T D 123 K, based about 0.1 A on the calculated and experimentally observed values of CQ 23 Na. This final model is the most representative of the data; hence, we conclude that the sodium atom may be increasingly displaced from its crystallographic position in a plane perpendicular to the direction of chain propagation (Fig. 9(b)). Motional anisotropy of the metal atom, known as the Gol’danskii–Karyagin effect,73,74 although not often observed, is not unexpected and has been observed in similar systems, such as Cp2 Fe75 – 77 and CpŁ 2 Fe.78 This model is also in line with the expansion of the unit cell along the a- and b-axes with increasing temperature, as indicated by ECMO calculations conducted for structures corresponding to temperatures of 123, 213 and 323 K (Fig. S9). As the lengths of the a and b unit cell axes are increased (due to an increasing temperature), less energy is required to displace the Na atom from its central position.

CONCLUSIONS A variety of experimental data for CpNa and CpNaÐTHF, in combination with theoretical calculations of NMR tensors, have been used to further our understanding of the relationships between structure, dynamic motion and sodium

Copyright  2007 John Wiley & Sons, Ltd.

NMR parameters in the solid state. Single-crystal XRD techniques have provided a suitable structure for the unstable CpNaÐTHF complex, which allowed us to present meaningful computational results regarding the 23 Na EFG, sodium CS and carbon CS tensors in this species. Solid-state NMR experiments at multiple fields have allowed for more precise measurements of the sodium CS tensors for both species and allowed us to conclude that the sodium CS tensor in CpNa is not axial, as previously assumed. VT powder XRD data, in combination with previous VT 23 Na NMR data and new computational results, has allowed us to discount prior models regarding the temperature-dependent structural changes in CpNa, and has allowed us to present a model which is consistent with all available data and is similar to what was recently observed in the Cp2 Fe and CpŁ 2 Fe metallocenes. This work also stresses the importance of using NMR, XRD and computational techniques for accurate modelling of solid-state dynamic processes. Finally, solid-state 23 Na NMR has proven to be useful in identifying impurity phases that are present when either of these species decompose upon exposure to air or moisture and serve to compliment the available pXRD data. The 23 Na chemical shift and quadrupolar interactions studied herein should prove very useful for structural elucidation of a wide range of species exhibiting alkali cation– interactions.

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

Investigation of structure and dynamics in sodium metallocenes

Supplementary material Supplementary electronic material for this paper is available in Wiley InterScience at: http://www.interscience.wiley. com/jpages/0749-1581/suppmat/

Acknowledgements This research was funded by Imperial Oil and the Natural Sciences and Engineering Research Council (NSERC - Canada). CMW would like to thank NSERC for a PGSM scholarship. RWS thanks the Canadian Foundation for Innovation (CFI), the Ontario Innovation Trust (OIT) and the University of Windsor for funding the Solid-State NMR Facility at the University of Windsor. The Centre for Catalysis and Materials Research (CCMR) at the University is thanked for additional funding. Dr. Victor Terskikh and the National Ultrahighfield NMR Facility for Solids in Ottawa are acknowledged for use of the 21.1 T NMR spectrometer. Mr. Aaron Rossini and Dr. Andy Lo are thanked for acquiring the high-field NMR data.

REFERENCES 1. Kealy TJ, Paulson PL. Nature 1951; 168: 1039. 2. Wilkinson G, Rosenblum M, Whiting MC, Woodward RB. J. Am. Chem. Soc. 1952; 74: 2125. 3. Fischer EO, Pfab W. Z. Naturforsch. 1952; 7b: 377. 4. Thiele J. Ber. Chem. Gesell. 1900; 33: 666. 5. Thiele J. Ber. Chem. Gesell. 1901; 34: 68. 6. Wardell JL. In Comprehensive Organometallic Chemistry, Wilkinson G, Stone FGA (eds). Pergamon Press: Oxford, 1982. 7. Beswick MA. In Comprehensive Organometallic Chemistry II, Abel EW, Stone FGA, Wilkinson G (eds). Elsevier Science: Oxford, 1995. 8. Weiss E. Angew. Chem. 1993; 105: 1565 (See also Angew. Chem. Int. Ed. Engl. 1993; 1532(1511): 1501–1523). 9. Harder S. Coord. Chem. Rev. 1998; 176: 17. 10. Harder S. Coord. Chem. Rev. 2000; 199: 331. 11. Herberich GE, Fischer A. Organometallics 1996; 15: 58. 12. Schaeffer WP, Cotter WD, Bercaw JE. Acta Crystallogr., Sect. C 1993; C49: 1489. 13. Rabe G, Roesky HW, Stalke D, Pauer F, Sheldrick GM. J. Organomet. Chem. 1991; 403: 11. 14. Jordan V, Behrens U, Olbrich F, Weiss E. J. Organomet. Chem. 1996; 517: 81. 15. Dinnebier RE, Olbrich F, van Smaalen S, Stephens PW. Acta Crystallogr., Sect. B 1997; B53: 153. 16. Dinnebier RE, Olbrich F, Bendele GM. Acta Crystallogr., Sect. C 1997; C53: 699. 17. Dinnebier RE, Behrens U, Olbrich F. Organometallics 1997; 16: 3855. 18. Lin G, Wong W-T. Polyhedron 1994; 13: 3027. 19. Dohmeier C, Baum E, Ecker A, Koeppe R, Schnoeckel H. Organometallics 1996; 15: 4702. 20. Jutzi P, Leffers W, Pohl S, Saak W. Chem. Ber. 1989; 122: 1449. 21. Chen H, Jutzi P, Leffers W, Olmstead MM, Power PP. Organometallics 1991; 10: 1282. 22. Davidson MG, Stalke D, Wright DS. Angew. Chem. 1992; 104: 1265 (See also Angew. Chem. Int. Ed. Engl. 1992; 1231(1269): 1226–1267). 23. Paver MA, Russell CA, Stalke D, Wright DS. Chem. Commun. 1993; 1349. 24. Armstrong DR, Herbst-Irmer R, Kuhn A, Moncrieff D, Paver MA, Russel CA, Stalke D, Steiner A, Wright DS. Angew. Chem. 1993; 105: 1807 (See also Angew. Chem. Int. Ed. Engl. 1993; 1832(1812): 1774–1806). 25. Harder S, Prosenc MH. Angew Chem. Int. Ed. Engl. 1996; 35: 97. 26. Harder S, Prosenc MH, Rief U. Organometallics 1996; 15: 118. 27. Fritz HP, Schneider R. Chem. Ber. 1960; 93: 1171. 28. Fritz HP, Schaefer L. Chem. Ber. 1964; 97: 1829. 29. Cox RH, Terry HW Jr, Harrison LW. J. Am. Chem. Soc. 1971; 93: 3297.

Copyright  2007 John Wiley & Sons, Ltd.

30. Ford WT. J. Organomet. Chem. 1971; 32: 27. 31. Aoyagi T, Shearer HMM, Wade K, Whitehead G. J. Organomet. Chem. 1979; 175: 21. 32. Aoyagi T, Shearer HMM, Wade K, Whitehead G. Chem. Commun. 1976; 164. 33. Larsen FH, Jakobsen HJ, Ellis PD, Nielsen NC. J. Magn. Reson. 1998; 131: 144. 34. Lipton AS, Sears JA, Ellis PD. J. Magn. Reson. 2001; 151: 48. 35. Madhu PK, Goldbourt A, Frydman L, Vega S. Chem. Phys. Lett. 1999; 307: 41. 36. Kentgens APM, Verhagen R. Chem. Phys. Lett. 1999; 300: 435. 37. Schurko RW, Hung I, Widdifield CM. Chem. Phys. Lett. 2003; 379: 1. 38. Willans MJ, Schurko RW. J. Phys. Chem. B 2003; 107: 5144. 39. Widdifield CM, Schurko RW. J. Phys. Chem. B 2005; 109: 6875. 40. Rietveld HM. J. Appl. Crystallogr. 1969; 2: 65. 41. Tanabe M, Vandermeulen GWM, Chan WY, Cyr PW, Vanderark L, Rider DA, Manners I. Nat. Mater. 2006; 5: 467. 42. Freemantle M. Chem. Eng. News 2006; 84: 9. 43. SMART, Program for Data Collection and Cell Refinement. Bruker AXS: Madison, 2001. 44. SAINT, Program for Data Reduction. Bruker AXS: Madison, 2001. 45. Sheldrick GM. SHELXTL: Program for Crystal Structure Solution. Bruker AXS: Madison, 2001. 46. Sheldrick GM. SHELXL-97: Program for Crystal Structure Refinement. Universitat Gottingen: Gottingen, 1997. 47. Kraus W, Nolze G. PowderCell for Windows, 2.4. Federal Institute for Materials Research and Testing: Berlin, 2000. 48. Dumazy Y, Amoureux J-P, Fernandez C. Mol. Phys. 1997; 90: 959. 49. Eichele K, Wasylishen RE. WSolids, 2.2.13. Dalhousie University: Halifax, 1998. 50. Bak M, Rasmussen JT, Nielsen NC. J. Magn. Reson. 2000; 147: 296. 51. Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Montgomery JA, Vreven T, Kudin KN, Burant JC, Millam JM, Iyengar SS, Tomasi J, Barone V, Mennucci B, Cossi M, Scalmani G, Rega N, Petersson GA, Nakatsuji H, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Klene M, Li X, Knox JE, Hratchian HP, Cross JB, Adamo C, Jaramillo J, Gomperts R, Stratmann RE, Yazyev O, Austin AJ, Cammi R, Pomelli C, Ochterski JW, Ayala PY, Morokuma K, Voth GA, Salvador P, Dannenberg JJ, Zakrzewski VG, Dapprich S, Daniels AD, Strain MC, Farkas O, Malick DK, Rabuck AD, Raghavachari K, Foresman JB, Ortiz JV, Cui Q, Baboul AG, Clifford S, Cioslowski J, Stefanov BB, Liu G, Liashenko A, Piskorz P, Komaromi I, Martin RL, Fox DJ, Keith T, AlLaham MA, Peng CY, Nanayakkara A, Challacombe M, Gill PMW, Johnson B, Chen H, Wong MW, Gonzalez C, Pople JA. Gaussian 03, Revision B.03. Gaussian: Pittsburgh, 2003. 52. Becke AD. Phys. Rev. A 1988; 38: 3098. 53. Lee C, Yang W, Parr RG. Phys. Rev. B 1988; 37: 785. 54. Huzinaga S, Klobukowski M. Chem. Phys. Lett. 1993; 212: 260. 55. Huzinaga S, Miguel B. Chem. Phys. Lett. 1990; 175: 289. 56. Cummins PL, Bacskay GB, Hush NS. Mol. Phys. 1987; 62: 193. 57. Brown RD, Head-Gordon MP. Mol. Phys. 1987; 61: 1183. 58. Jonsson P, Ynnermann A, Fischer CF, Godefroid MR, Olsen J. ¨ Phys. Rev. A 1996; 53: 4021. 59. Jeckelmann B, Beer W, Beltrami I, De Boer FWN, De Chambrier G, Goudsmit PFA, Kern J, Leisi HJ, Ruckstuhl W, Vacchi A. Nucl. Phys. A 1983; A408: 495. 60. Yei W, Sieradzan A, Havey MD. Phys. Rev. A 1993; 48: 1909. 61. Schwalm D, Warburton EK, Olness JW. Nucl. Phys. A 1977; A293: 425. 62. Sundholm D, Olsen J. Phys. Rev. Lett. 1992; 68: 927. 63. Ditchfield R. Mol. Phys. 1974; 27: 789. 64. Wolinski K, Hinton JF, Pulay P. J. Am. Chem. Soc. 1990; 112: 8251. 65. Mason J (ed). Multinuclear NMR. Plenum Press: New York, 1987. 66. Bryce DL, Adiga S, Elliott EK, Gokel GW. J. Phys. Chem. A 2006; 110: 13568.

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

S127

S128

C. M. Widdifield et al.

67. Zhang B, Jin Z, Wang Y, Wei G, Chen W. Chem. Res. Chin. Univ. 1992; 8: 15. 68. Sagnowski SF, Sulek Z, Stachura M, Ogar J. Z. Phys. B 1982; 46: 123. 69. Sagnowski SF, Ogar J. Phys. Status Solidi B 1981; 107: K125. 70. Wong A, Wu G. J. Phys. Chem. A 2000; 104: 11844. 71. Wong A, Whitehead RD, Gan Z, Wu G. J. Phys. Chem. A 2004; 108: 10551. 72. Jost S, Gunther H. Magn. Reson. Chem. 2003; 41: 373.

Copyright  2007 John Wiley & Sons, Ltd.

73. Gol’danskii VI, Gorodinskii GM, Karyagin SV, Korytko LA, Krizhanskii LM, Makarov EF, Suzdalev IP, Khrapov VV. Dokl. Akad. Nauk SSSR 1962; 147: 127. 74. Karyagin SV. Dokl. Akad. Nauk SSSR 1963; 148: 1102. 75. Herber RH. Phys. Organomet. Chem. 2002; 3: 207. 76. Herber RH, Nowik I. Solid State Sci. 2002; 4: 691. 77. Herber RH, Nowik I. Hyperfine Interact. 2003; 144/145: 249. 78. Herber RH, Nowik I, Kahlenberg V, Kopacka H, Schottenberger H. Eur. J. Inorg. Chem. 2006; 2006: 3255.

Magn. Reson. Chem. 2007; 45: S116–S128 DOI: 10.1002/mrc

Investigation of structure and dynamics in the sodium metallocenes ...

Solid-state 23Na NMR spectra of two organometallic complexes, cyclopentadienylsodium (CpNa) and the tetrahydrofuran (THF) solvate of CpNa (CpNa·THF), are presented. Analytical simulations of experimental spectra and calculated 23Na electric-field gradient (EFG) tensors confirm that both complexes are present.

563KB Sizes 0 Downloads 224 Views

Recommend Documents

Investigation on morphology of polymer structure in ...
Apr 24, 2016 - Polymer stabilization effect to increase the thermal stability of blue phases (BPs) for potential application in flat-panel display has been ... most LC materials, BPs are usually stable within a very narrow temperature range ... chain

Dividend Dynamics and the Term Structure of Dividend Strips
Dividend Dynamics and the Term. Structure of Dividend Strips. FREDERICO BELO, PIERRE COLLIN-DUFRESNE, and ROBERT S. GOLDSTEIN∗. ABSTRACT. Many leading asset pricing models are specified so that the term structure of dividend volatility is either fl

Dividend Dynamics and the Term Structure of Dividend ...
We thank the editor, Cam Harvey, as well as the associate editor and an ..... the per-year standard deviation of dividend growth across each horizon T for the two.

Electronic structure of sodium tungsten bronzes NaxWO3 by high ...
Apr 25, 2007 - A high metallic con- ... structure of NaxWO3 is of great interest from both techno- ... NaxWO3, Brown and Banks6 have shown that the crystal.

Complex networks: Structure and dynamics
Jan 10, 2006 - 255. 6.2. The Internet and the World Wide Web . ... Structure of the Internet . ..... Watts' pioneering book on the subject deals with the structure and the dynamics .... emerging, e.g., in mobile and wireless connected units.

PDF Online Structure and Dynamics
PDF Online Structure and Dynamics: An Atomic View of Materials. (Oxford Master Series in Physics) - Read Unlimited eBooks and. Audiobooks - By Martin T.

Population structure and evolutionary dynamics of ...
them to be surrounded by a cloud of isolates differing from them at one or .... tional parameters in Streptococcus pneumoniae from multilocus se- quence typing ...

Structure and Dynamics of Parallel b-Sheets ...
We study three systems (S1, S2, and S3) that differ in the initial staggering ... connecting loop, with V24–N27 being more extended in the S3 system. Additional ...

Population structure and evolutionary dynamics of ...
sequence data, described in the next section. A re-analysis of the electrophoretic data(6) was stimulated by a relatively small dataset for Neisseria gonorrhoeae ...

Dynamics of the Chilean Web structure
Dec 9, 2005 - (other non .cl sites hosted in Chile are estimated to number ... but there is no path to go back to MAIN; and. (d) other ... with 94,348 having a DNS server. Hence, in ..... site appeared at the end of 1993 in our CS depart- ment.

Sodium Bicarbonate vs Sodium Chloride for the ...
Sep 3, 2008 - To evaluate the effect of sodium bi- carbonate on the ... obtain consent, receipt of a sodium bi- carbonate .... Data Collection and Management.

Investigation and Treatment of Missing Item Scores in Test and ...
May 1, 2010 - This article first discusses a statistical test for investigating whether or not the pattern of missing scores in a respondent-by-item data matrix is random. Since this is an asymptotic test, we investigate whether it is useful in small

THE MODULE STRUCTURE OF HOCHSCHILD HOMOLOGY IN ...
(X). The wedge product yields a ring structure on S∗(TX[−1]) and hence on HT. ∗. (X), but the map I in general is not a isomorphism of rings. It was Kontsevich's ...

Task Dynamics and Resource Dynamics in the ...
source-dynamic components that have not been considered traditionally as ... energy for the act (Aleshinsky, 1986; Bingham, 1988; Bobbert,. 1988; Van Ingen .... As an alternative explanation, Kugler and Turvey (1987) suggested that periods ...

Task Dynamics and Resource Dynamics in the ...
the Royal Society, London, Series B, 97, 155-167. Hinton, G. E. (1984). Parallel computations for controlling an arm. Journal of Motor Behavior, 16, 171-194.

THE MODULE STRUCTURE OF HOCHSCHILD HOMOLOGY IN ...
1. Introduction. Let X be a smooth projective variety over the complex numbers and let ∆: X → X ×X denote .... By definition, the following diagram. U. L. ⊗OX F. IK.

International Integration and the Structure of Exports in ...
Oct 28, 2008 - trade data disaggregated at a relatively high level. The authors' ... international economic integration from detailed evidence of past balance of payments and trade data ..... The product in the electrical and electronic equip-.