Neutron diffraction with magnesium isotope substitution, high energy x-ray diffraction, and 29Si, 27Al, and 25Mg solid-state nuclear magnetic resonance (NMR) spectroscopy were used to measure the structure of glassy diopside (CaMgSi2O6), enstatite (MgSiO3), and four (MgO)x(Al2O3)y(SiO2)1−x−y glasses, with x = 0.375 or 0.25 along the 50 mol. % silica tie-line (1 − x − y = 0.5) or with x = 0.3 or 0.2 along the 60 mol. % silica tie-line (1 − x − y = 0.6). The bound coherent neutron scattering length of the isotope 25Mg was remeasured, and the value of 3.720(12) fm was obtained from a Rietveld refinement of the powder diffraction patterns measured for crystalline 25MgO. The diffraction results for the glasses show a broad asymmetric distribution of Mg–O nearest-neighbors with a coordination number of 4.40(4) and 4.46(4) for the diopside and enstatite glasses, respectively. As magnesia is replaced by alumina along a tie-line with 50 or 60 mol. % silica, the Mg–O coordination number increases with the weighted bond distance as less Mg2+ ions adopt a network-modifying role and more of these ions adopt a predominantly charge-compensating role. 25Mg magic angle spinning (MAS) NMR results could not resolve the different coordination environments of Mg2+ under the employed field strength (14.1 T) and spinning rate (20 kHz). The results emphasize the power of neutron diffraction with isotope substitution to provide unambiguous site-specific information on the coordination environment of magnesium in disordered materials.
I. INTRODUCTION
Magnesium is an enigmatic element within network-forming oxide glasses. It can adopt a variety of different local environments, where Mg–O coordination numbers in the range from four to six or more are found in the structural chemistry of its crystalline oxides,1–6 and there is debate about the structural role played by four-coordinated magnesium as it is sometimes regarded as a network-forming species.7 Definitive experimental information on the coordination environment of Mg2+ and its dependence on the glass composition is therefore desirable in order to clarify the structural role of magnesium in disordered network structures. Such information is, however, scarce because of (i) the chameleon-like nature of the magnesium coordination environment, (ii) limitations in x-ray scattering experiments that originate from the low atomic number of magnesium, and (iii) difficulty in interpreting the results obtained from 25Mg solid-state nuclear magnetic resonance (NMR) experiments.8 This isotope of magnesium has the only NMR active nucleus, but its natural abundance is small (10%) and the nucleus of this isotope has a low gyromagnetic ratio and significant quadrupolar moment (nuclear spin I = 5/2).9
In this paper, we investigate the structure of the diopside and enstatite composition glasses (CaO)0.25(MgO)0.25(SiO2)0.5 and (MgO)0.5(SiO2)0.5, respectively, along with four magnesium aluminosilicate (MgAS) glasses (MgO)x(Al2O3)y(SiO2)1−x−y, with x = 0.375 or 0.25 along the 50 mol. % silica tie-line (1 − x − y = 0.5), or with x = 0.3 or 0.2 along the 60 mol. % silica tie-line (1 − x − y = 0.6). The latter correspond to glass compositions for which the ratio R = x/y is either three or unity, respectively. These glasses were chosen for investigation because they cover a range of compositions where the role of Mg2+ is expected to change from a network modifying (R → ∞) to a predominantly charge-compensating (R = 1) role. They also serve as model systems for technological materials such as commercial display glass10,11 and magmatic materials of interest in the geosciences.12–15
The glass structure was probed by combining neutron diffraction with magnesium isotope substitution,16 high energy x-ray diffraction, and 29Si, 27Al, and 25Mg solid-state NMR spectroscopy. The diffraction and NMR experiments were performed on the same set of samples. The neutron diffraction method was selected because it is designed to deliver unambiguous site-specific structural information.17 The experiments pointed, however, to a systematic error in the published value for the bound coherent neutron scattering length of the isotope 25Mg.18 This parameter was therefore remeasured in neutron powder diffraction experiments on isotopically enriched crystalline 25MgO. The requirement in the neutron diffraction work of isotopically enriched samples helped to motivate a reassessment of solid-state 25Mg NMR for delivering unambiguous information on the coordination environment of magnesium in glass.
II. THEORY
A. Bragg powder diffraction
In a diffraction experiment on a powdered crystalline sample, kinematic theory shows that the intensity of elastically scattered neutrons is given by19
where 2θ is the scattering angle, Fhkl is the structure factor for the Miller indices h, k, l, the fractional coordinates xj, yj, zj give the site of atom j within a unit cell, bj is the coherent neutron scattering length of atom j, and SOFj is the site occupation factor. For crystalline MgO, which has the NaCl structure type, the fractional coordinates for Mg (0, 0, 0) and O (1/2, 1/2, 1/2) are known from its space group Fmm, and SOFj = 1 for both sites because of the well-defined 1:1 stoichiometry of the oxide.2 Hence, if bO is known accurately but there is uncertainty associated with the value for magnesium , a Rietveld refinement of the measured diffraction pattern with fixed scattering lengths will deliver a fitted value . The correct value for the magnesium scattering length follows from the expression , where the tilde indicates that the value will be sensitive to the isotopic enrichment.
B. Pair-distribution function analysis
In a neutron diffraction experiment, the measured total structure factor is given by17
where k is the magnitude of the scattering vector, cα is the atomic fraction of chemical species α, bα is the bound coherent neutron scattering length of chemical species α, and Sαβ(k) is the Faber–Ziman partial structure factor for the chemical species α and β. In an x-ray diffraction experiment, the scattering lengths bα in Eq. (2) are replaced by the k-dependent form factors fα(k) and the total structure factor is usually rewritten as
where the mean value . Neutral atom form factors were used in the x-ray diffraction data analysis.20
Let neutron diffraction experiments be performed on two glasses that are identical in every respect, except that one contains magnesium of natural isotopic abundance, natMg, and the other is isotopically enriched with 25Mg. If the measured total structure factors are denoted by and , respectively, then the difference function
where , eliminates all those pair-correlation functions that do not involve magnesium. The Mg-α correlations with α ≠ Mg can also be eliminated by forming the difference function given by
The difference functions ΔFMg(k) and ΔF(k) therefore simplify the complexity of correlations associated with a single total structure factor.
The real-space functions corresponding to FN(k), ΔFMg(k), and ΔF(k) are obtained by Fourier transformation and are denoted by DN(r), ΔDMg(r), and ΔD(r), respectively, where r is a distance in real space. For example, the total pair-distribution function is given by
where the normalization factor is obtained by taking the modulus of Eq. (2) after setting all the Sαβ(k) functions to zero and ⊗ is the one-dimensional convolution operator. M(k) is a window function given by M(k) = 1 for k ≤ kmax and M(k) = 0 for k > kmax, where kmax is the maximum k-value and M(r) is the real-space manifestation of M(k). If kmax is sufficiently large that M(k) does not truncate oscillations in FN(k), Eq. (6) will deliver the unmodified total pair-distribution function
where ρ is the atomic number density and gαβ(r) is the partial pair-distribution function for the chemical species α and β. Similarly,
where
and the normalization factor B is obtained by taking the modulus of Eq. (4) after setting all the Sαβ(k) functions to zero. It follows that, if the first peak in ΔDMg(r) is attributable to Mg–O correlations alone, then the Mg–O coordination number can be obtained from
where r1 and r2 give the lower and upper boundaries of the peak, respectively. In addition
where
and the normalization factor C is obtained by taking the modulus of Eq. (5) after setting all the Sαβ(k) functions to zero.
At r-values below the distance of closest approach between two atoms of chemical species α and β, the partial pair-distribution function gαβ(r) = 0. It follows that, at these small r-values, DN(r) = ΔDMg(r) = ΔD(r) = −4πρr, i.e., all three functions follow the so-called density line. Provided the datasets have been correctly normalized, the application of M(r) and the finite counting statistics will generate oscillations around the density line in the measured functions. If the low-r oscillations in one of these r-space functions are set to the values given by the density line and the function is Fourier transformed to k-space, this back Fourier transform should be in agreement with the measured dataset at all k-values.21 The level of agreement between a measured dataset and the back Fourier transform therefore gives a measure of the accuracy of the diffraction results. In the initial analysis of the datasets using the published value for the scattering length of 25Mg,18 this self-consistency check was not obeyed, which pointed to the need for a remeasurement of the scattering length for this isotope.
To accommodate the effect of the window function, each peak i in rgαβ(r) was represented by the Gaussian function
where , , and are the peak position, standard deviation, and coordination number of chemical species β around α, respectively. In the neutron diffraction work, a measured r-space function was fitted to a suitable sum of these Gaussian peaks, convoluted with M(r), using the procedure described in Ref. 22. The goodness-of-fit was assessed by the parameter Rχ.23
In the x-ray diffraction work, the total pair-distribution function was obtained from Eq. (6) after replacing by . The contribution of each Gaussian peak to SX(k) − 1 was calculated and then Fourier transformed to real space using the same M(k) function as used for the experimental data.22 A least squares procedure was used to fit an appropriate sum of these Fourier transforms to the function using the program PXFIT,80 where and . The fitted functions are presented as for ease of comparison with the neutron diffraction results. The goodness-of-fit was also assessed by the parameter Rχ23 as calculated for the representation of the x-ray datasets.
III. EXPERIMENT
A. Sample preparation and characterization
The enstatite samples were prepared using the method described in Ref. 16 from either natMgO (99%) or isotopically enriched 25MgO and SiO2 (99%). The isotopic enrichment of 25Mg was measured to be 1.00(2)% 24Mg, 98.49(3)% 25Mg, and 0.51(1)% 26Mg using Inductively Coupled Plasma Mass Spectrometry (ICP-MS).
The isotopically enriched diopside glass was made from powders of 25MgO (0.81% 24Mg, 98.79% 25Mg, 0.40% 26Mg), CaCO3 (99%), and SiO2 (99%) using the method described in Ref. 24. The powders were calcined overnight at 800 °C and were then melted in air at a processing temperature Tproc ∼1492 °C in a Pt/10%Rh crucible. The melt was quenched by immersing the bottom of the crucible in water. The melt and quench process was repeated to ensure a homogeneous glass. The diopside glass of natural isotopic abundance was prepared from powders of natMgO (Aldrich, ≥99.99%), CaCO3 (Aldrich, ≥99.999%), and SiO2 (Alfa Aesar, 99.9%). The powders were calcined overnight at 800 °C and mixed by shaking. A batch of mass ∼3 g was transferred to a Pt/10%Rh crucible, melted by heating from room temperature to Tproc = 1550 °C in about 1 h, and held at this temperature for 1 h. The melt was quenched by placing the bottom of the crucible onto a liquid-nitrogen cooled copper block. The sample was then ground, remelted in a furnace at 1550 °C for 1 h, and quenched again. The mass loss was 0.08% on the first melt and 0.15% on the second melt, which was attributed to the loss of water re-adsorbed during the preparation procedure.
The aluminosilicate samples are designated by MgAS_x_(100-x-y) with x and y expressed in mol. %. They were prepared from either natMgO (Aldrich, ≥99.99%) or isotopically enriched 25MgO (Isoflex, 0.32% 24Mg, 99.38% 25Mg, 0.30% 26Mg), Al2O3 (Sigma-Aldrich, 99.998%), and SiO2 (Alfa Aesar, 99.9%). The powders were calcined at 1000 °C for 2 h in a Pt or Pt/10%Rh crucible.
For the MgAS_37.5_50 and MgAS_25_50 samples, an appropriate mass of each powder was mixed by shaking. The mixture was then transferred to a Pt/10%Rh crucible and dried at 1000 °C for 10 min to remove water adsorbed during the weighing and mixing process. Each batch of mass ∼3 g was melted by heating from room temperature to a temperature Tproc of either 1550 °C (MgAS_37.5_50) or 1650 °C (MgAS_25_50) in about 1 or 1.25 h, respectively, and was then held at this temperature for 1 h. The melt was quenched by placing the bottom of the crucible onto a liquid-nitrogen cooled copper block. The mass loss on melting was 0.3%, which was attributed to the loss of re-adsorbed water.
The MgAS_30_60 and MgAS_20_60 glasses were made by adding SiO2 to either MgAS_37.5_50 or MgAS_25_50, respectively, after the latter had been finely ground and dried at 200 °C for 1 h. The powders were mixed by shaking. Each batch of mass ∼2 g was melted in a Pt/10%Rh crucible by putting that crucible into a furnace at 1650 °C for 1 h. The melt was quenched by placing the bottom of the crucible onto a liquid-nitrogen cooled copper block. Each sample was ground before the melting and quench process was repeated. The mass loss was ≤0.46% on the first melt and ≤0.29% on the second melt, which was attributed to the loss of re-adsorbed water.
The stoichiometry of the enstatite glasses was measured using electron microprobe analysis and the density was measured using Archimedes’s principle with toluene as the immersion medium.16 The stoichiometry of the other glasses was taken from the batch composition and their density was measured by He pycnometry (Table I).
The processing temperature Tproc and number density ρ of the diopside, enstatite, and MgAS glasses.
Sample . | x . | y . | Tproc (°C) . | ρ (Å−3) . |
---|---|---|---|---|
Diopside | ⋯ | ⋯ | 1492–1550 | 0.0786(1) |
Enstatite | 0.4939 | 0 | 1650 | 0.0806(1) |
MgAS_37.5_50 | 0.375 | 0.125 | 1550 | 0.0804(1) |
MgAS_25_50 | 0.250 | 0.250 | 1650 | 0.0795(1) |
MgAS_30_60 | 0.300 | 0.100 | 1650 | 0.0771(1) |
MgAS_20_60 | 0.200 | 0.200 | 1650 | 0.0766(1) |
Sample . | x . | y . | Tproc (°C) . | ρ (Å−3) . |
---|---|---|---|---|
Diopside | ⋯ | ⋯ | 1492–1550 | 0.0786(1) |
Enstatite | 0.4939 | 0 | 1650 | 0.0806(1) |
MgAS_37.5_50 | 0.375 | 0.125 | 1550 | 0.0804(1) |
MgAS_25_50 | 0.250 | 0.250 | 1650 | 0.0795(1) |
MgAS_30_60 | 0.300 | 0.100 | 1650 | 0.0771(1) |
MgAS_20_60 | 0.200 | 0.200 | 1650 | 0.0766(1) |
For each glass composition, the electronic structure will not depend on isotope enrichment, so the x-ray diffraction patterns measured for the natMg and 25Mg samples should be the same within statistical error. Figure 1 shows this is the case to a good level of approximation for the investigated MgAS materials. The agreement for the diopside glass is less satisfactory, which likely originates from differences in the sample preparation procedure for the natMg and 25Mg samples.
The SX(k) functions measured for diopside glass and for several of the MgAS glasses containing either natMg (black curves) or 25Mg (red curves). For each of the MgAS glasses, the curves overlay on the scale of the plot, as emphasized by taking the difference between them and multiplying by a factor of ten (magenta curves). Several of the curves are displaced vertically for clarity of presentation.
The SX(k) functions measured for diopside glass and for several of the MgAS glasses containing either natMg (black curves) or 25Mg (red curves). For each of the MgAS glasses, the curves overlay on the scale of the plot, as emphasized by taking the difference between them and multiplying by a factor of ten (magenta curves). Several of the curves are displaced vertically for clarity of presentation.
Crystalline samples of natMgO and 25MgO, of mass 195.6 and 312.3 mg, respectively, were prepared from the enstatite glasses. The silicate was placed in a polytetrafluoroethylene (PTFE) beaker and dissolved in 40% hydrofluoric acid. The resulting solution was evaporated, several milliliters of concentrated sulfuric acid were then added, and the solution was re-evaporated. The resulting MgSO4 was transferred to a corundum crucible and fired at 1200 °C for 3 h to obtain pure MgO. The isotopic abundance of natMg was measured to be 74.6(2)% 24Mg, 11.0(1)% 25Mg, and 14.4(1)% 26Mg using ICP-MS. In comparison, the standard isotopic abundance of natMg, as used in the preparation of standard tables of neutron scattering lengths,25,26 is 78.992(25)% 24Mg, 10.003(9)% 25Mg, and 11.005(19)% 26Mg.27
B. Neutron powder diffraction
The instrument D4c at the Institut Laue-Langevin28 was used to perform diffraction experiments on the powdered crystalline samples because of the high flux of incident neutrons and its ability to deliver diffraction patterns with excellent reproducibility and count-rate stability.29 The incident neutron wavelength λ was either 0.4971(5) or 0.6983(4) Å, as found from a Rietveld refinement of the diffraction pattern measured for a powdered nickel standard, and the λ/2 scattering from the monochromator was removed using either a Rh or Ir filter, respectively. In each experiment, powder diffraction patterns were measured for the 25MgO or natMgO sample in a cylindrical vanadium container of inner diameter 4.78 mm and wall thickness 0.3 mm, the empty container, the empty instrument, and a cylindrical vanadium rod of diameter 6.08 mm for normalization purposes. In these experiments, the longer wavelength leads to improved resolution of the Bragg peaks in a powder diffraction pattern at the expense of a reduction in kmax.
For each wavelength, the datasets were processed in two different ways. In the first procedure, the diffraction pattern for the sample in its container was corrected for background and container scattering by subtracting a linear combination of the diffraction patterns for the empty instrument and empty container. Corrections for sample attenuation and multiple scattering were then applied. In the second procedure, the total structure factor FN(k) was obtained using the procedure described in Ref. 21, which takes into account sample and container attenuation, multiple scattering from both the sample and container, and inelasticity corrections. Finally, the variable was changed from k to the scattering angle 2θ by using the expression k = (4π/λ) sin(θ).17
A Rietveld refinement was performed on each dataset, represented as a function of 2θ, using both the FullProf (version 7.60)30,31 and GSAS-II (version 5241)32 packages. In each refinement, the scattering length bO = 5.805(4) fm25,33 and a rock-salt structural model was employed. The zero-point shift and overall scale factor were initially refined, followed by the cell parameter a and thermal displacement parameters for the Mg and O atoms. The background originating from diffuse scattering was fitted using a twelfth-order Chebyshev polynomial. In FullProf, the peak shapes were fitted using the pseudo-Voigt function , which has both Lorentzian L(x) and Gaussian G(x) contributions, with the variable x = θ − θh, where θh is the Bragg peak position and is a weighting factor. The full-width at half-maximum (FWHM) of the Gaussian function is given by the Caglioti relation FWHM2 = U tan2 θ + V tan θ + W.34,35 The parameters U, V, W, and were obtained by refinement along with two asymmetry parameters.31 In GSAS, the peak shapes were fitted using an axial-divergence-broadened pseudo-Voigt function with the FWHM of the Gaussian function given by the Caglioti relation. The parameters U, V, W, and were refined along with the axial-divergence asymmetry parameter SH/L.
In a Rietveld refinement, the scattering length of magnesium cannot be obtained directly (Sec. II A). Instead, it was found by refining the occupation factor for the Mg atom after the magnesium scattering length is set to the published value of = 3.62(14) fm18 for the 25MgO sample or = 5.375(4) fm36 for the natMgO sample. The set magnesium scattering length is then multiplied by the refined occupation factor to give a corrected value, which is denoted by or , respectively, and corresponds to the isotopic enrichment of magnesium in the measured sample.
In total, for each of the 25MgO and natMgO samples, eight values of the magnesium scattering length were generated. The values did not show any obvious dependence on (i) the λ value chosen for the diffraction experiment, (ii) the procedure used to process the diffraction patterns, or (iii) the software package used for the Rietveld refinement. The mean value of the scattering length for a given sample was therefore calculated using all the available results and the error was found from the standard deviation about the mean.
C. 29Si and 27Al MAS NMR
The 29Si magic angle spinning (MAS) NMR spectra were recorded at room temperature at a magnetic flux density of 5.64 or 14.1 T using an Agilent DD2 or Bruker Avance Neo 600 spectrometer, respectively, equipped with commercial MAS NMR probes. The investigated samples were isotopically enriched with 25Mg. The spectra for the diopside and enstatite glasses were measured at 5.64 T using 4.0 mm rotors spun at 5 kHz, 90° pulses of 4 µs length, and a relaxation delay of 900 s.8 The spectra for the MgAS glasses were measured at 14.1 T using 2.5 mm rotors spun at 10 kHz, 90° pulses of 2.7 µs length, and a relaxation delay of 900 s.
The 27Al MAS NMR spectra were recorded at room temperature at a magnetic flux density of 14.1 T on the Bruker Avance Neo 600 spectrometer, using 2.5 mm rotors, spinning rates between 15 and 25 kHz, 90° pulses of 1.2 µs length, and a relaxation delay of 0.5 s. The samples were isotopically enriched with 25Mg or contained magnesium of natural isotopic abundance. The 27Al triple-quantum magic angle spinning (TQMAS) experiments37 utilized the standard three-pulse zero-filtering sequence, with pulse lengths of 3.9 and 1.7 µs for the creation and conversion of triple-quantum coherences, respectively, with a power level of 120 W. The soft detection pulse applied for the zero-to-single quantum conversion had a length of 10.2 µs with a power level of 1.5 W. The relaxation delay was 0.1 s. A series of free induction decay (FID) signals in the TQMAS measurements was Fourier transformed and sheared using the software TopSpin to obtain the separate isotropic (F1) and anisotropic (F2) dimensions. The isotropic chemical shift δiso and quadrupolar product PQ were derived from the sheared two-dimensional spectra using the expressions38
and
Here, δ1 and δ2 are the chemical shifts in the F1 and F2 dimensions obtained from integration over the regions of interest, respectively, CQ and η (0 ≤ η ≤ 1) are parameters that describe the magnitude and symmetry of the electric field gradient tensor interacting with the nuclear electric quadrupole moment, respectively, ν0 is the Zeeman frequency, and S is the spin quantum number of the nucleus.
The MAS NMR spectra were simulated using the DMfit program.39 Gaussian/Lorentzian line shapes were assumed for (i) the spin-1/2 29Si nucleus and (ii) the spinning sidebands associated with the |m| = 1/2 ↔ |m| = 3/2 satellite transitions for the 27Al nucleus (for spin-5/2 nuclei, the second-order quadrupolar broadening effects are negligible for these transitions40). In the above, m represents the magnetic quantum number. The 27Al central-transition line shapes were simulated with the Czjzek model,41 which is based on the assumption of a wide distribution of quadrupolar coupling strengths.
The 29Si and 27Al chemical shifts are reported relative to tetra-methylsilane (TMS) and 1M Al(NO3)3 aqueous solution, respectively, using solid kaolinite [δiso(29Si) = −91.5 ppm] and AlF3 [δiso(27Al) = −16 ppm] as secondary references, respectively.
D. 25Mg MAS and static NMR
The 25Mg MAS NMR spectra were measured at room temperature at a magnetic flux density of 14.1 T (36.8 MHz) in the form of rotor-synchronized Hahn spin echoes on the Bruker Avance Neo 600 spectrometer using 3.2 mm rotors spun at 20.0 kHz. Experiments were performed at two different power levels corresponding to nonselective nutation frequencies ν1 of 25 and 80 kHz (measured on cubic MgO). The effective 90° pulse lengths were optimized by maximizing the echo intensity in a tp − τ − 2tp − τ acquisition scheme, with τ = 50 µs and tp values of 9 and 3.5 µs, respectively, for the two nutation frequencies employed. The relaxation delay was 0.5 s.
The static 25Mg NMR spectra were measured using a Wideband Uniform Rate Smooth Truncation (WURST) pulse sequence42 combined with the Carr–Purcell–Meiboom–Gill (CPMG) echo train acquisition scheme.43,44 To cover a wide-frequency range, the signal was acquired at multiple distinct base frequencies, corresponding to resonance offsets from the center of the m = 1/2 ↔ m = − 1/2 transition of ±5000, ±10 000 ppm and up to ±30 000 ppm. These spectra were recorded with the objective of testing for the presence of noncentral (|m| = 1/2 ↔ |m| = 3/2 and |m| = 3/2 ↔ |m| = 5/2) satellite transitions, which are affected to different extents by the anisotropy of nuclear electric quadrupolar interactions. The WURST experiments were performed using excitation and refocusing pulses with an identical length of 50.0 µs, an excitation bandwidth of 1 MHz, and a recycle delay of 0.2 s. The pulse spacing was set to give a spikelet separation of 6.7 kHz upon Fourier transformation and the number of Meiboom–Gill loops was set to 256.
The simulations of the 25Mg MAS line shape used the Czjzek model and were performed using the ssNake45 software. The simulations of the 25Mg static line shapes also used the Czjzek model and were performed using in-house Matlab® code with SIMPSON46 as the kernel. Details of the simulation procedures are described elsewhere.8
The 25Mg chemical shifts are reported relative to aqueous MgCl2 solution using solid natMgO [δiso(25Mg) = 27 ppm] as a secondary reference.
E. Neutron diffraction with isotope substitution
The neutron diffraction with isotope substitution work on the amorphous materials employed the diffractometer D4c,28 chosen because of the attributes described in Sec. III B and its ability to access a wide k-range.
The incident neutron wavelength was λ = 0.4958(1) Å for the experiments on the diopside, MgAS_25_50 and MgAS_37.5_50 compositions, λ = 0.4955(1) Å for the experiment on the MgAS_20_60 composition, and λ = 0.4978(1) Å for the experiment on the MgAS_30_60 composition. In each experiment, the coarsely ground samples were held in a cylindrical vanadium container of inner diameter 4.8 mm (diopside or MgAS_30_60) or 6.8 mm (all other samples) and wall thickness 0.1 mm. The same container was used for both samples of a given glass composition in order to reduce the occurrence of systematic errors.21 Diffraction patterns were measured at room temperature (≃298 K) for each of the samples in its container, the empty container, the empty instrument, and a cylindrical vanadium rod of diameter 6.08 mm for normalization purposes. The relative counting times for the sample-in-container and empty container measurements were optimized in order to minimize the statistical error on the container corrected intensity.47 Data analysis followed the procedure described in Ref. 21.
F. X-ray diffraction
The x-ray diffraction experiments were performed at room temperature using beamline 6-ID-D at the Advanced Photon Source with an incident photon energy of 100.233 keV. A Varex 4343CT amorphous silicon flat panel detector was employed with the sample to detector distance set at 311 mm as found from the diffraction pattern measured for crystalline CeO2. Powdered glass samples were held in cylindrical Kapton polyimide tubes of 1.80(1) mm internal diameter and 0.051(6) mm wall thickness. Diffraction patterns were measured for each sample in its container, an empty container, and the empty instrument. The data were converted to one-dimensional diffraction patterns using FIT2D.48 The background scattering, beam polarization, attenuation and Compton scattering corrections were performed using PDFgetX2.49
IV. RESULTS
A. 25Mg neutron scattering length
Figure 2 shows the powder diffraction patterns for the crystalline 25MgO and natMgO samples measured at both neutron wavelengths, where the raw datasets were processed by using the first correction procedure described in Sec. III B. The diffraction patterns were refined using both the FullProf (see Fig. 2) and GSAS-II packages. As a consistency check, the raw datasets were also processed using the second correction procedure described in Sec. III B, and the corrected diffraction patterns were refined using both Rietveld refinement packages. For a given sample, the results from all these measurements were the same, within the experimental error.
Powder diffraction patterns measured for (a) and (b) 25MgO vs (c) and (d) natMgO crystalline samples at different incident neutron wavelengths and the Rietveld refinements using FullProf. Each panel shows the measured intensity Iobs (red circles), calculated intensity Icalc (black curve), and the difference of intensities Iobs − Icalc (blue curve), along with the positions of Bragg reflections allowed by the space group for the structure (vertical green bars). Rwp denotes the weighted profile R-factor.
Powder diffraction patterns measured for (a) and (b) 25MgO vs (c) and (d) natMgO crystalline samples at different incident neutron wavelengths and the Rietveld refinements using FullProf. Each panel shows the measured intensity Iobs (red circles), calculated intensity Icalc (black curve), and the difference of intensities Iobs − Icalc (blue curve), along with the positions of Bragg reflections allowed by the space group for the structure (vertical green bars). Rwp denotes the weighted profile R-factor.
Both the 25MgO and natMgO samples contain finite amounts of all three of the stable magnesium isotopes 24Mg, 25Mg, and 26Mg. The system of linear equations for extracting all three of the scattering length values is therefore under-determined, so we adopted an iterative approach to obtain a consistent set of values.
For the 25MgO sample, an average value = 3.745(10) fm was obtained from the Rietveld refinements of all the diffraction patterns, where the tilde indicates a contribution from small amounts of the 24Mg and 26Mg isotopes. To correct for this contribution, the scattering lengths = 5.66(3) fm26 and = 4.89(15) fm18 were initially employed, along with the isotopic abundance of our 25Mg sample as measured by ICP-MS (Sec. III A). Here, the reported value for was calculated from the value = 5.375(4) fm measured by neutron interferometry36 using the values = 3.62(14) fm and = 4.89(15) fm found in previous neutron powder diffraction work18 and the standard isotopic abundance of natMg reported in Ref. 27, where is the dominant isotope (Sec. III A). Hence, a new value for was obtained, leading to a new value for and, in turn, a new value for . This iterative procedure converged rapidly and the high isotopic enrichment of the 25MgO sample led to a small correction of the order 0.7%. The final scattering length values are listed in Table II.
Bound coherent neutron scattering lengths for natMg and the isotopes of magnesium.
Species . | b (fm) . | Reference . |
---|---|---|
natMg | 5.375(4) | 36 |
24Mg | 5.65(3) | This work |
25Mg | 3.720(12) | This work |
26Mg | 4.89(15) | 18 |
For the natMgO sample, an average value = 5.304(35) fm was obtained from the Rietveld refinements of all the diffraction patterns, where the tilde indicates the isotopic abundance of our natMg sample as measured by ICP-MS (Sec. III A). In comparison, the value = 5.328(31) fm is obtained from the scattering lengths for the magnesium isotopes listed in Table II and the isotopic abundance of our natMg sample. The discrepancy between these scattering lengths is 0.45%, which is within the precision of the measurements.
For completeness, the value = 5.373(29) fm is obtained from the scattering lengths for the magnesium isotopes listed in Table II and the standard isotopic abundance of natMg reported in Ref. 27 (Sec. III A), which is in agreement with the value = 5.375(4) fm measured using neutron interferometry.36 The scattering lengths listed in Table II are therefore self-consistent, although the values for 24Mg and 26Mg should be remeasured.
Finally, the FN(k) function measured for the 25MgO sample using λ = 0.4976 Å, and corrected using the measured value = 3.745(10) fm that takes into account the isotopic enrichment, is shown in Fig. 3. Here, the wavelength was found from the diffraction pattern measured for the powdered nickel standard by fitting the five Bragg peaks that occur in the range 12° ≤ 2θ ° using a procedure in which the wavelength and zero-angle 2θ offset are varied, keeping the lattice parameter determining the Bragg peak positions fixed and allowing the parameters describing the intensity and shape of the Lorentzian–Gaussian peak profiles to vary. The 0.1% discrepancy with the value found from the Rietveld refinement of the same nickel diffraction data is well within the precision of the measurements. For comparison with the glass structures, the corresponding function is shown in Fig. 4 and was fitted by constraining the coordination numbers of the fitted Gaussian peaks to the values expected for the measured crystal structure.2 The first peak in originates from Mg–O correlations and is symmetrical, corresponding to a bond distance of 2.102(2) Å and coordination number = 6.00.
The FN(k) function for crystalline 25MgO measured using λ = 0.4976 Å. The measured function is shown by the black curve with vertical error bars, where the error bars are smaller than the curve thickness at most k-values.
The FN(k) function for crystalline 25MgO measured using λ = 0.4976 Å. The measured function is shown by the black curve with vertical error bars, where the error bars are smaller than the curve thickness at most k-values.
The function for crystalline 25MgO. The measured function (black solid circles) was obtained by Fourier transforming the FN(k) function shown in Fig. 3 with kmax = 23.55 Å−1 and ρ = 0.1070 Å−3. The black solid curve gives the fitted function and the other curves show the contributions from the Mg–O (magenta solid curves), O–O (green broken curves), and Mg–Mg (red solid curves) correlations. The displaced green solid curve shows the residual for the fitted range 1.5–10.0 Å (Rχ = 0.0295).
The function for crystalline 25MgO. The measured function (black solid circles) was obtained by Fourier transforming the FN(k) function shown in Fig. 3 with kmax = 23.55 Å−1 and ρ = 0.1070 Å−3. The black solid curve gives the fitted function and the other curves show the contributions from the Mg–O (magenta solid curves), O–O (green broken curves), and Mg–Mg (red solid curves) correlations. The displaced green solid curve shows the residual for the fitted range 1.5–10.0 Å (Rχ = 0.0295).
B. Solid-state NMR
1. 29Si MAS NMR
Figure 5 shows the 29Si MAS NMR spectra for the isotopically enriched samples. The signal-to-noise ratio is low because of the small sample size. The resonances are poorly resolved and, for the MgAS glasses, no sensible peak deconvolution was possible. Such poor resolution is expected from the large variety of possible silicon local coordination environments generated by the interaction of silicon with both the network modifier and aluminum species. The resonance positions are consistent with four-coordinated Si atoms.51
The 29Si MAS NMR spectra measured for 25Mg enriched magnesium silicate and aluminosilicate glasses. The datasets for enstatite and diopside are from Ref. 8. The red broken vertical lines mark the center of gravity of each peak shape.
The 29Si MAS NMR spectra measured for 25Mg enriched magnesium silicate and aluminosilicate glasses. The datasets for enstatite and diopside are from Ref. 8. The red broken vertical lines mark the center of gravity of each peak shape.
In alumina free glasses, the introduction of a network modifier oxide to silica leads to the breakage of Si–O–Si linkages and the formation of non-bridging oxygen (NBO) atoms. The tetrahedral [SiØ4/2]0 units of pure silica glass, where Ø denotes a bridging oxygen (BO) atom, are thereby converted to tetrahedral [SiØ3/2O]−, [SiØ2/2O2]2−, [SiØ1/2O3]3−, and [SiO4]4− units. These silicon species are designated by Qn, where n is the number of Si-BO bonds within a tetrahedron, and can be identified and quantified from a deconvolution of the measured 29Si MAS NMR spectra.51 The results for the diopside and enstatite glasses are reported in Ref. 8.
The spectra for vitreous aluminosilicates are, however, more complicated.52,53 In the standard model for (MO)x(Al2O3)y(SiO2)1−x−y glasses containing divalent modifier cations M2+ with R = x/y ≥ 1, which comprises the glass compositions relevant to this study, all the aluminum is present in the form of [AlØ4/2]− units. These units can form Si–O–Al linkages and their formal negative charge is compensated by the positive charge associated with M2+ ions. The M2+ species that fulfill this role are referred to as charge-compensating cations. All the NBO atoms are associated with the silicon atoms such that the probability of an Si-NBO bond is given by pSi-NBO = NNBO/4NSi, where NNBO and NSi are the numbers of NBO and Si atoms, respectively.
Both the ligation of Si to NBO atoms and its ligation to [AlØ4/2]− units are expected to shift the 29Si resonance toward higher (less negative) frequencies.54 Along a tie-line with fixed mol. % silica in the regime where R ≥ 1, the NBO to silicon atom ratio NNBO/NSi will decrease as MO is replaced by Al2O3, thus shifting the 29Si resonance toward lower (more negative) frequencies. In contrast, as the mole fraction of alumina increases there is a greater propensity for Si–O–Al linkages, thus shifting the 29Si resonance in the opposite direction, i.e., toward higher (less negative) frequencies. In consequence, there are two competing effects on the shift in position of the 29Si resonance.
The composition dependence of NNBO/NSi can be calculated from the glass composition.50 The number of Si–O–Al linkages per silicon atom NSi–O–Al/NSi can be estimated by first assuming a binomial distribution of Si-BO and Si-NBO bonds such that the fraction of Qn species is given by
A binomial distribution of Si–O–Si and Si–O–Al linkages, in which the probability of an Si–O–Al linkage is given by pSi–O–Al = NAl/(NSi + NAl) where NAl is the number of Al atoms,55 can then be used to find the fraction of Qn species that form m Si–O–Al bonds (m ≤ n). These species are denoted by and their fraction is given by
In this way, the overall fraction of the Si-centered tetrahedral units that form species can be deduced and the ratio NSi–O–Al/NSi can be calculated.
Figure 6 compares the R dependence of the ratios NNBO/NSi, NSi–O–Al/NSi, and NNBO/NSi–O–Al calculated for the standard model along both the 50 and 60 mol. % silica tie-lines. The ratio NNBO/NSi → 0 as R → 1 and the ratio NNBO/NSi–O–Al varies linearly (50 mol. % silica) or almost linearly (60 mol. % silica) with R. The results confirm that the observed shift of the 29Si resonance toward smaller (more negative) chemical shifts in the investigated MgAS glasses as R → 1 (Table III) is associated with a decrease in the NNBO/NSi–O–Al ratio. This shift is smaller for the 60 mol. % silica tie-line where, for the investigated R-range, the fractional change in NNBO/NSi–O–Al is concomitantly smaller.
The R dependence of the ratios NSi–O–Al/NSi (black curves), NNBO/NSi (black broken curves), and NNBO/NSi–O–Al (red curves) for the standard model (p = 1) of aluminosilicate glass. The ratio NNBO/NSi for the GYZAS model of magnesium aluminosilicates with p = 0.7750 is also given (blue broken curves). The ratios are calculated for tie-lines with either 50 mol. % silica (left column) or 60 mol. % silica (right column).
The R dependence of the ratios NSi–O–Al/NSi (black curves), NNBO/NSi (black broken curves), and NNBO/NSi–O–Al (red curves) for the standard model (p = 1) of aluminosilicate glass. The ratio NNBO/NSi for the GYZAS model of magnesium aluminosilicates with p = 0.7750 is also given (blue broken curves). The ratios are calculated for tie-lines with either 50 mol. % silica (left column) or 60 mol. % silica (right column).
Center of gravity of the 29Si MAS NMR spectra measured for 25Mg enriched silicate and aluminosilicate glasses. For the latter, the ratio R = x/y is given. The results for diopside and enstatite are taken from Ref. 8.
Glass . | δCG (±3 ppm) . | R . |
---|---|---|
Diopside | −84 | ⋯ |
Enstatite | −83 | ⋯ |
25MgAS_37.5_50 | −86 | 3 |
25MgAS_25_50 | −95 | 1 |
25MgAS_30_60 | −91 | 3 |
25MgAS_20_60 | −93 | 1 |
Glass . | δCG (±3 ppm) . | R . |
---|---|---|
Diopside | −84 | ⋯ |
Enstatite | −83 | ⋯ |
25MgAS_37.5_50 | −86 | 3 |
25MgAS_25_50 | −95 | 1 |
25MgAS_30_60 | −91 | 3 |
25MgAS_20_60 | −93 | 1 |
The above considerations are, nevertheless, approximate because the high cation field strength of Mg2+ will promote the formation of a fraction of higher-coordinated aluminum species when R ≥ 150 and because NBO atoms associated with Al species are neglected. The influence of these changes in speciation on the 29Si chemical shift is not well documented. Nevertheless, its effect on the R dependence of NNBO/NSi can be predicted by the so-called GYZAS model of Ref. 50, in which the fraction of Al(IV) species is set by an empirical parameter p found from 27Al MAS NMR experiments: fAl(IV) = (1 + p)/2 when R ≥ 1. As compared to the standard model, the ratio NNBO/NSi is no longer zero at R = 1 (Fig. 6), i.e., the shift in the 29Si resonance toward lower frequencies originating from the loss of NBO atoms will be comparatively smaller.
2. 27Al MAS NMR
Figure 7 shows the central transition in the measured 27Al MAS NMR spectra and the line shape fits based on the parameters listed in Table IV. The asymmetrical shape of each spectrum is characteristic of a wide distribution of quadrupolar coupling parameters. Most of the datasets could be fitted by assuming two distinct aluminum sites, with Czjzek line shape components having isotropic chemical shifts near 63 and 38 ppm. According to these chemical shifts, the first site is assigned to four-coordinated aluminum and the second site suggests the formation of five-coordinated aluminum.52,53 For the 25MgAS_25_50 sample, there is also evidence for a small fraction of six-coordinated aluminum.
The central transition in the 27Al MAS NMR spectra measured for the aluminosilicate glasses containing either (a) isotopically enriched 25Mg or (b) magnesium with its natural isotopic abundance. The spectra were deconvoluted using the Czjzek model.
The central transition in the 27Al MAS NMR spectra measured for the aluminosilicate glasses containing either (a) isotopically enriched 25Mg or (b) magnesium with its natural isotopic abundance. The spectra were deconvoluted using the Czjzek model.
27Al spectral parameters for the samples containing 25Mg vs natMg. MAS-Czjzek: Parameters obtained from a deconvolution analysis of the 27Al central-transition spectra using the Czjzek model for infinite spinning speed;41,45 average isotropic chemical shift , width of the Czjzek distribution σ, average quadrupolar coupling constant (defined in the sense of a modulus as per Ref. 56), relative intensity I, and average coordination number . SATRAS: Average isotropic chemical shift obtained from analysis of the spinning sidebands corresponding to satellite transitions. TQMAS: Average isotropic chemical shift and second-order quadrupolar effect obtained from the center of gravity of the TQMAS cross peaks using Eqs. (14) and (15).
MAS-Czjzek . | SATRAS . | TQMAS . | |||||||
---|---|---|---|---|---|---|---|---|---|
. | σ . | . | I . | . | . | . | |||
Glass . | Site . | (±1 ppm) . | (±0.1 MHz) . | (±0.1 MHz) . | (±2%) . | . | (±1 ppm) . | (±1 ppm) . | (±0.1 MHz) . |
natMgAS_37.5_50 | Al(IV) | 63 | 4.0 | 7.9 | 84 | 4.16(7) | ⋯ | 60 | 5.7 |
Al(V) | 38 | 4.1 | 8.1 | 16 | ⋯ | 31 | 3.2 | ||
25MgAS_37.5_50 | Al(IV) | 64 | 3.9 | 7.9 | 83 | 4.17(7) | 65 | 63 | 6.1 |
Al(V) | 38 | 4.1 | 8.1 | 17 | 38 | 36 | 4.7 | ||
natMgAS_25_50 | Al(IV) | 64 | 3.9 | 7.9 | 75 | 4.25(7) | ⋯ | 59 | 7.1 |
Al(V) | 39 | 4.1 | 8.1 | 25 | ⋯ | 31 | 2.9 | ||
25MgAS_25_50 | Al(IV) | 64 | 4.0 | 7.9 | 75 | 4.29(7) | 68 | 63 | 6.1 |
Al(V) | 38 | 3.9 | 7.9 | 21 | 39 | 36 | 4.7 | ||
Al(VI) | 15 | 4.1 | 8.1 | 4 | 13 | 11 | 4.6 | ||
natMgAS_30_60 | Al(IV) | 62 | 3.7 | 7.3 | 81 | 4.19(7) | ⋯ | 60 | 8.9 |
Al(V) | 37 | 4.3 | 7.9 | 19 | ⋯ | 30 | 6.4 | ||
25MgAS_30_60 | Al(IV) | 61 | 3.9 | 7.9 | 80 | 4.20(7) | 66 | 61 | 8.9 |
Al(V) | 38 | 4.6 | 9.1 | 20 | 42 | 30 | 6.4 | ||
natMgAS_20_60 | Al(IV) | 61 | 4.1 | 8.1 | 78 | 4.22(7) | ⋯ | 60 | 6.1 |
Al(V) | 36 | 4.0 | 7.9 | 22 | ⋯ | 31 | 3.7 | ||
25MgAS_20_60 | Al(IV) | 63 | 4.0 | 7.9 | 79 | 4.21(7) | 65 | 61 | 6.2 |
Al(V) | 36 | 4.0 | 7.9 | 21 | 39 | 35 | 3.5 |
MAS-Czjzek . | SATRAS . | TQMAS . | |||||||
---|---|---|---|---|---|---|---|---|---|
. | σ . | . | I . | . | . | . | |||
Glass . | Site . | (±1 ppm) . | (±0.1 MHz) . | (±0.1 MHz) . | (±2%) . | . | (±1 ppm) . | (±1 ppm) . | (±0.1 MHz) . |
natMgAS_37.5_50 | Al(IV) | 63 | 4.0 | 7.9 | 84 | 4.16(7) | ⋯ | 60 | 5.7 |
Al(V) | 38 | 4.1 | 8.1 | 16 | ⋯ | 31 | 3.2 | ||
25MgAS_37.5_50 | Al(IV) | 64 | 3.9 | 7.9 | 83 | 4.17(7) | 65 | 63 | 6.1 |
Al(V) | 38 | 4.1 | 8.1 | 17 | 38 | 36 | 4.7 | ||
natMgAS_25_50 | Al(IV) | 64 | 3.9 | 7.9 | 75 | 4.25(7) | ⋯ | 59 | 7.1 |
Al(V) | 39 | 4.1 | 8.1 | 25 | ⋯ | 31 | 2.9 | ||
25MgAS_25_50 | Al(IV) | 64 | 4.0 | 7.9 | 75 | 4.29(7) | 68 | 63 | 6.1 |
Al(V) | 38 | 3.9 | 7.9 | 21 | 39 | 36 | 4.7 | ||
Al(VI) | 15 | 4.1 | 8.1 | 4 | 13 | 11 | 4.6 | ||
natMgAS_30_60 | Al(IV) | 62 | 3.7 | 7.3 | 81 | 4.19(7) | ⋯ | 60 | 8.9 |
Al(V) | 37 | 4.3 | 7.9 | 19 | ⋯ | 30 | 6.4 | ||
25MgAS_30_60 | Al(IV) | 61 | 3.9 | 7.9 | 80 | 4.20(7) | 66 | 61 | 8.9 |
Al(V) | 38 | 4.6 | 9.1 | 20 | 42 | 30 | 6.4 | ||
natMgAS_20_60 | Al(IV) | 61 | 4.1 | 8.1 | 78 | 4.22(7) | ⋯ | 60 | 6.1 |
Al(V) | 36 | 4.0 | 7.9 | 22 | ⋯ | 31 | 3.7 | ||
25MgAS_20_60 | Al(IV) | 63 | 4.0 | 7.9 | 79 | 4.21(7) | 65 | 61 | 6.2 |
Al(V) | 36 | 4.0 | 7.9 | 21 | 39 | 35 | 3.5 |
To confirm the presence of five-coordinated aluminum, Fig. 8 shows representative SATellite TRansition Spectra (SATRAS) for the isotopically enriched samples. As previously discussed,57 the allowed m = ±1/2 ↔ m = ±3/2 Zeeman transitions of I = 5/2 nuclei are broadened by the anisotropy of the nuclear electric quadrupolar interactions. However, second-order effects on these Zeeman transitions are rather weak, such that magic angle spinning transforms the inhomogeneously broadened line shape into a manifold of sharp spinning sidebands separated from each other by multiples of the spinning frequency. Thus, owing to the absence of second-order quadrupolar broadening effects, the spectroscopic resolution for these satellite-MAS-peak patterns is superior to the resolution for the central-transition line shape, as observed in Fig. 8.
The 27Al SATRAS data for the isotopically enriched aluminosilicate glasses. For each glass composition, the third spinning sidebands on each side of the peak for the central-transition are shown. The spectra were deconvoluted using DMfit.39 Note that the improved resolution helps to distinguish between the four- and five-coordinated Al species.
The 27Al SATRAS data for the isotopically enriched aluminosilicate glasses. For each glass composition, the third spinning sidebands on each side of the peak for the central-transition are shown. The spectra were deconvoluted using DMfit.39 Note that the improved resolution helps to distinguish between the four- and five-coordinated Al species.
The 27Al MAS NMR results therefore support the presence of five-coordinated Al species in all the MgAS glasses of the present study, with a contribution to the area of a spectrum that varies between 16% and 25%. The results for the MgAS_20_60 glass are in good agreement with those previously obtained.50,58 The average coordination number can be deduced from the fractional areas found from the fitted spectra and are listed in Table IV. This information was used in fitting the neutron and x-ray diffraction data (Sec. IV C).
Figure 9 shows the complementary spectra measured by 27Al TQMAS NMR. They confirm the presence of both four- and five-coordinated Al, and the application of Eqs. (14) and (15) yields the average isotropic chemical shifts and second-order quadrupolar effects listed in Table IV. For four-coordinated Al, the and values obtained from TQMAS NMR are in reasonable agreement. The values extracted from TQMAS NMR tend to be smaller than the values extracted from the MAS-Czjzek fits because of the lower triple-quantum excitation efficiency observed for the Al species at the upper end of the quadrupole coupling constant distribution. For this reason, the values from the MAS-Czjzek fits are considered more representative. For five-coordinated Al, the values are considerably underestimated because of the low signal-to-noise ratio. For the isotopically enriched sample 25MgAS_25_50, the small peak at δ2 ≃ 5 ppm in the TQMAS spectrum indicates the presence of some six-coordinated Al, which is confirmed by our analysis of the MAS NMR spectrum. The amount of Al(VI) listed in Table IV may, however, be an overestimate because the presence of Al(VI) is not clear from the SATRAS data for this sample (Fig. 8). In general, for a given composition, the interaction parameters extracted from the 27Al NMR spectra measured for the isotopically enriched 25Mg vs natMg glasses are the same, within experimental errors.
The 27Al TQMAS NMR spectra for the aluminosilicate glasses containing either (a) isotopically enriched 25Mg or (b) magnesium with its natural isotopic abundance. Broken lines correspond to the diagonal of the 2D spectra where the δ1:δ2 ratio is 1:1.
The 27Al TQMAS NMR spectra for the aluminosilicate glasses containing either (a) isotopically enriched 25Mg or (b) magnesium with its natural isotopic abundance. Broken lines correspond to the diagonal of the 2D spectra where the δ1:δ2 ratio is 1:1.
3. 25Mg MAS and static NMR
Figure 10 shows the solid-state 25Mg MAS NMR spectra, acquired with a Hahn-echo sequence. The MAS NMR line shape for all the glasses is similar and is dominated by a wide chemical shift dispersion due to a distribution of quadrupolar coupling strengths whose effects upon the NMR line shape can be treated in the regime of second-order perturbation theory. The entire line shape could be successfully modeled using a Czjzek distribution, implemented within the ssNake program.45 The simulations were performed assuming the finite MAS rate regime, where 16 spinning sidebands were simulated by the Carousel averaging approach implemented in the software. An unconstrained fitting approach was used to determine the best-fit parameters. A Gaussian line-broadening (LB) parameter was employed to account for a distribution of isotropic chemical shifts.
The 25Mg MAS NMR spectra for the 25Mg enriched magnesium silicate and aluminosilicate glasses. The measured datasets, obtained by rotor-synchronized Hahn spin echo acquisition at a rotor frequency of 20.0 kHz (acquisition after two rotor cycles), are shown by black curves and the simulations obtained from a distribution of electric field gradients given by the Czjzek model are shown by red curves. The datasets for diopside and enstatite glasses are adopted from Ref. 8.
The 25Mg MAS NMR spectra for the 25Mg enriched magnesium silicate and aluminosilicate glasses. The measured datasets, obtained by rotor-synchronized Hahn spin echo acquisition at a rotor frequency of 20.0 kHz (acquisition after two rotor cycles), are shown by black curves and the simulations obtained from a distribution of electric field gradients given by the Czjzek model are shown by red curves. The datasets for diopside and enstatite glasses are adopted from Ref. 8.
The simulation parameters are listed in Table V, where the mean chemical shift takes a value in the range from 12 to 28 ppm. There is imprecision in the value of this parameter that originates from the large linewidth. In crystalline compounds, a chemical shift between zero and 30 ppm relative to aqueous MgCl2 solution is typical of magnesium in a six-coordinated environment Mg(VI), whereas a chemical shift of 40–80 ppm is typical of magnesium in a four-coordinated environment Mg(IV).9 The results therefore suggest that the Mg2+ ions in the different glasses are predominantly six-coordinated. As discussed in Sec. V, this interpretation is in stark contrast to the results found in the diffraction work.
Deconvolution analysis for the central transition of the 25Mg MAS NMR spectra, measured for the isotopically enriched glasses at a finite spinning rate of 20 kHz, using the Czjzek model as implemented in the ssNake software.45 The fitted parameters were unconstrained. LB is a line-broadening parameter corresponding to the convolution of the calculated spectrum with a Gaussian function. The datasets for diopside and enstatite glasses are taken from Ref. 8.
Glass . | (±5 ppm) . | σ (±0.2 MHz) . | (±0.4 MHz) . | LB (±0.5 kHz) . |
---|---|---|---|---|
Diopside | 13 | 4.2 | 8.3 | 2.0 |
Enstatite | 13 | 4.4 | 8.6 | 3.0 |
MgAS_37.5_50 | 28 | 4.5 | 8.9 | 2.5 |
MgAS_25_50 | 12 | 4.5 | 8.9 | 3.0 |
MgAS_30_60 | 13 | 4.7 | 9.3 | 2.0 |
MgAS_20_60 | 20 | 4.5 | 8.9 | 2.5 |
Glass . | (±5 ppm) . | σ (±0.2 MHz) . | (±0.4 MHz) . | LB (±0.5 kHz) . |
---|---|---|---|---|
Diopside | 13 | 4.2 | 8.3 | 2.0 |
Enstatite | 13 | 4.4 | 8.6 | 3.0 |
MgAS_37.5_50 | 28 | 4.5 | 8.9 | 2.5 |
MgAS_25_50 | 12 | 4.5 | 8.9 | 3.0 |
MgAS_30_60 | 13 | 4.7 | 9.3 | 2.0 |
MgAS_20_60 | 20 | 4.5 | 8.9 | 2.5 |
We have also recorded the static 25Mg NMR spectrum for the isotopically enriched sample 25MgAS_30_60 using the WURST-CPMG pulse sequence with a stepwise acquisition, with the carrier frequency varied in steps of 360 kHz (10 000 ppm). The spectrum is shown in Fig. 11 on two different scales, showing the central-transition line (left column) and the spikelet patterns visible on the high-frequency side of the central resonance (right column), which can be attributed to the satellite transition manifold. The line shape (spikelet envelope) is broadly consistent with a simulation based on the parameters obtained from the fit to the MAS NMR spectrum.
Left column: The static 25Mg WURST-CPMG spikelet spectra for the isotopically enriched sample 25MgAS_30_60 (black curve) and the simulation using the Czjzek distribution (red curve) obtained by fixing the parameter values at those obtained from the simulations of the 25Mg MAS NMR spectra (Fig. 10). Right column: Same spectrum and fit as in the left column but on an expanded scale, showing the resonances corresponding to the 25Mg satellite transitions. The arrows indicate the resonance offsets used for the stepwise acquisition of the wideline spectrum.
Left column: The static 25Mg WURST-CPMG spikelet spectra for the isotopically enriched sample 25MgAS_30_60 (black curve) and the simulation using the Czjzek distribution (red curve) obtained by fixing the parameter values at those obtained from the simulations of the 25Mg MAS NMR spectra (Fig. 10). Right column: Same spectrum and fit as in the left column but on an expanded scale, showing the resonances corresponding to the 25Mg satellite transitions. The arrows indicate the resonance offsets used for the stepwise acquisition of the wideline spectrum.
C. Neutron diffraction with isotope substitution
The scattering lengths used in the analysis of the neutron diffraction data for the glassy materials were bAl = 3.449(5) fm, bSi = 4.1491(10) fm, bO = 5.803(4) fm,26 = 5.375(4) fm (Table II), and taking into account the isotopic enrichment, = 3.739(12) fm for diopside, = 3.745(10) fm for enstatite, and = 3.729(12) fm for the MgAS glasses. The measured neutron total structure factors are shown in Fig. 12. The contrast in shape with the x-ray total structure factors (Fig. 1) reflects a difference in weighting of the Sαβ(k) functions.
The measured FN(k) functions for the enstatite, diopside, and MgAS glasses containing either natMg or 25Mg. For each glass, the points with black vertical error bars give the measured function and the red curve gives the back Fourier transform of the corresponding function (Figs. 18 and 19) after the low-r oscillations, below the distance of closest approach between two atoms, are set to the theoretical density-line limit. For each glass, the vertical error bars are smaller than the curve thickness at most k values.
The measured FN(k) functions for the enstatite, diopside, and MgAS glasses containing either natMg or 25Mg. For each glass, the points with black vertical error bars give the measured function and the red curve gives the back Fourier transform of the corresponding function (Figs. 18 and 19) after the low-r oscillations, below the distance of closest approach between two atoms, are set to the theoretical density-line limit. For each glass, the vertical error bars are smaller than the curve thickness at most k values.
The measured ΔFMg(k) functions are illustrated in Fig. 13 and show contrast between the total structure factors measured for each glass. The first peak in the corresponding functions (Fig. 14) is attributed to nearest-neighbor Mg–O correlations and encompasses typical Mg–O nearest-neighbor distances. For example, the Mg–O bond length is 1.923(1) Å for the MgO4 polyhedra of crystalline MgAl2O4,1 is in the range 2.09(9)–2.31(38) Å for the MgO6 polyhedra of crystalline MgO,2 MgSiO3,3 β-Mg2SiO4,4 and Mg0.5AlSiO4,5 and is 2.27(8) Å for the MgO8 polyhedra of crystalline Mg3Al2Si3O12.6 Unlike crystalline MgO (Fig. 4), the first Mg–O peak is broad and asymmetric and was therefore represented by two Gaussian functions. A third Gaussian function was used in the fitting procedure to constrain the high-r side of these Mg–O peaks.
The measured ΔFMg(k) functions for the enstatite, diopside, and MgAS glasses. For each glass, the points with black vertical error bars give the measured function and the red curve gives the back Fourier transform of the corresponding function (Fig. 14) after the low-r oscillations, below the distance of closest approach between two atoms, are set to the theoretical density-line limit.
The measured ΔFMg(k) functions for the enstatite, diopside, and MgAS glasses. For each glass, the points with black vertical error bars give the measured function and the red curve gives the back Fourier transform of the corresponding function (Fig. 14) after the low-r oscillations, below the distance of closest approach between two atoms, are set to the theoretical density-line limit.
The fitted functions for the enstatite, diopside, and MgAS glasses. In each panel, the black solid circles give the measured function, the black solid curve gives the fitted function, and the other curves show the contributions from the Mg–O (magenta solid curves) and Mg-β (β ≠ O) (cyan broken curve) correlations. The latter was introduced to constrain the peaks fitted at smaller r-values. The displaced green solid curve shows the residual.
The fitted functions for the enstatite, diopside, and MgAS glasses. In each panel, the black solid circles give the measured function, the black solid curve gives the fitted function, and the other curves show the contributions from the Mg–O (magenta solid curves) and Mg-β (β ≠ O) (cyan broken curve) correlations. The latter was introduced to constrain the peaks fitted at smaller r-values. The displaced green solid curve shows the residual.
The fitted functions are shown in Fig. 14. The fitted parameters are listed in Table VI along with the weighted mean bond distance given by
where gMgO(r) was obtained from the sum of the Gaussian functions fitted to the first peak and r1 and r2 define the overall extent of this peak. For each glass, the overall Mg–O coordination number (sum) obtained from the fitting procedure is in agreement with the value found directly from by using Eq. (10) to integrate over the first peak after the application of a Lorch modification function.59,60 The value of (sum) = 4.46(4) for enstatite glass is in accord with the value of 4.50(2) measured previously using neutron diffraction with magnesium isotope substitution.16
The Mg–O peak parameters obtained from the fitted functions for the diopside, enstatite, and MgAS glasses. The error associated with rMgO,i is typically ±0.004 Å (i = 1) or ±0.008 Å (i = 2). The error associated with σMgO,i is typically ±0.004 Å (i = 1) or ±0.007 Å (i = 2). The error associated with (i = 1 or 2) is typically ±0.03. The weighted mean Mg–O distance was obtained from the fitted peaks by applying Eq. (18) and has a typical error of ±0.008 Å. The overall Mg–O coordination number (sum) was obtained using the full precision of the measurements and has a typical error of ±0.04. The goodness-of-fit parameter Rχ is also given.
Glass . | First peak . | . | . | Second peak . | . | . | Overall . | . | . |
---|---|---|---|---|---|---|---|---|---|
rMgO, 1 (Å) . | σMgO, 1 (Å) . | . | rMgO, 2 (Å) . | σMgO, 2 (Å) . | . | (Å) . | (sum) . | Rχ . | |
Diopside | 1.975 | 0.050 | 2.61 | 2.143 | 0.069 | 1.80 | 2.036 | 4.40 | 0.074 |
Enstatite | 1.984 | 0.077 | 2.59 | 2.188 | 0.120 | 1.87 | 2.056 | 4.46 | 0.022 |
MgAS_37.5_50 | 2.003 | 0.092 | 3.17 | 2.186 | 0.152 | 1.59 | 2.051 | 4.76 | 0.042 |
MgAS_25_50 | 2.010 | 0.080 | 2.43 | 2.150 | 0.136 | 2.47 | 2.070 | 4.90 | 0.071 |
MgAS_30_60 | 2.004 | 0.098 | 3.40 | 2.204 | 0.157 | 1.36 | 2.047 | 4.77 | 0.040 |
MgAS_20_60 | 2.007 | 0.073 | 2.17 | 2.168 | 0.156 | 2.91 | 2.086 | 5.07 | 0.044 |
Glass . | First peak . | . | . | Second peak . | . | . | Overall . | . | . |
---|---|---|---|---|---|---|---|---|---|
rMgO, 1 (Å) . | σMgO, 1 (Å) . | . | rMgO, 2 (Å) . | σMgO, 2 (Å) . | . | (Å) . | (sum) . | Rχ . | |
Diopside | 1.975 | 0.050 | 2.61 | 2.143 | 0.069 | 1.80 | 2.036 | 4.40 | 0.074 |
Enstatite | 1.984 | 0.077 | 2.59 | 2.188 | 0.120 | 1.87 | 2.056 | 4.46 | 0.022 |
MgAS_37.5_50 | 2.003 | 0.092 | 3.17 | 2.186 | 0.152 | 1.59 | 2.051 | 4.76 | 0.042 |
MgAS_25_50 | 2.010 | 0.080 | 2.43 | 2.150 | 0.136 | 2.47 | 2.070 | 4.90 | 0.071 |
MgAS_30_60 | 2.004 | 0.098 | 3.40 | 2.204 | 0.157 | 1.36 | 2.047 | 4.77 | 0.040 |
MgAS_20_60 | 2.007 | 0.073 | 2.17 | 2.168 | 0.156 | 2.91 | 2.086 | 5.07 | 0.044 |
The measured ΔF(k) functions for the enstatite, diopside, and MgAS glasses are shown in Fig. 15 and the corresponding ΔD′(r) functions are shown in Fig. 16.
The measured ΔF(k) functions for the enstatite, diopside, and MgAS glasses. For each glass, the points with black vertical error bars give the measured function and the red curve gives the back Fourier transform of the corresponding ΔD′(r) function (Fig. 16) after the low-r oscillations, below the distance of closest approach between two atoms, are set to the theoretical density-line limit.
The measured ΔF(k) functions for the enstatite, diopside, and MgAS glasses. For each glass, the points with black vertical error bars give the measured function and the red curve gives the back Fourier transform of the corresponding ΔD′(r) function (Fig. 16) after the low-r oscillations, below the distance of closest approach between two atoms, are set to the theoretical density-line limit.
The fitted ΔD′(r) functions for the enstatite, diopside, and MgAS glasses. In each panel, the black solid circles give the measured function, the black solid curve gives the fitted function, and the other curves show the contributions from the Si–O (blue broken curve), Al–O (red solid curve, only for the MgAS glasses), O–O (green broken curve), and Ca–O (cyan solid curve, only for diopside) correlations. The displaced green solid curve shows the residual. The O–O correlations were introduced to constrain the peaks fitted at smaller r-values.
The fitted ΔD′(r) functions for the enstatite, diopside, and MgAS glasses. In each panel, the black solid circles give the measured function, the black solid curve gives the fitted function, and the other curves show the contributions from the Si–O (blue broken curve), Al–O (red solid curve, only for the MgAS glasses), O–O (green broken curve), and Ca–O (cyan solid curve, only for diopside) correlations. The displaced green solid curve shows the residual. The O–O correlations were introduced to constrain the peaks fitted at smaller r-values.
For enstatite glass, the first peak in ΔD′(r) originates solely from Si–O correlations and was fitted to a single Gaussian function. A second Gaussian function was used to fit the O–O correlations, which are expected at a mean distance of 2.65 Å for a tetrahedral SiO4 unit, where is the mean intra-tetrahedral Si–O bond distance. The fitted function is shown in Fig. 16 and the fitted parameters are listed in Table VII. The diffraction results are consistent with = 4, as found from 29Si MAS NMR experiments on enstatite glass.8,61,62
The Si–O and Al–O peak parameters obtained from the fitted ΔD′(r) functions for the diopside, enstatite, and MgAS glasses. Fixed parameters are marked with an asterisk, where the values were obtained from the 27Al MAS NMR experiments (Table IV). The errors associated with and σSiO are typically ±0.003 and ±0.005 Å, respectively, and the errors associated with and σAlO are typically ±0.005 Å. The goodness-of-fit parameter Rχ is also given.
Glass . | (Å) . | σSiO (Å) . | . | (Å) . | σAlO (Å) . | . | Rχ . |
---|---|---|---|---|---|---|---|
Diopsidea | 1.627 | 0.059 | 4.00* | ⋯ | ⋯ | ⋯ | 0.063 |
Enstatite | 1.617 | 0.059 | 4.00* | ⋯ | ⋯ | ⋯ | 0.115 |
MgAS_37.5_50 | 1.618 | 0.045 | 4.00* | 1.765 | 0.051 | 4.17* | 0.021 |
MgAS_25_50 | 1.620 | 0.051 | 4.00* | 1.766 | 0.072 | 4.27* | 0.021 |
MgAS_30_60 | 1.614 | 0.048 | 4.00* | 1.765 | 0.053 | 4.20* | 0.028 |
MgAS_20_60 | 1.614 | 0.047 | 4.00* | 1.768 | 0.065 | 4.22* | 0.026 |
Glass . | (Å) . | σSiO (Å) . | . | (Å) . | σAlO (Å) . | . | Rχ . |
---|---|---|---|---|---|---|---|
Diopsidea | 1.627 | 0.059 | 4.00* | ⋯ | ⋯ | ⋯ | 0.063 |
Enstatite | 1.617 | 0.059 | 4.00* | ⋯ | ⋯ | ⋯ | 0.115 |
MgAS_37.5_50 | 1.618 | 0.045 | 4.00* | 1.765 | 0.051 | 4.17* | 0.021 |
MgAS_25_50 | 1.620 | 0.051 | 4.00* | 1.766 | 0.072 | 4.27* | 0.021 |
MgAS_30_60 | 1.614 | 0.048 | 4.00* | 1.765 | 0.053 | 4.20* | 0.028 |
MgAS_20_60 | 1.614 | 0.047 | 4.00* | 1.768 | 0.065 | 4.22* | 0.026 |
Nearest-neighbor Ca–O correlations: = 2.374(5) Å, σCaO = 0.123(6) Å and = 5.01(5).
For the MgAS glasses, there is also a contribution to the first peak in ΔD′(r) from the Al–O correlations, which causes this peak to become more asymmetric as the alumina content of the glass increases (Fig. 16). These correlations were represented by an additional Gaussian function with set to the value found from the 27Al MAS NMR experiments (Table IV). There is little change to aluminum speciation with the glass composition, so there is little change to the mean Al–O distance (Table VII).
For diopside glass, in addition to the contributions to ΔD′(r) from nearest-neighbor Si–O and O–O correlations, there is also a contribution from nearest-neighbor Ca–O correlations. To help assess their effect, the functions were fitted using (i) a single Gaussian function for the nearest-neighbor Si–O correlations with the coordination number set at the value = 4 found from 29Si MAS NMR experiments,8,63 (ii) two Gaussian functions for the Mg–O correlations with the coordination numbers set at the values found from fitting (Table VI), and (iii) single Gaussian functions for both the Ca–O and O–O correlations. The x-ray datasets were used because of their enhanced sensitivity to the calcium pair-correlation functions: The atomic number of Ca (Z = 20) is larger than for the other chemical species present in the glass. The fitted function is shown in Fig. 17 and the fitted parameters are listed in Tables VIII and IX. A Ca–O coordination number of 5.01–5.06 was obtained with an associated bond distance of 2.353(5)–2.357(5) Å. The neutron ΔD′(r) function (Fig. 16) was then fitted following the procedure adopted for enstatite, but with an additional Gaussian function for the Ca–O correlations and constrained to the range of values found in the x-ray diffraction work. The results give with = 2.374(5) Å (Table VII), which compare to for crystalline diopside with Ca–O distances in the range 2.352–2.717 Å.64 The value for the glass may, however, be larger than indicated if the distribution of nearest neighbors is asymmetric, as found for the coordination environment of magnesium in the same material.
The fitted functions for the glasses investigated by x-ray diffraction. In each panel, the black solid circles give the function measured for the sample containing 25Mg, the black solid curve gives the fitted function, and the other curves show the contributions from the Si–O (blue broken curve), Al–O (red solid curve, only for the MgAS glasses), O–O (green broken curve), and Ca–O (cyan solid curve, only for diopside) correlations. The displaced green solid curve shows the residual. The O–O correlations were introduced to constrain the peaks fitted at smaller r-values.
The fitted functions for the glasses investigated by x-ray diffraction. In each panel, the black solid circles give the function measured for the sample containing 25Mg, the black solid curve gives the fitted function, and the other curves show the contributions from the Si–O (blue broken curve), Al–O (red solid curve, only for the MgAS glasses), O–O (green broken curve), and Ca–O (cyan solid curve, only for diopside) correlations. The displaced green solid curve shows the residual. The O–O correlations were introduced to constrain the peaks fitted at smaller r-values.
The Si–O and Al–O peak parameters obtained from the fitted functions for diopside and several of the MgAS glasses. Fixed parameters are marked with an asterisk, where the values were obtained from the 27Al MAS NMR experiments (Table IV). The errors associated with and σSiO are typically ±0.001 Å and the errors associated with and σAlO are typically ±0.003 Å. The goodness-of-fit parameter Rχ is also given.
Glass . | (Å) . | σSiO (Å) . | . | (Å) . | σAlO (Å) . | . | Rχ . |
---|---|---|---|---|---|---|---|
natDiopsidea | 1.625 | 0.060 | 4.00* | ⋯ | ⋯ | ⋯ | 0.095 |
25Diopsideb | 1.626 | 0.065 | 4.00* | ⋯ | ⋯ | ⋯ | 0.116 |
natMgAS_37.5_50 | 1.626 | 0.063 | 4.00* | 1.780 | 0.089 | 4.17* | 0.053 |
25MgAS_37.5_50 | 1.630 | 0.067 | 4.00* | 1.780 | 0.113 | 4.17* | 0.039 |
natMgAS_25_50 | 1.639 | 0.074 | 4.00* | 1.784 | 0.120 | 4.27* | 0.036 |
25MgAS_25_50 | 1.643 | 0.075 | 4.00* | 1.785 | 0.127 | 4.27* | 0.038 |
natMgAS_20_60 | 1.628 | 0.067 | 4.00* | 1.780 | 0.116 | 4.22* | 0.062 |
25MgAS_20_60 | 1.631 | 0.070 | 4.00* | 1.779 | 0.125 | 4.22* | 0.061 |
Glass . | (Å) . | σSiO (Å) . | . | (Å) . | σAlO (Å) . | . | Rχ . |
---|---|---|---|---|---|---|---|
natDiopsidea | 1.625 | 0.060 | 4.00* | ⋯ | ⋯ | ⋯ | 0.095 |
25Diopsideb | 1.626 | 0.065 | 4.00* | ⋯ | ⋯ | ⋯ | 0.116 |
natMgAS_37.5_50 | 1.626 | 0.063 | 4.00* | 1.780 | 0.089 | 4.17* | 0.053 |
25MgAS_37.5_50 | 1.630 | 0.067 | 4.00* | 1.780 | 0.113 | 4.17* | 0.039 |
natMgAS_25_50 | 1.639 | 0.074 | 4.00* | 1.784 | 0.120 | 4.27* | 0.036 |
25MgAS_25_50 | 1.643 | 0.075 | 4.00* | 1.785 | 0.127 | 4.27* | 0.038 |
natMgAS_20_60 | 1.628 | 0.067 | 4.00* | 1.780 | 0.116 | 4.22* | 0.062 |
25MgAS_20_60 | 1.631 | 0.070 | 4.00* | 1.779 | 0.125 | 4.22* | 0.061 |
Nearest-neighbor Ca–O correlations: = 2.357(5) Å, σCaO = 0.137(4) Å and = 5.06(12).
Nearest-neighbor Ca–O correlations: = 2.353(5) Å, σCaO = 0.142(3) Å and = 5.01(13).
The Mg–O peak parameters obtained from the fitted functions for diopside and several of the MgAS glasses. Fixed parameters are marked by an asterisk and were obtained from the measured functions (Table VI). The errors associated with rMgO,i and σMgO,i (i = 1, 2) are typically ±0.005 and ±0.007 Å, respectively. The weighted mean Mg–O distance was obtained from the fitted peaks by applying Eq. (18) and has a typical error of ±0.008 Å.
Glass . | First peak . | . | . | Second peak . | . | . | Overall . | . |
---|---|---|---|---|---|---|---|---|
rMgO, 1 (Å) . | σMgO, 1 (Å) . | . | rMgO, 2 (Å) . | σMgO, 2 (Å) . | . | (Å) . | (sum) . | |
natDiopside | 1.976 | 0.073 | 2.61* | 2.092 | 0.116 | 1.80* | 2.016 | 4.40* |
25Diopside | 1.970 | 0.071 | 2.61* | 2.112 | 0.134 | 1.80* | 2.019 | 4.40* |
natMgAS_37.5_50 | 2.015 | 0.076 | 3.17* | 2.208 | 0.089 | 1.59* | 2.069 | 4.76* |
25MgAS_37.5_50 | 2.020 | 0.088 | 3.17* | 2.218 | 0.107 | 1.59* | 2.074 | 4.76* |
natMgAS_25_50 | 2.063 | 0.068 | 2.43* | 2.200 | 0.114 | 2.47* | 2.124 | 4.90* |
25MgAS_25_50 | 2.072 | 0.074 | 2.43* | 2.197 | 0.125 | 2.47* | 2.127 | 4.90* |
natMgAS_20_60 | 2.053 | 0.050 | 2.17* | 2.200 | 0.099 | 2.91* | 2.129 | 5.07* |
25MgAS_20_60 | 2.063 | 0.064 | 2.17* | 2.200 | 0.114 | 2.91* | 2.133 | 5.07* |
Glass . | First peak . | . | . | Second peak . | . | . | Overall . | . |
---|---|---|---|---|---|---|---|---|
rMgO, 1 (Å) . | σMgO, 1 (Å) . | . | rMgO, 2 (Å) . | σMgO, 2 (Å) . | . | (Å) . | (sum) . | |
natDiopside | 1.976 | 0.073 | 2.61* | 2.092 | 0.116 | 1.80* | 2.016 | 4.40* |
25Diopside | 1.970 | 0.071 | 2.61* | 2.112 | 0.134 | 1.80* | 2.019 | 4.40* |
natMgAS_37.5_50 | 2.015 | 0.076 | 3.17* | 2.208 | 0.089 | 1.59* | 2.069 | 4.76* |
25MgAS_37.5_50 | 2.020 | 0.088 | 3.17* | 2.218 | 0.107 | 1.59* | 2.074 | 4.76* |
natMgAS_25_50 | 2.063 | 0.068 | 2.43* | 2.200 | 0.114 | 2.47* | 2.124 | 4.90* |
25MgAS_25_50 | 2.072 | 0.074 | 2.43* | 2.197 | 0.125 | 2.47* | 2.127 | 4.90* |
natMgAS_20_60 | 2.053 | 0.050 | 2.17* | 2.200 | 0.099 | 2.91* | 2.129 | 5.07* |
25MgAS_20_60 | 2.063 | 0.064 | 2.17* | 2.200 | 0.114 | 2.91* | 2.133 | 5.07* |
As a check on the internal consistency of the neutron diffraction datasets, each function was fitted using the results from the difference functions as starting parameters, subject to the constraints that , is given by the value found from the 27Al MAS NMR experiments, and for each peak is given by the value found from the corresponding function. For each glass composition, the functions (Figs. 18 and 19) could be represented with only minor adjustments to the peak positions (on average ≲0.6%) and peak widths (on average ≲3.9% for σSiO, σAlO and σMgO,1 vs ≲6.1% for σMgO, 2) (Tables X and XI). The fitted functions demonstrate the benefit of using neutron diffraction with magnesium isotope substitution in removing the overlap between (i) the Mg–O and Ca–O nearest-neighbor correlations for diopside and (ii) the Mg–O and Al–O correlations for the MgAS glasses.
The fitted functions for enstatite and two MgAS glasses along the 50 mol. % silica tie-line for the samples containing either natMg (left column) or 25Mg (right column). In each panel, the black solid circles give the measured function, the black solid curve gives the fitted function, and the other curves show the contributions from the Si–O (blue broken curve), Al–O (red solid curve, only for the MgAS glasses), Mg–O (magenta solid curves), and O–O (green broken curve) correlations. The displaced green solid curve shows the residual. The O–O correlations were introduced to constrain the peaks fitted at smaller r-values.
The fitted functions for enstatite and two MgAS glasses along the 50 mol. % silica tie-line for the samples containing either natMg (left column) or 25Mg (right column). In each panel, the black solid circles give the measured function, the black solid curve gives the fitted function, and the other curves show the contributions from the Si–O (blue broken curve), Al–O (red solid curve, only for the MgAS glasses), Mg–O (magenta solid curves), and O–O (green broken curve) correlations. The displaced green solid curve shows the residual. The O–O correlations were introduced to constrain the peaks fitted at smaller r-values.
The fitted functions for diopside and two MgAS glasses along the 60 mol. % silica tie-line for the samples containing either natMg (left column) or 25Mg (right column). In each panel, the black solid circles give the measured function, the black solid curve gives the fitted function, and the other curves show the contributions from the Si–O (blue broken curve), Al–O (red solid curve, only for the MgAS glasses), Mg–O (magenta solid curves), Ca–O (cyan solid curve, only for diopside), and O–O (green broken curve) correlations. The displaced green solid curve shows the residual. The O–O correlations were introduced to constrain the peaks fitted at smaller r-values.
The fitted functions for diopside and two MgAS glasses along the 60 mol. % silica tie-line for the samples containing either natMg (left column) or 25Mg (right column). In each panel, the black solid circles give the measured function, the black solid curve gives the fitted function, and the other curves show the contributions from the Si–O (blue broken curve), Al–O (red solid curve, only for the MgAS glasses), Mg–O (magenta solid curves), Ca–O (cyan solid curve, only for diopside), and O–O (green broken curve) correlations. The displaced green solid curve shows the residual. The O–O correlations were introduced to constrain the peaks fitted at smaller r-values.
The Si–O and Al–O peak parameters obtained from the fitted functions measured for diopside, enstatite, and the MgAS glasses. Fixed parameters are marked with an asterisk, where the values were obtained from the 27Al MAS NMR experiments (Table IV). The errors associated with and σSiO are typically ±0.003 and ±0.005 Å, respectively, and the errors associated with and σAlO are typically ±0.005 Å. The goodness-of-fit parameter Rχ is also given.
Glass . | (Å) . | σSiO (Å) . | . | (Å) . | σAlO (Å) . | . | Rχ . |
---|---|---|---|---|---|---|---|
natDiopsidea | 1.622 | 0.058 | 4.00* | ⋯ | ⋯ | ⋯ | 0.064 |
25Diopsideb | 1.624 | 0.058 | 4.00* | ⋯ | ⋯ | ⋯ | 0.054 |
natEnstatite | 1.616 | 0.058 | 4.00* | ⋯ | ⋯ | ⋯ | 0.087 |
25Enstatite | 1.617 | 0.058 | 4.00* | ⋯ | ⋯ | ⋯ | 0.074 |
natMgAS_37.5_50 | 1.618 | 0.044 | 4.00* | 1.773 | 0.052 | 4.17* | 0.037 |
25MgAS_37.5_50 | 1.618 | 0.044 | 4.00* | 1.772 | 0.052 | 4.17* | 0.026 |
natMgAS_25_50 | 1.619 | 0.050 | 4.00* | 1.767 | 0.072 | 4.27* | 0.029 |
25MgAS_25_50 | 1.620 | 0.047 | 4.00* | 1.771 | 0.071 | 4.27* | 0.013 |
natMgAS_30_60 | 1.614 | 0.048 | 4.00* | 1.775 | 0.052 | 4.20* | 0.031 |
25MgAS_30_60 | 1.614 | 0.048 | 4.00* | 1.774 | 0.052 | 4.20* | 0.025 |
natMgAS_20_60 | 1.615 | 0.047 | 4.00* | 1.775 | 0.065 | 4.22* | 0.022 |
25MgAS_20_60 | 1.615 | 0.047 | 4.00* | 1.775 | 0.065 | 4.22* | 0.019 |
Glass . | (Å) . | σSiO (Å) . | . | (Å) . | σAlO (Å) . | . | Rχ . |
---|---|---|---|---|---|---|---|
natDiopsidea | 1.622 | 0.058 | 4.00* | ⋯ | ⋯ | ⋯ | 0.064 |
25Diopsideb | 1.624 | 0.058 | 4.00* | ⋯ | ⋯ | ⋯ | 0.054 |
natEnstatite | 1.616 | 0.058 | 4.00* | ⋯ | ⋯ | ⋯ | 0.087 |
25Enstatite | 1.617 | 0.058 | 4.00* | ⋯ | ⋯ | ⋯ | 0.074 |
natMgAS_37.5_50 | 1.618 | 0.044 | 4.00* | 1.773 | 0.052 | 4.17* | 0.037 |
25MgAS_37.5_50 | 1.618 | 0.044 | 4.00* | 1.772 | 0.052 | 4.17* | 0.026 |
natMgAS_25_50 | 1.619 | 0.050 | 4.00* | 1.767 | 0.072 | 4.27* | 0.029 |
25MgAS_25_50 | 1.620 | 0.047 | 4.00* | 1.771 | 0.071 | 4.27* | 0.013 |
natMgAS_30_60 | 1.614 | 0.048 | 4.00* | 1.775 | 0.052 | 4.20* | 0.031 |
25MgAS_30_60 | 1.614 | 0.048 | 4.00* | 1.774 | 0.052 | 4.20* | 0.025 |
natMgAS_20_60 | 1.615 | 0.047 | 4.00* | 1.775 | 0.065 | 4.22* | 0.022 |
25MgAS_20_60 | 1.615 | 0.047 | 4.00* | 1.775 | 0.065 | 4.22* | 0.019 |
Nearest-neighbor Ca–O correlations: = 2.358(5) Å, σCaO = 0.156(6) Å and = 5.01*.
Nearest-neighbor Ca–O correlations: = 2.350(5) Å, σCaO = 0.150(6) Å and = 5.01*.
The Mg–O peak parameters obtained from the fitted functions for diopside, enstatite, and the MgAS glasses. Fixed parameters are marked by an asterisk and were obtained from the measured functions (Table VI). The error associated with rMgO,i is typically ±0.004 Å (i = 1) or ±0.007 Å (i = 2). The error associated with σMgO,i is typically ±0.004 Å (i = 1) or ±0.008 Å (i = 2). The weighted mean Mg–O distance was obtained from the fitted peaks by applying Eq. (18) and has a typical error of ±0.008 Å.
Glass . | First peak . | . | . | Second peak . | . | . | Overall . | . |
---|---|---|---|---|---|---|---|---|
rMgO,1 (Å) . | σMgO,1 (Å) . | . | rMgO,2 (Å) . | σMgO,2 (Å) . | . | (Å) . | (sum) . | |
natDiopside | 1.977 | 0.076 | 2.61* | 2.143 | 0.146 | 1.80* | 2.033 | 4.40* |
25Diopside | 1.978 | 0.074 | 2.61* | 2.143 | 0.200 | 1.80* | 2.031 | 4.40* |
natEnstatite | 1.984 | 0.084 | 2.59* | 2.115 | 0.140 | 1.87* | 2.029 | 4.46* |
25Enstatite | 1.982 | 0.087 | 2.59* | 2.084 | 0.130 | 1.87* | 2.016 | 4.46* |
natMgAS_37.5_50 | 2.009 | 0.086 | 3.17* | 2.185 | 0.153 | 1.59* | 2.056 | 4.76* |
25MgAS_37.5_50 | 2.010 | 0.086 | 3.17* | 2.178 | 0.155 | 1.59* | 2.055 | 4.76* |
natMgAS_25_50 | 2.022 | 0.074 | 2.43* | 2.148 | 0.153 | 2.47* | 2.075 | 4.90* |
25MgAS_25_50 | 2.036 | 0.065 | 2.43* | 2.159 | 0.179 | 2.47* | 2.086 | 4.90* |
natMgAS_30_60 | 2.006 | 0.084 | 3.40* | 2.212 | 0.105 | 1.36* | 2.053 | 4.77* |
25MgAS_30_60 | 2.009 | 0.081 | 3.40* | 2.215 | 0.100 | 1.36* | 2.056 | 4.77* |
natMgAS_20_60 | 2.023 | 0.055 | 2.17* | 2.175 | 0.144 | 2.91* | 2.099 | 5.07* |
25 MgAS_20_60 | 2.027 | 0.050 | 2.17* | 2.173 | 0.138 | 2.91* | 2.100 | 5.07* |
Glass . | First peak . | . | . | Second peak . | . | . | Overall . | . |
---|---|---|---|---|---|---|---|---|
rMgO,1 (Å) . | σMgO,1 (Å) . | . | rMgO,2 (Å) . | σMgO,2 (Å) . | . | (Å) . | (sum) . | |
natDiopside | 1.977 | 0.076 | 2.61* | 2.143 | 0.146 | 1.80* | 2.033 | 4.40* |
25Diopside | 1.978 | 0.074 | 2.61* | 2.143 | 0.200 | 1.80* | 2.031 | 4.40* |
natEnstatite | 1.984 | 0.084 | 2.59* | 2.115 | 0.140 | 1.87* | 2.029 | 4.46* |
25Enstatite | 1.982 | 0.087 | 2.59* | 2.084 | 0.130 | 1.87* | 2.016 | 4.46* |
natMgAS_37.5_50 | 2.009 | 0.086 | 3.17* | 2.185 | 0.153 | 1.59* | 2.056 | 4.76* |
25MgAS_37.5_50 | 2.010 | 0.086 | 3.17* | 2.178 | 0.155 | 1.59* | 2.055 | 4.76* |
natMgAS_25_50 | 2.022 | 0.074 | 2.43* | 2.148 | 0.153 | 2.47* | 2.075 | 4.90* |
25MgAS_25_50 | 2.036 | 0.065 | 2.43* | 2.159 | 0.179 | 2.47* | 2.086 | 4.90* |
natMgAS_30_60 | 2.006 | 0.084 | 3.40* | 2.212 | 0.105 | 1.36* | 2.053 | 4.77* |
25MgAS_30_60 | 2.009 | 0.081 | 3.40* | 2.215 | 0.100 | 1.36* | 2.056 | 4.77* |
natMgAS_20_60 | 2.023 | 0.055 | 2.17* | 2.175 | 0.144 | 2.91* | 2.099 | 5.07* |
25 MgAS_20_60 | 2.027 | 0.050 | 2.17* | 2.173 | 0.138 | 2.91* | 2.100 | 5.07* |
Finally, the functions for the MgAS glasses were fitted using the same strategy as for the neutron total pair-distribution functions, i.e., by setting , to the value measured by 27Al MAS NMR, and for each of the Gaussian peaks to the value found from . Again, the measured functions (Fig. 17) could be fitted with only minor adjustments to the peak widths and positions (Tables VIII and IX). As compared to the functions, there is increased overlap between the Mg–O and O–O correlations, which emphasizes the benefit of using neutron diffraction with isotope substitution to gain unambiguous information on the coordination environment of magnesium.
V. DISCUSSION
In all the investigated glasses, the measured functions show that the local coordination environment of magnesium is characterized by a broad and asymmetric distribution of Mg–O distances (Fig. 14), indicating a wide distribution of coordination environments for the Mg2+ ions. This finding is consistent with molecular dynamics studies of enstatite glass where Mg2+ is found to be four-, five-, or six-coordinated, and the distortion of the polyhedral units is quantified using an orientational order parameter.14 The variety leads to a broad distribution of electric field gradient components at the magnesium position, as inferred from 25Mg solid-state NMR experiments on the diopside and enstatite glasses8 and confirmed by the asymmetry of the spectra measured for the MgAS glasses in this work. It follows from the diffraction study that the majority of Mg2+ ions are not six-coordinated, in contrast to the expectation originating from the chemical shifts observed in unconstrained fits to the 25Mg MAS NMR spectra (Table V) and in contrast to the findings of previous work using 25Mg MAS65–67 or TQMAS NMR spectroscopy68,69 on the same or related silicate glasses.
At the heart of this issue is the strong effect on the solid-state 25Mg NMR spectra of a wide distribution of quadrupolar coupling strengths. This distribution reflects the variation in local magnesium coordination environments and results in excessive loss of site-specific resolution in the measured spectra. In the case of enstatite glass, for example, the MAS NMR spectrum could be adequately fitted by constraining the chemical shift to 50 ppm, which is typical of four-coordinated Mg species.8 A satisfactory fit to the spectrum could also be obtained by assuming a 1:1 superposition of Mg species with chemical shifts of 13 and 50 ppm to mimic magnesium in a five-coordinated environment. The present study confirms that, for the 25Mg MAS NMR spectra collected at 14.1 T with a magic angle spinning rate of 20 kHz, the distribution of Mg2+ coordination numbers cannot be revealed. In future work, this issue might be addressed by much higher magnetic flux densities (28.2 T) and spinning speeds (50 kHz). In the end, the ultimate resolution limit will be dominated by the width of the distribution of the isotropic chemical shifts associated with each of the Mg coordination states, such as Mg(IV), Mg(V), and Mg(VI). If these widths are much larger than the differences between the respective average shifts of these coordination environments, no peak resolution will be observable, no matter how high the magnetic field strength. Furthermore, other resolution limitations associated with chemical shift anisotropy could enter at higher magnetic field strengths.
The Mg–O coordination number increases with the weighted mean Mg–O distance as the alumina content is increased along either the 50 or 60 mol. % silica tie-line (Table VI). For example, along the 50 mol. % silica tie-line, = 4.46(4) for enstatite, which corresponds to the alumina free end-member composition, and increases as magnesia is replaced by alumina to reach the value = 4.90(4) in MgAS_25_50 glass when R = 1. In contrast, Guignard and Cormier70 applied the reverse Monte Carlo method to the neutron and x-ray diffraction datasets measured for four different MgAS glasses along the 50 mol. % silica tie-line and found little change from with the glass composition.
It should be noted that the values reported in this work account for the area under the first peak and shoulder in the measured functions, extending to the first minimum at rmin ≃ 2.5 Å. More remote oxygen atoms may, however, produce a large-r tail to gMgO(r) that contributes toward the Mg-β peak near 2.8 Å in (Fig. 14). These oxygen neighbors may also affect the 25Mg NMR chemical shifts. In the case of enstatite glass, the measured function does not suggest a pronounced tail, as emphasized by plotting the function
which takes the value of zero at rmin (Fig. 20). Nevertheless, molecular dynamics simulations of enstatite glass using an aspherical ion model14 do show a tail on the high-r side of the first peak in gMgO(r) and contributions toward the Mg-β peak from the Mg–O, Mg–Si, and Mg–Mg correlations (Fig. 20).
The ΔGMg(r) function for enstatite glass as measured using neutron diffraction with magnesium isotope substitution (circles) or simulated using molecular dynamics (black curve).14 The contributions to the simulated function from the weighted gMgO(r), gMgSi(r), and gMgMg(r) functions are also given.
The ΔGMg(r) function for enstatite glass as measured using neutron diffraction with magnesium isotope substitution (circles) or simulated using molecular dynamics (black curve).14 The contributions to the simulated function from the weighted gMgO(r), gMgSi(r), and gMgMg(r) functions are also given.
In diopside and the other investigated glasses, Mg–O coordination numbers of less than five indicate the presence of four-coordinated Mg2+ ions. For example, if only four- and five-coordinated species are present in the glass structure, and if fMg(IV) denotes the fraction of four-coordinated Mg2+ ions, it follows that the mean coordination number . In the case of diopside, the measured value = 4.40(4) then implies that 60% of the magnesium is four-coordinated.
There is, however, no evidence in support of four-coordinated magnesium adopting a network-forming role (Sec. I). For example, the Qn speciation found from 29Si MAS NMR experiments on glassy diopside,8,63 enstatite,8,61,62 and wollastonite (CaSiO3)63,71,72 is similar for all three materials, giving a mean value . Hence, there is nothing to distinguish between the structural roles played by Ca2+, which is usually regarded as a network modifier, and Mg2+. Additionally, the entropy of mixing found from viscosity measurements on the pyroxene CaSiO3–MgSiO3, garnet Ca3Al2Si3O12–Mg3Al2Si3O12, and anorthite CaAl2Si2O8–MgAl2Si2O8 systems, in which Ca2+ was systematically replaced by Mg2+, is consistent with an ideal mixing hypothesis for these cations.73,74 The measurements do not, therefore, indicate any substantial difference between the structural roles played by the Ca2+ and Mg2+ ions.
As noted in Sec. IV B 1, the Mg2+ ions in the diopside and enstatite glasses generate Si-NBO linkages and therefore adopt a network-modifying role. When alumina is introduced to the glass network, a fraction of the Mg2+ ions will also adopt a predominantly charge-compensating role by balancing the charge on the [AlØ4/2]− species.50 Along a tie-line with constant mol. % silica, the fraction of Mg2+ ions that adopt a network-modifying role will therefore decrease as magnesia is replaced by alumina. The diffraction results indicate, therefore, a difference in size between the coordination environments of Mg2+ ions that take a network-modifying vs a predominantly charge-compensating role. In this context, it is important to point out that, for the alumina containing glasses, a given Mg2+ ion may well have in its coordination sphere both Si-NBO atoms and the oxygen atoms from [AlØ4/2]− units and, therefore, play a dual network-modifying/charge-compensating role. The anionic charge on this Al(IV) unit may be compensated by multiple Mg2+ ions, each contributing a small part of the required charge and each, likewise, interacting with multiple oxygen atoms.
VI. CONCLUSIONS
Neutron powder diffraction was used to remeasure the bound coherent neutron scattering length of the magnesium isotope 25Mg. The value = 3.720(12) fm was obtained, which is 2.8% larger than the literature value.18
The method of neutron diffraction with isotope substitution was used to measure the structure of diopside and enstatite composition glasses, two MgAS glasses along the 50 mol. % silica tie-line with R = 3 or R = 1, and two MgAS glasses along the 60 mol. % silica tie-line with the same R-values. As compared to the literature value for ,18 the new scattering length led to much improved self-consistency between the measured difference functions and their back Fourier transforms (Sec. II B). The diffraction results reveal a broad and asymmetric distribution of Mg–O nearest neighbors. In the diopside and enstatite glasses, where the Mg2+ ions adopt a network-modifying role, is 4.40(4) and 4.46(4), respectively. As magnesia is replaced by alumina along a 50 or 60 mol. % silica tie-line, the mean Mg–O coordination number increases with the weighted Mg–O bond distance as Mg2+ ions also adopt the role of charge-compensators to the [AlØ4/2]− units, particularly in those glasses for which R = 1.
The glass structure was also investigated by 25Mg MAS and static NMR experiments. The measured spectra could be represented by employing the Czjzek distribution model using an identical set of parameters for a given glass composition, which lends support to the strategy employed to interpret the NMR results.8 However, the chemical shifts obtained from unconstrained fits of the measured MAS NMR spectra appear to be more consistent with an Mg–O coordination number of six for all the investigated glasses, rather than a value in the range 4.40(4)–5.07(4) measured by diffraction. This dichotomy finds explanation in the broad distribution of electric field gradients associated with the different sites of the Mg2+ ions that originates from structural disorder. In the investigated glasses, the concomitant quadrupolar broadening of the measured spectra makes it impossible to resolve the contributions from the different magnesium species, at least under the field strength (14.1 T) and spinning rate (20 kHz) employed in the MAS NMR experiments.
The present work demonstrates the power of neutron diffraction with magnesium isotope substitution to provide unambiguous site-specific information on the glass structure. In the investigated materials, it removed the uncertainties associated with overlap between (i) the Mg–O and Al–O correlations in the or functions for the MgAS glasses, (ii) the Mg–O and Ca–O correlations in the or functions for diopside glass, and (iii) the Mg–O and O–O correlations in the functions for all the glasses. The results thereby clarify the way in which these total pair-distribution functions should be interpreted, which will empower systematic investigations of the structure of MgAS glasses over a broader composition range to explore, e.g., the network-modifying vs charge-compensating role of the Mg2+ ions in more detail.
ACKNOWLEDGMENTS
We thank Kristina Brix and Ralf Kautenburger (Saarland University, Saarbrücken, Germany) for performing the ICP-MS experiments and Alain Bertoni (Grenoble, France) for help with the D4c experiments.
H.M. was supported by Corning Inc. (Agreement No. CM00002159/SA/01). R.M.D.S. was supported by the Royal Society (Grant No. RGF/EA/180060). A.Z. was supported by a Royal Society-EPSRC Dorothy Hodgkin Research Fellowship. L.V.D.G. acknowledges funding and support from the EPSRC Centre for Doctoral Training in Condensed Matter Physics (CDT-CMP), Grant No. EP/L015544/1, the Science and Technology Facilities Council (STFC) and Diamond Light Source Ltd. (Reference No. STU0173). P.S.S. and A.Z. are grateful to Corning Inc. for the award of Gordon S. Fulcher Distinguished Scholarships under which this work was conceived. Funding of this work by FAPESP, Project No. 2013/07793-6, is gratefully acknowledged. M.d.O., Jr. and H.E. acknowledge the National Council for Scientific and Technological Development (CNPq, Grant Nos. 311069/2020-7 and 310870/2020-8, respectively). H.D. acknowledges personal fellowship support from CNPq and Nippon Electric Glass. This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No. DE-AC02-06CH11357. We acknowledge use of the Inorganic Crystal Structure Database accessed via the Chemical Database Service funded by the Engineering and Physical Sciences Research Council (EPSRC) and hosted by the Royal Society of Chemistry.
P.S.S., A.Z., R.E.Y., and B.G.A. designed the diffraction project on the glass structure. H.K. and H.E.F. designed the project to measure the neutron scattering lengths. L.C. provided the isotopically enriched samples of diopside and enstatite and the raw data sets for the enstatite glass. H.M. prepared the other glass samples and F.G. prepared the crystalline MgO samples. F.G. and H.E.F. performed the neutron powder diffraction experiments and F.G. and H.M. analyzed the results. M.d.O., Jr., H.D., and H.E. performed the NMR experiments and analyzed the results. A.Z., H.M., R.M.D.S., H.E.F., and P.S.S. performed the neutron diffraction experiments on the glasses and H.M., R.M.D.S., L.V.D.G., and A.Z. analyzed the results. C.J.B. performed the x-ray diffraction experiments and R.M.D.S. analyzed the results. P.S.S., M.d.O., Jr., and H.E. wrote the paper with input from all coauthors.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Hesameddin Mohammadi: Formal analysis (equal); Investigation (equal); Writing – review & editing (supporting). Rita Mendes Da Silva: Formal analysis (equal); Investigation (equal); Writing – review & editing (supporting). Anita Zeidler: Conceptualization (equal); Data curation (supporting); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Software (equal); Supervision (equal); Writing – review & editing (supporting). Lawrence V. D. Gammond: Formal analysis (supporting); Investigation (supporting); Writing – review & editing (supporting). Florian Gehlhaar: Formal analysis (equal); Investigation (equal). Marcos de Oliveira, Jr.: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Software (equal); Visualization (equal); Writing – original draft (supporting); Writing – review & editing (supporting). Hugo Damasceno: Formal analysis (supporting); Investigation (supporting). Hellmut Eckert: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Validation (equal); Writing – original draft (equal); Writing – review & editing (equal). Randall E. Youngman: Conceptualization (equal); Funding acquisition (equal). Bruce G. Aitken: Conceptualization (equal); Funding acquisition (equal). Henry E. Fischer: Conceptualization (supporting); Investigation (supporting); Supervision (equal). Holger Kohlmann: Conceptualization (supporting); Supervision (equal). Laurent Cormier: Resources (supporting). Chris J. Benmore: Investigation (supporting). Philip S. Salmon: Conceptualization (equal); Data curation (lead); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Software (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (lead); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are openly available in University of Bath Research Data Archive at https://doi.org/10.15125/BATH-01192.75 The measured neutron diffraction datasets for crystalline 25MgO and natMgO are available from Ref. 76. The measured neutron diffraction datasets for the glasses are available from Refs. 77–79.