Oxygen Speciation in Multicomponent Silicate ... - ACS Publications

May 5, 2017 - the glass structure is composed of two regions: network structure and breakage region ..... for their technical support with the NMR exp...
0 downloads 0 Views 1MB Size
Subscriber access provided by University of Florida | Smathers Libraries

Letter

Oxygen Speciation in Multicomponent Silicate Glasses Using Through Bond Double Resonance NMR Spectroscopy Sohei Sukenaga, Pierre Florian, Koji Kanehashi, Hiroyuki Shibata, Noritaka Saito, Kunihiko Nakashima, and Dominique Massiot J. Phys. Chem. Lett., Just Accepted Manuscript • Publication Date (Web): 05 May 2017 Downloaded from http://pubs.acs.org on May 6, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry Letters is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Oxygen Speciation in Multicomponent Silicate Glasses Using Through Bond Double Resonance NMR Spectroscopy Sohei Sukenaga1*, Pierre Florian2*, Koji Kanehashi3, Hiroyuki Shibata1, Noritaka Saito4, Kunihiko Nakashima4 and Dominique Massiot2 1

Institute of Multidisciplinary Research for Advanced Materials (IMRAM), Tohoku University,

2-1-1, Katahira, Aoba-ku, Sendai 9808577, Japan 2

CEMHTI UPR3079, CNRS, Université d’Orléans, 1D avenue de la Recherche Scientifique

45071 Orléans cedex 2, France 3

Materials Characterization Research Lab., Advanced Technology Research Laboratories,

Nippon Steel & Sumitomo Metal Corporation (NSSMC), 20-1 Shintomi, Futtsu, 2938511, Japan 4

Department Materials Science and Engineering, Kyushu University, 744 Motooka Nishi-ku

Fukuoka, 8190935, Japan *e-mail addresses of corresponding authors: [email protected] (S.S.) and [email protected] (P.F.)

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 24

ABSTRACT

The description of the structure of aluminosilicate glasses is more than often centered on its cationic constituents and oxygen ions determine their connectivity, directly impacting the physical properties of those disordered materials. A very powerful approach to ascertain this short- to medium-range order is to use

17

O NMR but up to now the speciation of the chemical

bonds was only ambiguously achieved for multicomponent glasses. Here, we propose to directly probe the very scarcely explored through-bond correlations using

17

O{27Al} and

17

O{23Na}

solid-state nuclear magnetic resonance (NMR) double-resonance experiments. Our approach allows quantifying the strongly overlapping components of the 17O NMR spectra of a quaternary aluminosilicate glass. We observe a cooperative location of alkali and aluminum ions in the neighborhood of bridging oxygens, which is consistent with the modified random network model where the glass structure is composed of two regions: network structure and breakage region (i.e. channel). (147/150 words)

TOC GRAPHICS

ACS Paragon Plus Environment

2

Page 3 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Silicate glasses are widely used in everyday’s life but are also candidate systems for advanced materials (e.g. biomaterials, solid-state electrolytes1) as well as by-products of steel-making industries, such as quenched metallurgical slags2. When new glass materials are designed, it is important to capture the features of their non-crystalline structure, which can be strongly linked to their macroscopic properties. Greaves3 proposed the modified random network (MRN) model, which divides the non-crystalline structure into two regions: network structure region and channel region (see Figure 1). The network structure region consists of framework species (e.g. SiO4, AlO4 tetrahedra) connected by bridging oxygens (BOs). The channel region is the breakage region of the structure composed of non-bridging oxygens (NBOs) and non-framework cations (e.g. alkali and alkaline-earth cations), which act as network modifier. Non-framework cations can also be incorporated in the network structure region as charge compensators when there are negative charges on the BOs, e.g. BOs between Si4+ and Al3+ (Si-OBO-Al).4 If this partitioning concept is accepted, it is possible to interpret the properties of the glass materials as the sum of the properties of the two regions. Recently, combining X-ray and/or neutron scattering experiments with reverse Monte Carlo simulation techniques gave a precise 3-dimentional structures for single- and two-component oxide glasses (e.g. silica glass,5-6 potassium-silicate glass7). However, most practical glass materials are multicomponent systems2, 8 to control their properties (e.g. melting temperature and melt’s viscosity). As of now, it is difficult to determine their precise structure by diffraction-based techniques alone because of the increasing number of overlapping pair distribution functions for the multicomponent system.

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 24

_______________________________________________________________________

Figure 1. Schematic illustrations of network structure and channel region in a plane. The Network structure is composed of framework species Si and M (e.g. SiO4 and AlO4 tetrahedra). The channel is the breakage region of the non-crystalline structure. Non-framework cations are disturbed both in the channel and in the network structure region as network modifiers and charge compensators to Si-O-M bridging oxygen, respectively.

_____________________________________________________________________________ Probing the detailed chemical state of oxygen atoms will help us know which types of cations form bonds with BOs and NBOs, and

17

O solid-state NMR spectroscopy is one of the most

appropriate techniques to detect this short- to medium-range order.

17

O NMR spectroscopy has

been applied to a large number of model two or three component silicate glasses and the amounts of NBOs and BOs have been successfully quantified.9-10 Recently, 17O NMR experiments11 were applied to quaternary oxide glasses, but molecular dynamics simulations had to be used to

ACS Paragon Plus Environment

4

Page 5 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

interpret the experimental spectra. To the best of our knowledge, there have been no quantitative direct experimental studies showing the cationic distribution in multicomponent silicate glasses. Double-resonance NMR techniques, requiring doubly tunable probes, are methods of choice to disentangle complex spectra. These experiments can be carried out using dipolar through-space interaction using a variety of sequences,12-13 with the difficulty of sorting out the few close-by nuclei from the many remote coupled ones and performing complex radio-frequency based manipulations of quadrupolar spins.14 Alternatively through-bond indirect J-coupling existing between chemically bonded spin bearing nuclei can be used but its magnitude is usually small.15 Heading for a two quadrupolar spin system, we decided to investigate sequences based on scalar couplings J which directly explore the chemically bonded units and can be applied to quadrupolar nuclei using heteronuclear multiple quantum correlation (HMQC) experiments16-18 or refocused insensitive nuclei enhanced by polarization transfer (INEPT)19-21 which require only few rf-pulses and have received much less practical attention. To adapt this sequence to the presence of quadrupolar nuclei,22 we carefully choose low radio-frequency fields to ensure central-transition selective pulses23-24 and performed full synchronization with the rotor’s rotation. These J-based through-bond approach has been seldom used for inorganic materials15, 17, 25

whereas through space dipolar-recoupling approaches13, 26-28 have been largely preferred.

In the present study, we examined the cationic distribution in the neighborhood of BOs and NBOs by applying a variety of

17

O NMR techniques to a ternary calcium aluminosilicate glass

and a quaternary sodium calcium aluminosilicate glass, which are typical systems of man-made glasses29 and quenched metallurgical slags.30

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 24

_______________________________________________________________________ Table 1. Nominal and examined composition$ (in mol%) of the glass samples used for the NMR experiments.

Component (mol%) Sample CAS

NCAS

CaO

SiO2

Al2O3

Na2O

Nominal

36.5

51.0

12.5

-

Examined

37.2 ± 0.3

50.3 ± 0.2

12.5 ± 0.2

-

Nominal

31.9

44.7

12.6

10.8

Examined

32.3 ± 0.3

44.0 ± 0.6

13.0 ± 0.1

10.8 ± 0.3

$

Chemical compositions of the glasses were examined 5 times using wavelength dispersive X-ray spectrometry (WDS); averaged compositions are presented in this table.

_______________________________________________________________________ We synthesised

17

O-enriched calcium aluminosilicate (CAS) and sodium calcium

aluminosilicate (NCAS) glasses by the melt and quench method (detailed procedure is shown in the “Experimental methods” section). Powder X-ray diffraction (XRD) and electron-probemicro-analyzer (EPMA) ensure their glassy nature and chemical composition (see Figure S1 and Table 1). In addition, the 29Si and

27

Al solid-state NMR experiments (see Figures S4 and S5)

showed expected behaviors with only small amounts of aluminium in five-fold coordination detected.

ACS Paragon Plus Environment

6

Page 7 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

_______________________________________________________________________

Figure 2.

17

O solid-state NMR spectra of the CAS and the NCAS glasses obtained at 20.0 T. In all

figures, original observed data are displayed as thick grey lines, while the sum of the simulated lines are described as black dashed lines. The coloured lines represent the Gaussian fit curves for each component. a, 17O single-resonance spectrum of the CAS glass. b, 17O{27Al} double-resonance spectrum of the CAS glass. c, 17O single-resonance spectrum of the NCAS glass. d, 17O{23Na} double-resonance spectrum of the NCAS glass. e, 17O{27Al} double-resonance spectrum of the NCAS glass.

_____________________________________________________________________________ Figure 2a shows the

17

O NMR single-resonance spectrum of the CAS glass. This spectrum

shows two main peaks around 110 ppm and at 55 ppm. The former is close to that of the NBO signals (i.e. Si-ONBO:Ca) reported in calcium aluminosilicate glasses31-32 while the latter is related to BOs31-32 but cannot determine from the this spectrum which type of oxygen atoms bond with the aluminum. The through-bond

17

O{27Al} double-resonance (INEPT) spectrum of

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the CAS glass (Figure 2b) displays the signal of the

17

Page 8 of 24

O atoms directly bonded to aluminum

atoms and a unique signal is found at ~53 ppm, in the BOs range. It indicates that the majority of the aluminum atoms have bonds with BOs only while NBOs link with silicon and calcium cations. It clearly means that the aluminum cations are in the network structure region. According to the Loewenstein rule,33 aluminosilicate glass can have two major types of BOs (Si-OBO-Si and Si-OBO-Al) and a minor species of Al-OBO-Al34-35. The 17O triple-quantum (3Q) MAS NMR experiment of the CAS glass confirmed this with undetectable amount of Al-OBO-Al (see Figure S2) as well as line shapes dominated at 20.0 T by chemical shift distributions (i.e. Gaussian). So the peak position and width derived for the Si-ONBO:Ca and Si-OBO-Al species were used to simulate the 17O quantitative single-resonance spectrum adding a Gaussian line for the Si-OBO-Si species, as shown in Figure 2a (the detailed fitting procedure is described in the supporting information). The area fraction of the simulated curve for the Si-ONBO:Ca signal (28 %) is in excellent agreement with the nominal NBO concentration (28.2 ± 0.5 %) derived from the composition, validating our methodology. We experimentally confirm that aluminum atoms create Si-OBO-Al species charge compensated by calcium. Consequently, in the CAS glass, the aluminium cations are concentrated in the neighborhood of BOs while calcium cations are distributed near both BOs and NBOs.

ACS Paragon Plus Environment

8

Page 9 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

__________________________________________________________________________

Figure 3. Experimental and simulated

17

O{27Al} INEPT excitation build-up for the CAS glass. a,

Schematics of the pulse sequence used in this study. b, Experimental (colored lines) and simulated (dashed grey lines) intensity as a function of τ’. c, Experimental (blue dots) and calculated (dotted black line)

17

O{27Al} INEPT spectra for the excitation delays used in this study, along with the undistorted

Gaussian line (red line). Calculations assume a linear correlation between δiso and J and absence of quadrupolar broadening (see text for details).

_____________________________________________________________________________ Before going to the quaternary oxide glass, we discuss the possible J-based edited spectrum’s limitations. Using selective irradiation conditions we don’t expect any impact of the quadrupolar interaction on the evolution under the scalar couplings.22 On the other hand, the disordered nature of the materials studied here produces a distribution of NMR interactions including scalar couplings. It shows in the build-up of the

17

O{27Al} INEPT intensity as a function the second

delay (figure 3a) of this sequence where the zero crossings are strongly dependent upon the

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

position in the

17

Page 10 of 24

O line (figure 3b). For an ISn spin-1/2 system displaying a unique scalar

coupling JIS the INEPT intensity as a function of time τ’ (see figure 3a) classically36 goes as  ⁄ sin 2′ cos 2′

(1)

where T2 is the relaxation time of the heteronuclear coherences. With n = 1 (i.e. no Al-O-Al linkages, vide supra) and no second quadrupolar broadening (see

17

O MQMAS figure S3) the

first zero crossing appears at ′ = 1⁄2 and hence its spread evidences a distribution of scalar couplings. We can moreover see a strong correlation between 1⁄2 and the position  in the Gaussian

17

O line, a situation reminiscent of the linear dependence observed between 1H-13C

scalar coupling and carbon chemical shift in proteins.37 Assuming such a linear trend  ! = " + $ which we introduce in equation (1) we can calculate the 17O{27Al} INEPT build-up of a Gaussian line shape (position %& and full width at half maximum Δ(/). The results of the simulation using n = 1 displayed on figure 3b appears to account very well for the observed build-up, validating the assumptions used and leading to δav = 54.0 ppm, ∆ν1/2 = 24.0 ppm, T2 = 28.1 ms, A = -0.469 Hz/ppm and B = 61.9 Hz. Interestingly, the slope A (the first report we are aware of for inorganic compounds) is consistent with the one observed for JCH in RNA (0.710 Hz/ppm) and proteins (0.365 Hz/ppm) underlying a very general common origin of both isotropic chemical shifts and scalar couplings. The most probable JAlO is found at 35 Hz (see also supplementary section S7) but its distribution implies that a single set of (τ, τ’) excitation times may not equally excite the full line shape. This is illustrated in figure 3c where the excitation times use here (i.e. τ = τ’ = 5 ms) lead to a small lack of intensity on left, high frequency, side of the spectrum. It is hence expected that our reconstruction of the

17

O spectrum is slightly

overestimating the amount of Si-O-Si environments but the simulation of figure 3c suggests that the error is reasonable small.

ACS Paragon Plus Environment

10

Page 11 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

We now turn our attention to the quaternary oxide glass (NCAS) and figure 2c shows the 17O single-resonance spectrum of the NCAS glass. The two main peaks, assigned to the Si-ONBO:Ca (at 110 ppm) and the BOs (Si-OBO-Si and Si-OBO-Al, at ~55 ppm), are visible in this sodiumcontaining system, too. However, the depth of the dip around 80 ppm is shallower than in the CAS glass, indicating the presence of additional component(s). Similarly, the width of the BOs’ signal is wider than that of the CAS glass, indicating that NCAS glass has a wider variety of oxygen types than CAS glass because of the presence of sodium cations. The

17

O{23Na} NMR

INEPT experiments (Figure 2d) shows the oxygen bonded to sodium cations and displays two signals around 85 ppm and 40 ppm. The former could be assigned to NBOs linked to at least one sodium cation and has a broader line shape (peak width = 40 ppm) than that of Si-ONBO:Ca (31 ppm). In general, NBO could be coordinated by 3 (or 4) modifier cations as well as Si4+ in the silicate glasses38-39 forming Si-O:(CaxNay) species. In this NCAS glass, the breadth of the NBONa signal could be due to peak overlap of those various Si-O:(CaxNay) environments (labelled as Si-ONBO:(Ca, Na) in Figures 2c and 2d). On the other hand, the peak at 40 ppm is close to the BOs’ signal observed in sodium aluminosilicate glasses.39 This signal should hence contain SiOBO-Si and Si-OBO-Al, which bonds with sodium atoms (i.e. Si-OBO-Si:Na and Si-OBO-Al:Na). Consequently, the 17O{23Na} NMR experiments show that sodium cations bond with both NBOs and BOs. Similar to the case of CAS glass, the 17O{27Al} signal of NCAS is found at ~50 ppm, indicating that almost all of the aluminium atoms have bonds with only BOs (Figure 2e). The 17

O{27Al} signal of NCAS is broader than that of CAS, showing that Si-OBO-Al:Na were created

with the addition of sodium oxide. Using peak positions and widths obtained from our doubleresonance experiments, a fit of the

17

O single-resonance spectrum of the NCAS glass by six

Gaussian curves reproduces its line shape and the total area fraction of NBOs (36 %) agrees well

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 24

with the nominal values (35.4 ± 0.8 %) obtained from the chemical composition (for detailed fitting procedure see supporting information). Here, the area fraction of signals from the BOs bonding with sodium cations (14 %) is larger than that from Si-ONBO:(Ca, Na) (10 %). According to Stebbins’ estimation4 the average negative charges on Si-OBO-Al and NBO are -0.25 and -1 respectively. Assuming that Si-OBO-Al:Na bonds with one sodium atom while Si-ONBO:(Ca, Na) is coordinated by one sodium and two calcium atoms on average,40 the ratio of the positive charge amount of sodium cations used for BOs to that for NBOs is approximately 0.54/0.46 while this ratio for calcium cations is approximately 0.25/0.75 when sodium and calcium cations are in six-fold and seven-fold coordinations41 (for details see supporting information section S9). These ratios show that the average BO/NBO ratios in the sodium and calcium coordination spheres are approximately 3.5/2.5 and 1.9/5.1 with the assumption that the negative charges of Si-OBO-Al:Ca and Si-OBO-Si(unlinked

with Na)

are fully compensated by calcium cations, clearly

indicating that the positive charges of sodium cations are used to preferentially compensate the negative charges on Si-OBO-Al compared to calcium cations. This behavior has been inferred from recent studies,11, 39, 41 but we quantified it directly from a measurement of NBO species’ population and without relying on molecular dynamics simulations. This preferential cationic behavior is also expected for Ca/Mg aluminosilicate glasses.42-43 If the channel and network region separately exists, sodium cations have a stronger locational preference than calcium cations in the Al/Si network along with aluminum while calcium cations have a greater tendency than sodium cations to be in the channels formed by NBOs located on silicon cations, as pictorially described in Figure 4. A recent study44 shows that ionic conductivity (i.e., mobility of alkali atom) in alkali and alkaline-earth aluminosilicate glass is higher when that alkali atom acts as a charge compensator to BOs, compared to the case of alkali atoms concentrated in the

ACS Paragon Plus Environment

12

Page 13 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

channel as modifiers. Therefore, the preference of sodium toward the network structure will obviously impact macroscopic properties such as ionic conductivity of melts at elevated temperatures. Here, we show that the through-bond NMR double resonance technique performed at a high magnetic field allows direct observation and quantification of the various species present in the glass. Specifically, the 17O-based experiment clearly evidenced non-random cationic distribution between in the neighborhood of BOs and NBOs, which is a key to understanding the chemical and physical properties of glasses and melts. Moreover, this methodology can be used to investigate the structure of complex oxide glasses, which is necessary for creating advanced glass materials with high performance. (2228/2500 words)

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 24

_______________________________________________________________________

Figure 4. Pictorial description of a possible microstructure of the NCAS glass in a plane. Fewer aluminium cations have bonds with NBOs. Sodium cations have a stronger preference than calcium cations to be in the network structure region while calcium cations have a greater tendency than sodium cations to be in the channel region.

____________________________________________________________________________

Experimental methods

ACS Paragon Plus Environment

14

Page 15 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Sample preparation. The chemical compositions of the glass samples are described in Table 1. CAS and NCAS glasses were made by melting mixtures of CaCO3, Al2O3, Na2CO3 (Sigma Aldrich, Inc.) and 17O enriched SiO2 at 1873 K for 90 min under an Ar atmosphere (the mixture was wrapped in a Pt foil). The

17

O enriched SiO2 was prepared by using 40% enriched H217O

(ISOTEC) with SiCl4 (Sigma Aldrich, Inc.) according to the literature protocol.45 The melts were then quenched on a copper plate. The quenched vitreous samples were crushed into powder and used for the NMR experiments.

17

O single- and double-resonance experiment. NMR spectra were collected on a Bruker

Avance 850 MHz spectrometer at the magnetic field of 20 T with Larmor frequencies of 221.67 MHz (27Al), 224.98 MHz (23Na), and 115.28 MHz (17O). Samples were spun at the magic angle in a 4 mm or 3.2 mm thin-wall ZrO2 rotor at a spinning speed of 14 or 22 kHz. 27Al, 23Na and 17

O chemical shifts were referenced to 1 M

27

Al(NO3)3 solution, 1 M

23

NaCl solution, and tap

water (H217O), respectively. 17

O single-resonance NMR spectra were collected using a rotor synchronised full-echo

acquisition pulse sequence; the lengths of delays were set to 1 ms. The length of the selective 90° pulse (T90) and 180° pulses (T180) for the 17O nuclei in the solid samples were found at 11.5 and 23.0 µs for CAS glass (12.0 and 24.0 µs for NCAS glass), respectively. Recycle delays for the CAS and the NCAS glass were 20 s. 17

O{27Al} and

17

O{23Na} through-bond double-resonance signals were obtained using the

refocused INEPT19-20 pulse sequence presented in Figure 3a. INEPT evolution delays (τ and τ’) were synchronized with the rotation speed of 22 kHz. The period of these delays was experimentally optimized taking into account the spin-spin relaxation time (T2) of each nucleus

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 24

(T2(27Al) ≈ 50 ms, T2(23Na) ≈ 45 ms, T2(17O) ≈ 160 ms (BO) and 200 ms (NBO)) and the INEPT signal intensity variation with τ’ (see supporting information Figure S6 and S7). Selective T90 of the 27Al 17O, 23Na and 29Si were estimated to be 8.5 µs, 12 µs, 13 µs, and 13.7 µs, respectively. The recycle delays for all INEPT experiments were 1 s.

ASSOCIATED CONTENT Supporting Information The following files are available free of charge at ACS publication website. XRD pattern of the samples, additional results of NMR experiment (MQMAS NMR, J-coupling coupling constant and so on), and procedure to determine the experimental condition (PDF)

AUTHOR INFORMATION Corresponding Authors *E-mail: [email protected] (S.S.) *E-mail: [email protected] (P.F.)

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT

ACS Paragon Plus Environment

16

Page 17 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

S.S. and H.S. acknowledge financial support from the Japan Society for Promotion of Science (JSPS) as the institutional program for young researcher overseas visit (JSPS Clean Energy) and JSPS KAKENHI Grant Number J16H04543. Also, S.S. would like to thank the Faculty of Engineering, Kyushu University for financial support. The authors (K.N. and H.S.) are grateful for financial support from the Cooperative Research Program of the “Network Joint Research Center for Materials and Devices”, IMRAM Tohoku University. We are grateful to Dr. Vincent Sarou-Kanian and Dr. Aydar Rakhmatullin, CEMHTI-CNRS, for their technical support with the NMR experiments. We would like to thank Mr. Masanori Tashiro, Tohoku University, for the EPMA analysis on the glass samples.

AUTHOR CONTRIBUTION S.S., P.F., and D.M. designed and directed the present study. S.S. synthesized the 17O-enriched glass samples. P.F. developed an idea on the INEPT experiment for the spin pairs of 17O{27Al} and

17

O{23Na}. P.F. and S.S. contributed the NMR experiments and analyzed the experimental

results. K.K. conducted the 17O 3QMAS NMR experiments at 11.4 T at NSSMC (Futtsu, Japan). S.S., P.F., D.M, H.S., K.K., N.S., and K.N. considered the implication of the structural information. S.S. and P.F. mainly wrote this manuscript. All the authors commented on this manuscript.

REFERENCES

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 24

(1) Du, J. C.; Chen, C. H. Structure and Lithium Ion Diffusion in Lithium Silicate Glasses and at Their Interfaces with Lithium Lanthanum Titanate Crystals. J. Non-Cryst. Solids 2012, 358, 3531-3538.

(2) Rawlings, R. D.; Wu, J. P.; Boccaccini, A. R. Glass-Ceramics: Their Production from Wastes-a Review. J. Mater. Sci. 2006, 41, 733-761.

(3) Greaves, G. N. EXAFS and the Structure of Glass. J. Non-Cryst. Solids 1985, 71, 203-217.

(4) Stebbins, J. F.; Wu, J. S.; Thompson, L. M. Interactions between Network Cation Coordination and Non-Bridging Oxygen Abundance in Oxide Glasses and Melts: Insights from Nmr Spectroscopy. Chem. Geol. 2013, 346, 34-46.

(5) Ohno, H.; Kohara, S.; Umesaki, N.; Suzuya, K. High-Energy X-Ray Diffraction Studies of Non-Crystalline Materials. J. Non-Cryst. Solids 2001, 293, 125-135.

(6) Kohara, S.; Suzuya, K. Intermediate-Range Order in Vitreous SiO2 and GeO2. J. Phys.: Condens. Matter 2005, 17, S77-S86.

(7) Majerus, O.; Cormier, L.; Calas, G.; Beuneu, B. A Neutron Diffraction Study of Temperature-Induced Structural Changes in Potassium Disilicate Glass and Melt. Chem. Geol. 2004, 213, 89-102.

ACS Paragon Plus Environment

18

Page 19 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

(8) Phillips, J. C.; Kerner, R. Structure and Function of Window Glass and Pyrex. J. Chem. Phys. 2008, 128.

(9) Stebbins, J. F.; Xu, Z. NMR Evidence for Excess Non-Bridging Oxygen in an Aluminosilicate Glass. Nature 1997, 390, 60-62.

(10) Stebbins, J. F.; Oglesby, J. V.; Lee, S. K. Oxygen Sites in Silicate Glasses: A New View from Oxygen-17 NMR. Chem. Geol. 2001, 174, 63-75.

(11) Pedone, A.; Gambuzzi, E.; Menziani, M. C. Unambiguous Description of the Oxygen Environment in Multicomponent Aluminosilicate Glasses from O-17 Solid State NMR Computational Spectroscopy. J. Phys. Chem. C 2012, 116, 14599-14609.

(12) Hu, B.; Trebosc, J.; Amoureux, J. P. Comparison of Several Hetero-Nuclear Dipolar Recoupling Nmr Methods to Be Used in MAS HMQC/HSQC. J. Magn. Reson. 2008, 192, 112122.

(13) Eckert, H.; Elbers, S.; Epping, J. D.; Janssen, M.; Kalwei, M.; Strojek, W.; Voigt, U. Dipolar Solid State Nmr Approaches Towards Medium-Range Structure in Oxide Glasses. New Techniques in Solid-State NMR 2005, 246, 195-233.

(14) Ashbrook, S. E.; Wimperis, S. Spin-Locking of Half-Integer Quadrupolar Nuclei in Nuclear Magnetic Resonance of Solids: Second-Order Quadrupolar and Resonance Offset Effects. J. Chem. Phys. 2009, 131.

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 24

(15) Massiot, D.; Fayon, F.; Deschamps, M.; Cadars, S.; Florian, P.; Montouillout, V.; Pellerin, N.; Hiet, J.; Rakhmatullin, A.; Bessada, C. Detection and Use of Small J Couplings in Solid State Nmr Experiments. C. R. Chim. 2010, 13, 117-129.

(16) Muller, L. Sensitivity Enhanced Detection of Weak Nuclei Using Heteronuclear Multiple Quantum Coherence. J. Am. Chem. Soc. 1979, 101, 4481-4484.

(17) Iuga, D.; Morais, C.; Gan, Z. H.; Neuville, D. R.; Cormier, L.; Massiot, D. NMR Heteronuclear Correlation between Quadrupolar Nuclei in Solids. J. Am. Chem. Soc. 2005, 127, 11540-11541.

(18) Lesage, A.; Sakellariou, D.; Steuernagel, S.; Emsley, L. Carbon-Proton Chemical Shift Correlation in Solid-State Nmr by through-Bond Multiple-Quantum Spectroscopy. J. Am. Chem. Soc. 1998, 120, 13194-13201.

(19) Morris, G. A.; Freeman, R. Enhancement of Nuclear Magnetic-Resonance Signals by Polarization Transfer. J. Am. Chem. Soc. 1979, 101, 760-762.

(20) Fyfe, C. A.; Wongmoon, K. C.; Huang, Y.; Grondey, H. Inept Experiments in Solid-State Nmr. J. Am. Chem. Soc. 1995, 117, 10397-10398.

(21) Xue, X. Y. Determination of J Coupling Constants between Spin-1/2 and Quadrupolar Nuclei in Inorganic Solids from Spin Echo and Refocused INEPT Experiments: A Case Study on Alpo4 Berlinite. Solid State Nucl. Magn. Reson. 2010, 38, 62-73.

ACS Paragon Plus Environment

20

Page 21 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

(22) Amoureux, J. P.; Trebosc, J.; Wiench, J.; Pruski, M. HMQC and Refocused-INEPT Experiments Involving Half-Integer Quadrupolar Nuclei in Solids. J. Magn. Reson. 2007, 184, 114.

(23) Vega, S. Fictitious Spin 1/2 Operator Formalism for Multiple Quantum NMR. J. Chem. Phys. 1978, 68, 5518-5527.

(24) Smith, M. E.; van Eck, E. R. H. Recent Advances in Experimental Solid State NMR Methodology for Half-Integer Spin Quadrupolar Nuclei. Prog. Nucl. Magn. Reson. Spectrosc. 1999, 34, 159-201.

(25) Tricot, G.; Raguenet, B.; Silly, G.; Ribes, M.; Pradel, A.; Eckert, H. P-O-B-3 Linkages in Borophosphate Glasses Evidenced by High Field B-11/P-31 Correlation NMR. Chem. Commun. 2015, 51, 9284-9286.

(26) van Wullen, L.; Tricot, G.; Wegner, S. An Advanced NMR Protocol for the Structural Characterization of Aluminophosphate Glasses. Solid State Nucl. Magn. Reson. 2007, 32, 44-52.

(27) Ren, J. J.; Zhang, L.; Eckert, H. Medium-Range Order in Sol-Gel Prepared Al2o3-Sio2 Glasses: New Results from Solid-State Nmr. J. Phys. Chem. C 2014, 118, 4906-4917.

(28) LaComb, M.; Rice, D.; Stebbins, J. F. Network Oxygen Sites in Calcium Aluminoborosilicate Glasses: Results from O-17{Al-27} and O-17{B-11} Double Resonance NMR. J. Non-Cryst. Solids 2016, 447, 248-254.

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 24

(29) Moir, G. K.; Glasser, F. P. Phase-Equilibria in Glass-Forming Region of System Na2OCaO-Al2O3-SiO2. Phys. Chem. Glasses 1976, 17, 45-53.

(30) Li, M.; Utigard, T.; Barati, M. Kinetics of Na2O and B2O3 Loss from CaO-SiO2-Al2O3 Slags. Metal. Mater. Trans. B 2015, 46, 74-82.

(31) Allwardt, J. R.; Lee, S. K.; Stebbins, J. F. Bonding Preferences of Non-Bridging O Atoms: Evidence from O-17 MAS and 3QMAS NMR on Calcium Aluminate and Low-Silica CaAluminosilicate Glasses. Am. Mineral. 2003, 88, 949-954.

(32) Stebbins, J. F.; Dubinsky, E. V.; Kanehashi, K.; Kelsey, K. E. Temperature Effects on NonBridging Oxygen and Aluminum Coordination Number in Calcium Aluminosilicate Glasses and Melts. Geochim. Cosmochim. Acta 2008, 72, 910-925.

(33) Loewenstein, W. The Distribution of Aluminum in the Tetrahedra of Silicates and Aluminates. Am. Mineral. 1954, 39, 92-96.

(34) Lee, S. K.; Stebbins, J. F. The Degree of Aluminum Avoidance in Aluminosilicate Glasses. Am. Mineral. 1999, 84, 937-945.

(35) Lee, S. K.; Stebbins, J. F. Al-O-Al and Si-O-Si Sites in Framework Aluminosilicate Glasses with Si/Al=1: Quantification of Framework Disorder. J. Non-Cryst. Solids 2000, 270, 260-264.

ACS Paragon Plus Environment

22

Page 23 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

(36) Keeler, J. Understanding NMR Spectroscopy. Second Edition ed.; John Wiley & Sons Ltd: The Atrium, Chichester, West Sussex, PO19 8SQ, United Kingdom, 2010; p 511.

(37) Zwahlen, C.; Legault, P.; Vincent, S. J. F.; Greenblatt, J.; Konrat, R.; Kay, L. E. Methods for Measurement of Intermolecular Noes by Multinuclear NMR Spectroscopy: Application to a Bacteriophage Lambda N-Peptide/Boxb Rna Complex. J. Am. Chem. Soc. 1997, 119, 6711-6721.

(38) Florian, P.; Vermillion, K. E.; Grandinetti, P. J.; Farnan, I.; Stebbins, J. F. Cation Distribution in Mixed Alkali Disilicate Glasses. J. Am. Chem. Soc. 1996, 118, 3493-3497.

(39) Lee, S. K.; Sung, S. The Effect of Network-Modifying Cations on the Structure and Disorder in Peralkaline Ca-Na Aluminosilicate Glasses: O-17 3QMAS NMR Study. Chem. Geol. 2008, 256, 326-333.

(40) Lee, S. K.; Stebbins, J. F. Nature of Cation Mixing and Ordering in Na-Ca Silicate Glasses and Melts. J. Phys. Chem. B 2003, 107, 3141-3148.

(41) Cormier, L.; Neuville, D. R. Ca and Na Environments in Na2O-CaO-Al2O3-SiO2 Glasses: Influence of Cation Mixing and Cation-Network Interactions. Chem. Geol. 2004, 213, 103-113.

(42) Kelsey, K. E.; Allwardt, J. R.; Stebbins, J. F. Ca-Mg Mixing in Aluminosilicate Glasses: An Investigation Using O-17 MAS and 3QMAS and Al-27 MAS NMR. J. Non-Cryst. Solids 2008, 354, 4644-4653.

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 24

(43) Neuville, D. R.; Cormier, L.; Montouillout, V.; Florian, P.; Millot, F.; Rifflet, J. C.; Massiot, D. Structure of Mg- and Mg/Ca Aluminosilicate Glasses: Al-27 NMR and Raman Spectroscopy Investigations. Am. Mineral. 2008, 93, 1721-1731.

(44) Amma, S.; Lanagan, M. T.; Kim, S. H.; Pantano, C. G. Ionic Conductivity in SodiumAlkaline Earth-Aluminosilicate Glasses. J. Am. Ceram. Soc. 2016, 99, 1239-1247.

(45) Geissberger, A. E.; Bray, P. J. Determinations of Structure and Bonding in Amorphous SiO2 Using O-17 NMR. J. Non-Cryst. Solids 1983, 54, 121-137.

ACS Paragon Plus Environment

24