Next Article in Journal
Assessing Hydrogen Embrittlement in Pipeline Steels for Natural Gas-Hydrogen Blends: Implications for Existing Infrastructure
Next Article in Special Issue
Synthesis and Crystal Structure of Ilmenite-Type Silicate with Pyrope Composition
Previous Article in Journal
Taxifolin Adsorption on Nitrogenated Graphenes: Theoretical Insights
Previous Article in Special Issue
High-Pressure Synthesis of the Iodide Carbonate Na5(CO3)2I
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of the Structure of Zn and Na Borophosphate Glasses Using X-ray and Neutron Scattering Techniques

1
Institut für Physik, Universität Rostock, D-18051 Rostock, Germany
2
Department of Materials Science and Engineering, Missouri University of Science and Technology, Rolla, MO 65409, USA
3
Deutsches Elektronen-Synchrotron DESY, D-22603 Hamburg, Germany
4
ISIS Facility, Rutherford Appleton Laboratory, Chilton, Didcot, Oxon OX11 0QX, UK
*
Author to whom correspondence should be addressed.
Current address: Institute of Physics, Faculty of Science, Pavol Jozef Šafárik University, 040 01 Košice, Slovakia.
Solids 2024, 5(3), 355-374; https://doi.org/10.3390/solids5030024
Submission received: 26 April 2024 / Revised: 7 June 2024 / Accepted: 20 June 2024 / Published: 1 July 2024

Abstract

:
The atomic structures of Zn and Na borophosphate glasses were studied using X-ray and neutron scattering techniques. Peaks assigned to the B−O, P−O, and O−O distances confirm that only BO4 units co-exist with the PO4 tetrahedra. The Zn−O and Na−O coordination numbers are found to be a little larger than four. The narrowest peaks of the Zn−O first-neighbor distances exist for the glasses along a line connecting the Zn(PO3)2 and BPO4 compositions (50 mol% P2O5), which is explained by networks of ZnO4, BO4, and PO4 tetrahedra with twofold coordinated oxygens. The calculated amounts of available oxygen support this interpretation. Broadened peaks occur for glasses with lower P2O5 contents, which is consistent with the presence of threefold coordinated oxygens. The two distinct P−O peak components of the Zn and Na borophosphate glasses differ in their relative abundances. This is interpreted as follows: Na+ cations coordinate oxygens in some P−O−B bridges, which is something not seen for the Zn2+ ions.

Graphical Abstract

1. Introduction

The addition of B2O3 to phosphate glasses improves their chemical durability [1] and makes these glasses suitable for a variety of applications. Dissolution rates and thermal stability are optimized for biomedical use [2]. No tendency toward devitrification was observed in the rare-earth doped fibers of Na/Nb borophosphate glasses [3]. Borophosphate glasses with alkali modifiers are of particular interest due to their promising ionic conductivities and large glass-forming ranges extending into the P2O5- or B2O3-rich compositional regions [4]. The observation of maxima in ionic conductivity, for example, in the glasses (A2O)0.35[(B2O3)x(P2O5)1−x]0.65 at x = ~0.4, gave rise to the introduction of the term mixed network-former effect [5] (A2O—alkali oxide). A striking feature of the mixed glassy networks with compositions rich in P2O5 is that the boron forms exclusively tetrahedral BO4 units with the corners mostly linked to phosphate tetrahedra but also to some neighboring BO4 units [6,7,8,9,10,11]. The higher cationic mobility with increasing x is accompanied by increasing cross-link densities of the borophosphate networks. At the B2O3-rich end, the BO4 fractions decrease, replaced by BO3 units, to those values known according to the boron anomaly of binary alkali borate glasses [12].
The BO4 formation in the P2O5-rich glasses with predominantly B−O−P bridges in the tetrahedral corners expresses the usual effect of additional oxygen in the phosphate networks, namely the disruption of P−O−P bridges [13]. The BPO4 crystals [14,15] illustrate the mutual benefit for the PO4 and BO4 tetrahedra. They possess balanced valences of 1.25 and 0.75 vu (valence units) in their four bonds with P−O and B−O bond lengths of ~0.152 and ~0.147 nm, respectively. The few known borophosphate crystals, e.g., Na3BP2O8 [16], Na5B2P3O13 [17], and Zn3BO3PO4 [18], have structures where the phosphate networks are already fully dissolved (no P−O−P bridges). Crystalline analogs for the glassy borophosphate networks are missing in the regions of samples that are rich in P2O5.
The present work aims to clarify the structural role of ZnO in the borophosphate networks and compare that role to the one played by Na2O in two similar compositions. The glass-forming range of binary ZnO-B2O3 glasses is small [19] and ternary glasses with small P2O5 contents (<30 mol%) are difficult to obtain [13,20,21,22]. Raman spectroscopy, X-ray photoelectron spectroscopy (XPS), high-pressure liquid chromatography (HPLC), and 31P or 11B MAS NMR have provided detailed information on the structural evolution of the Zn-borophosphate networks whose chemical compositions are given in Figure 1a. B−O−B bridges alongside P−O−P bridges contradict the usual scheme of fully depolymerized phosphate networks [23]. This behavior was explained by the stronger acidity of the glass-former B2O3 [13], when compared, e.g., with Al2O3 in phosphate networks [24]. Most studies on the mixed network-former effect use a series of constant modifier contents, [6,10,11] as shown in Figure 1a. Other series have fixed O/P ratios, e.g., for the Zn alumino- and borophosphate glasses [13,24]. The O/P ratios determine the mean number of PO4 corners in P−O−P bridges provided that all oxygen belongs to at least one PO4.
For describing the borophosphate networks, the labels P m B n and B m P n were introduced to describe the PO4 and BO4 (BO3) units with the bridges P−O−P, B−O−B, and P−O−B [6,7,8]. The superscript n is the total number of bridging corners of the P- or B-centered units, while the subscript m gives the number of heteroatomic bridges. The corresponding fractions of groups can be widely resolved by 31P and 11B MAS NMR [6,7,8,9,10,11] and were useful for developing quantitative descriptions of the compositional dependence of the network structures. These models, however, are not used to analyze the Zn- and Na-centered oxygen polyhedra in the borophosphate glasses, which is the focus of the present work. Here, the critical parameter will be the number of P−O bonds that do not link to neighboring PO4 tetrahedra, which will ultimately determine the coordination environments of the Zn- and Na-ions. Therefore, besides the P m B n and B m P n units, the PO4 units will still be characterized by the common Qn nomenclature, where the superscript n gives the number of PO4 corners in P−O−P bridges with possible n = 3, 2, 1, or 0 [23]. The oxygens of binary phosphate glasses are divided into OB (bridging in P−O−P) and ONB (non-bridging in other bonds). In detail, the OB atoms of the borophosphate glasses can form P−O−P, P−O−B, and B−O−B bridges, where some of them have modifier cations in their vicinity. A complete description of all combinations of nearest neighbors is a challenging task. The knowledge of the binary systems is helpful to start.
In binary (ZnO)x(P2O5)1−x glasses with 0.33 ≤ x ≤ 0.5, the Zn−O coordination number (NZnO—the number of oxygens surrounding a Zn site) decreases from six to four with increasing zinc oxide content [25,26,27]. The minimum NZnO of four is reached at 50 mol% P2O5 [28]. In this compositional range, the NZnO value follows the ONB/Zn ratio = 2/x, indicating real Zn−O−P bridges [25]. The ONB/Zn ratio of the binary system is given on the left axis of Figure 1b. For glasses with x > 0.5, increasing fractions of ONB must share two Zn neighbors. The P−ONB bonds of the Qn with n = 3, 2, 1, and 0 show decreasing bond valences with 2.0, 1.5, 1.33, and 1.25 vu (valence unit) [23]. Thus, the electron charges on the ONB increase, and are to be balanced by the Zn-ions. The Q1 and Q0 groups supply a sufficient charge to their ONB to be shared by two Zn2+ that still form ZnO4. The bond valence is not strictly fixed for a given P−ONB in a Qn. The P−ONB bond distances show variations according to the demands of their local environments. For example, the bond valence P−ONB of the Q3 is a little less than 2.0 vu if this ONB coordinates a Zn [26]. For x > 0.5, the ONB atoms shared by two Zn cause some elongated Zn−O bonds if compared with those in Zn−O−P bridges. This effect broadens the peak of Zn−O distances. Also, the Zn−O coordination can exceed the number four in the binary glasses of the phosphate-rich compositions (x < 0.5) [29]. A broadening of the Zn−O peaks is expected, as well, because of longer bonds associated with the ZnO5 or ZnO6 that account for the increasing NZnO > 4 [25,26,27].
This work will mainly focus on the variations in the Zn−O distance peaks; the P and B environments and the borophosphate networks were characterized in a preceding work [13]. In addition to the glasses along a line from Zn(PO3)2 to BPO4 with fixed 50 mol% P2O5, two series with fixed O/P ratios are chosen for the diffraction study. An earlier measured sample of a binary Zn borate glass is adopted for discussing the change in the ZnOm units for a glass being free of P2O5. Two additional samples contain Na2O instead of ZnO to investigate the different effects of the Na+ and Zn2+ cations on details of the P−O bond lengths.

2. Materials and Methods

2.1. Samples

The zinc borophosphate glasses were prepared as described in a preceding work (Freudenberger and Brow [13]). The raw materials ZnO (99.5+%, Acros Organics, Waltham, MA USA), H3BO3 (cer. ACS, Fisher Scientific, Rochester, NY USA), and H3PO4 (85%, Fisher) were mixed to obtain glasses with the nominal compositions as given in Figure 1b and Table S1 (Supplementary Material). The samples are labeled zbp01 and so on according to their numbers given in Figure 1b. Two Na borophosphate glasses with labels nbp08 and nbp13 were prepared using Na2CO3 (99.5%, Alfa Aesar, Ward Hill, MA USA) as a raw material. The mixtures were calcined in platinum crucibles for 16 h at 350 °C. The duration of melting was two hours at temperatures in the range of 1000–1225 °C. The melts were quenched in preheated carbon molds. The clear glasses were annealed for one hour at temperatures of 400–550 °C (each at 10 °C below the respective glass transition temperature) [13]. The chemical analyses of the Zn borophosphate glasses of the preceding study [13] gave a loss of P2O5 content from 0 to 6 mol% with uncertainties of ~2.5 mol%. It was decided to start the data analyses with the batch compositions. The mass densities were measured using the Archimedes method. Two samples (zbp04, zbp11) exceed the range of homogeneous glasses that was reported in [22].

2.2. X-ray Scattering

The X-ray scattering (XRD) experiments were performed at the synchrotron (storage ring PETRA III at DESY Photon Science Hamburg/Germany). The hard photons at the beamline P02.1 have fixed radiation energies of ~60 keV. The exact wavelength of the incident beam (0.02080 nm) and the sample–detector distance (249.9 mm) were calibrated with the diffraction pattern from CeO2 powder. The beam size was 0.5 × 0.5 mm2. Thus, the beam is narrower than the silica capillaries (diameter 1 mm and wall thickness ~10 µm) that contain the powdered sample material. The beam stop between the sample and detector was fixed on a Kapton foil. The image-plate detector (Perkin Elmer 1621; Perkin Elmer Optoelectronics, Wiesbaden, Germany) used in these experiments is sensitive to photons of radiation energies > 20 keV. The samples were irradiated five times for one second with a full duration of the measurement equal to one minute per sample. The two-dimensional scattering images were merged to functions of the scattering angle 2θ. Further details on the instrument and first data analysis are described elsewhere [30]. The scattering intensities were corrected for container scattering, background, and absorption. The final intensities Icorr(Q) were normalized to the structure-independent scattering, which makes use of the chemical compositions of the samples and the tabulated atomic data of coherent and Compton scattering Icompt(Q) [31,32]. The final X-ray structure factors SX(Q) are given by the following:
S X ( Q ) = I c o r r Q · N < f 2 Q >   I c o m p t ( Q ) / < f Q > 2
where <…> means the average of the sample composition and f(Q) is the coherent atomic scattering amplitude. The normalization factor N allows SX(Q) to oscillate around unity. Smoothly changing corrections improve the behavior in the upper end of the Q-range. The scattering data of the zbp14 sample were obtained in an earlier laboratory experiment using Ag Kα radiation. A recent work [19] improved the knowledge of the range of binary Zn borate glasses from 54 to 70 mol% ZnO. Glasses with a smaller ZnO content show the separation of a B2O3-rich phase during melting. The Raman spectrum of the clear part of the glass with a 50 mol% batch is identical to that of the 54 mol% glass [19]. Hence, 54 mol% ZnO is also used for analyzing the data of the zbp14 sample instead of its batch composition of 50 mol%.

2.3. Neutron Scattering

The neutron scattering experiments were performed using the GEM instrument of the neutron spallation source ISIS of the Rutherford Appleton Laboratory (Chilton/UK). Measurements were made for six of the fifteen samples (zbp05, zbp07, zbp09, zbp11, zbp13, nbp13). The powdered sample material was loaded into thin-walled vanadium cylinders (diameter: 10.33 mm for all samples). The cylinder wall is a vanadium foil which is only 0.025 mm thick. The duration of data acquisition was at least 5 h per sample. A vanadium rod was used to determine the incident energy spectrum that is needed for the data normalization in the time-of-flight regime. The data were corrected using standard procedures for container and background scattering, attenuation, multiple scattering, and inelasticity effects [33]. Since natural boron with 20% of the strongly absorbing isotope 10B was used, a strong wavelength-dependent neutron absorption occurs. The corrections proved to be reliable even for up to 8% of boron atoms in the zbp11 sample. The differential scattering cross-sections of the detector groups 2, 3, 4, and 5 (scattering angles 14°–109°) are normalized to the mean scattering that is calculated from the sample compositions and tabulated neutron scattering lengths. Finally, the neutron structure factors SN(Q) were obtained in analogy to Equation (1).

3. Results

The corrected and normalized X-ray structure factors of the borophosphate glasses are shown in Figure S1. The SX(Q) data of the samples zbp04 and zbp11, which are close to the border of glass formation [22], show small Bragg reflections of tiny crystalline fractions. Statistical noise is no issue in the measurements using image plates. The neutron scattering experiments of six samples cover large Q-ranges, which bears the advantage of an excellent real-space resolution. The neutron diffraction results are shown in Figure 2 by the interference functions, which are weighted by the factor Q. The experimental data are compared with model curves composed of damped sinusoidal functions. These oscillating functions in the scattering intensities are the pendant to Gaussian peaks in the real-space correlation functions T(r). The model curves are calculated by the parameters given in Table S2 as obtained from the subsequently described Gaussian fitting. The T(r) functions are Fourier transforms of the experimental S(Q) data with the following:
T r = 4 π r ρ 0 + 2 π   0 Q m a x Q S Q 1 M Q sin Q r d Q     .
The number densities of atoms, ρ0, are calculated from the mass densities and glass compositions given in Table S1. The T(r) functions are calculated with Qmax = 183 nm−1 (X-rays), 400 nm−1 (neutrons), and M(Q) = 1. According to Lorch, a damping M(Q) with Qmax = 200 nm−1 (X-rays) and 500 nm−1 (neutrons) was used for a second series of T(r) functions. The Gaussian fitting procedures of the first-neighbor peaks were applied to both series of T(r) functions, where the effects of truncation at Qmax, M(Q) damping, and the Q-dependence of the partial weighting factors in the case of X-rays are taken into account [34,35]. Good fits for the first-neighbor peaks were obtained for all X-ray T(r) functions, whereas the fits of the Zn−O (Na−O) peaks in the neutron T(r) functions were not as good for some cases. Several comparisons of the model T(r) functions with the experimental T(r) functions are exemplified in Figure 3, Figure 4 and Figure S2–S4. The resulting peak parameters are listed in Table S2.
The combination of X-ray and neutron scattering does allow the separation of the B−O and P−O bonds due to the different changes in scattering contrast. Nevertheless, NBO = 4 was fixed in this analysis, which is consistent with previous 11B MAS NMR studies that showed that BO4 units were dominant in the structures of the full compositional range of the investigated borophosphate glasses [8,13,21,22]. In scattering experiments, the limit of detection for the presence of any BO3 fractions is of poorer certainty (±10%). The bond length of 0.147 nm for the BO4 is taken from related crystals [14,15,16,17] and is used together with a reasonable peak width to fit the scattering data.
For the binary ZnO-P2O5 glasses (zbp01, zbp05), the P−O peak is based on two bond lengths, P−ONB and P−OB, which differ by ~0.010 nm [26,36]. These fractions are directly determined from the P−O peaks in the T(r) of zbp05 from neutron diffraction (Figure 3). The bond fractions are also calculated from the O/P ratio that corresponds to the average n of the Qn distributions [23]. Here, the corresponding relation is also applied to the ternary glasses, assuming that the P−O bonds in P−O−B bridges are only insignificantly longer than those in P−O−Zn bridges. The necessary changes in the bond fractions to achieve excellent fits are small for the P−O peaks of all samples. However, the bond fractions calculated from O/P ratios are not correct for glasses with B−O−B bridges. Additional P−O−P bridges must also form [13]. The observed P−O bonds of the Na borophosphate glasses show slightly different behavior. Both deviations are discussed in Section 4.3.
Finally, the two P−O distances give a satisfactory approximation of the P−O peaks for all samples. Two samples show total NPO values of 3.6, which differs from the expected 4.0 by more than the common uncertainty. That can be due to a slightly changed composition for the zbp11 sample because its NZnO appears quite large. The deviation for zbp06 must be due to a normalization problem. In the case of all other samples, the NPO values are quite reasonable, which gives support to the use of the batch compositions for these analyses. The O−O peaks at 0.25 nm belong to the edges of the BO4 and PO4 units. The numbers of the tetrahedral edges were calculated from the fractions of the BO4 and PO4 units according to the batch compositions (Table S1). The calculation of the NOO value of the Zn borate glass zbp14 takes into account the fractions of the BO4 and BO3 units according to NBO = 3.3, which equals the result from 11B MAS NMR [19]. The successful fits of the huge O−O peaks in the case of the neutron T(r) functions support the choice of this approach. The edge length of the BO4 tetrahedron was fixed to 0.241 nm as taken from a crystal [14], and that of the PO4, which was close to 0.252 nm, was adjusted a little for each case.
Good fits of the metal cation–oxygen distances are possible up to ~0.27 nm. That is the lower limit of the B−O, P−O, and O−O second-neighbors with unknown numbers and distances. The overlaps with the O−O distances of the BO4 or PO4 tetrahedral edges are no issue in the fits. The contrast change of X-ray and neutron scattering differs for the O−O and Zn−O (Na−O) correlations. In addition, the calculated O−O peaks gave reasonable fits.
The scattering weight of the Zn−O partial correlation is larger for X-rays than for neutrons. Therefore, T(r) fitting in the range of the Zn−O distances was made to the X-ray correlation functions independent of the neutron data. The high-energy XRD has several advantages, including low absorption, and thus little demand for precise sample positioning. On the other hand, the real-space resolution of the neutron data is so good that termination effects are nearly vanishing. Oscillations caused by termination at Qmax in Equation (2) are visible only for the narrow B−O and P−O peaks, but not for the others (cf. Figure 3). The positions and shapes of the obtained Zn−O peaks are reproduced in the neutron data. The full peak heights are not reached for the zbp11 and zbp13 samples (Figure S3). The Zn−O partial of zbp11 has a small scattering weight for neutrons (only 5%) and large corrections for the highly absorbing boron content were needed. It is noteworthy that the fits to the T(r) functions obtained with Lorch damping gave equivalently good results. Figure S4 shows both variants of fits for the sodium borophosphate glass nbp13. The Na−O distances overlap with the O−O peaks. The X-ray data show a good fit for the broad Na−O peak with four oxygen neighbors at distances centered at 0.238 nm (Table S2). The neutron T(r) functions show unphysical features at ~0.185 nm, as indicated in Figure S4.
The present work aims to clarify how the glass composition affects the Zn−O environment. At first glance, the results are disappointing in that interesting changes are not observed. The Zn−O coordination numbers given in Figure S5 are almost constant at ~4.4, showing a slight increase toward smaller ZnO contents. The NZnO values of zbp06 and zbp11 are considered outliers. It is noteworthy that these are also the samples with small NPO values = 3.6. Commonly, the bond lengths increase with NZnO and the mean distances can be used as a criterion for the NZnO values. The ZnO4, ZnO5, and ZnO6 units have mean bond lengths of 0.194, 0.205, and 0.209 nm, respectively, as found in related crystal structures [37,38,39]. The Zn−O distances in the glasses have been reported with 0.194 nm for ZnO4 [28] and 0.209 nm for ZnO6 units [26]. In the present work, the Zn−O distances vary from 0.196 to 0.199 nm, with changes only a little outside the uncertainty interval (cf. Figure 5); these lengths were consistent with the obtained coordination numbers, which were found to be a little larger than four. A careful inspection of the bond lengths of the series with constant O/P = 3.25 (Figure 5b) identifies the shortest bonds for the zbp09 sample. This sample is also a member of the series of constant 50 mol% P2O5 (Figure 5a), where systematic distance changes were not observed. The Zn borate glass (zbp14) has an NZnO value of 4.3 and the largest detected Zn−O bond length of all samples at 0.200 nm (cf. Table S2). The Gaussian fit of this sample was repeated using the batch composition of 50 mol% ZnO instead of the 54 mol% taken from [19]. The two NZnO values do not differ, which is explained by a compensation of the compositional Zn decrease and O increase in the calculation of the Zn−O weighting factors. In the case of using 50 mol% ZnO, however, an NBO of only 2.9 was obtained.
The Zn−O peaks in the total X-ray T(r) functions (Figure 4) do not show significant changes. A further way to compare them is by considering the model peaks, which are here presented by the radial distribution functions. These Zn−O model peaks were calculated with the parameters obtained from Gaussian fitting and these peaks are compared with the bond lengths in the related crystals. This approach was recently used for binary phosphate glasses [27], where it was shown that the termination effects of the Fourier transformation (broadening and oscillations) and the influence of the changed scattering weights as the ZnO content changes are widely eliminated. The full widths at half maximum (fwhm) of these model peaks are shown in Figure 6. Again, a minimum is found for the zbp09 sample in the series in Figure 6b, whereas the values in Figure 6a do not show any clear trend. The Zn−O model peaks of all samples are shown for the three series of constant P2O5 content or constant O/P ratio (cf. Figure 7). For better comparability, the peaks of the glasses are given with identical heights. The peaks of the ZnO4, ZnO5, and ZnO6 units from the crystals show significant differences in the bond lengths. For all three series, the peaks of the glasses are close to that of the ZnO4 units known from the β-ZnP2O6 crystal [37]. The ZnO4 polyhedron of β-ZnP2O6 has a fifth oxygen nearby. Hence, the small contributions at ~0.230 nm found for all samples can be interpreted as real distances. Significant differences exist at ~0.215 nm. The peaks of the glasses with a constant 50 mol% P2O5 content shown in Figure 7a have small numbers of such distances and change very little. The samples outside the line of 50 mol% (cf. Figure 1) have significantly more Zn−O bonds at ~0.215 nm as becomes visible in Figure 7b,c. The differences appear very pronounced for the samples shown in Figure 7c. The zbp12, zbp13, and zbp14 samples have visibly more distances at ~0.215 nm than zbp04, due to bond lengths that belong to fractions of ZnO5 or ZnO6 units or shared ONB. This behavior is also found for zbp05, zbp06, zbp07, and zbp11 in Figure 7b. Among them, zbp11 is the only sample with a P2O5 content > 50 mol% (cf. Figure 1b). According to the large uncertainties of the measured coordination numbers, the behavior shown by the bond lengths is a better basis for discussing the Zn−O environments in the Zn borophosphate glasses. To support this statement, the fractions of the long bonds are calculated with Nlong/(Nshort + Nlong) and given in Figure S6. The Nshort and Nlong values are the numbers of oxygen neighbors at 0.194 and ~0.212 nm, as given in Table S2. Similar to the fwhm in Figure 6, the Nlong fractions for the series with constant 50 mol% P2O5 (Figure S6a) do not show systematic changes, whereas the minimum of long bonds in the series with the O/P ratio of 3.25 is found for the zbp09 sample (Figure S6b).

4. Discussion

4.1. Fully Tetrahedral Networks and the Ratio ONB/Zn

The Zn borophosphate glasses have the potential to form continuous disordered networks of corner-connected tetrahedra with all oxygen atoms being two-fold coordinated. The Zn metaphosphate glass (50 mol% P2O5) has been considered a continuous tetrahedral network for a long time now [40], which is manifested in its properties [41]. This classification is not entirely accurate because the Zn−O coordination number is slightly larger than four, which is also reflected in the corresponding bond lengths (Figure S4 and Figure 5). The ZnO4 units of the β-Zn(PO3)2 crystal [37] show a fifth Zn−O bond of 0.28 nm in length. Similarly, a few longer distances exist for the investigated glasses. Nevertheless, the metaphosphate glass zbp01 and the borophosphate glasses of 50 mol% P2O5 (zbp02, zbp09, zbp03, and zbp04) show the narrowest Zn−O peaks, together with the smallest distance contributions at ~0.212 nm (Figure 7a and Figure S6); this is evidence that a great majority of Zn ions are in tetrahedral units that have their four corners in Zn−O−P linkages.
The Zn−O peaks of the samples possessing less than 50 mol% P2O5 show increased contributions at ~0.212 nm that indicate ONB shared by two Zn. A significant increase in the corresponding NZnO is not detected. Two samples (zbp10, zbp11) of the series with the fixed O/P ratio of 3.25 have more than 50 mol% P2O5. The zbp11 sample shows significantly increased values of NZnO, rZnO, and peak width (fwhm). In this case, the ONB/Zn ratio is larger than four, and some larger units than ZnO4 are formed. Similar to the binary ZnO-P2O5 system [25,26,42], the borophosphate networks behave in such a way that avoids PO4 units (Q3, Q2) with terminal P=O double bonds. The possible second bond partner of this oxygen is Zn but not B, for reasons of bond valences. This is in contrast to what has been found for the Zn aluminophosphate glasses, where the Al3+ cations form AlO5 and AlO6 units for the P2O5-rich and ZnO-poor compositions [24]. This structural flexibility of Al allows the glass-forming range to reach the binary Al2O3-P2O5 border. However, similar to the BPO4, the range of the AlPO4 composition is excluded for its strong crystallization tendency.
On the side of the binary B2O3-P2O5 system, BPO4 crystals [14,15] are known, but disordered networks of PO4 and BO4 units are not formed. The structure of BPO4 consists of P 4 B 4 and B 4 P 4 units, each connected via their four corners by P−O−B bridges. Figure 8a shows the bond valences in the BPO4 crystals formed by these units. The tetrahedral network in the Zn(PO3)2 metaphosphate glass consists of P 0 B 2 and ZnO4 units (Figure 8b). If one connects the Zn(PO3)2 and the BPO4 compositions (50 mol% P2O5) in Figure 1, their mixture could create networks of ZnO4, BO4, and PO4 units whose corners are connected exclusively by twofold coordinated oxygens. That behavior receives support from the narrow Zn−O distance peaks (Figure 7a). However, the question arises as to how the structures made of Zn(PO3)2 and BPO4 with mixed P 0 B 2 and P 4 B 4 groups can be arranged, whereby the isolated ZnO4 and BO4 units must be provided with the necessary bond valences. Starting from the side of the ZnO-rich glasses, the problem is solved using a structural speciation reaction:
P 0 B 2 + P 4 B 4 2 P 2 B 3
where the oxygens of the P 2 B 3 (a Q1 unit) are shared with one ZnO4, one PO4, and two BO4 units, as shown in Figure 8c. The Q1 groups terminate the phosphate chains. Most ZnO4 corners interact with the remaining Q2 groups while the BO4 tetrahedra connect the Q1 and Q0 groups. Neither the ZnO4 nor the BO4 could exist in isolated sites alone with only Q1 neighbors (with P 0 B 1 or P 3 B 4 , respectively). The ZnO4 would have to share ONB and the BO4 would not receive sufficient bond valence. Approaching the composition of the zbp04 sample, nearly all Q2 have changed to Q1 groups. A high ordering of the ZnO4 and BO4 is required so that both groups can receive the necessary bond valences. Fortunately, a further solution exists for the bond valences, which stabilizes the disordered borophosphate networks, whereby some of the oxygens that, for example, would form P−O−B bonds, are instead incorporated into B−O−B bonds, necessitating the concomitant formation of new P−O−P bonds to produce a weak polymerization of the phosphate network [13]. This rearrangement is described by the following:
P 4 B 4 + B 4 P 4 P 3 B 4 + B 3 P 4 .
Figure 8d illustrates an arrangement of a P 3 B 4 with a B 3 P 4 unit. 11B MAS NMR has determined the fractions of the B 3 P 4 in addition to the B 4 P 4 units [13]. Hence, the O/P ratio is no longer exactly related to the Qn distribution. Four of the Zn borophosphate glasses used in [13] have ~50 mol% P2O5. The glass with the highest B2O3 fraction (18.5 mol%) was found with the highest B 3 P 4 fraction (88%). The other samples had ~35% B 3 P 4 in addition to the B 4 P 4 units. Reactions according to Equation (3) take place in glasses of high Q2 fractions (high ZnO content), whereas reactions according to Equation (4) dominate when the Q2 units are already in the minority (equal amounts of ZnO and B2O3). The fraction of oxygens that are available for coordinating the Zn2+ cations is preserved when the fraction of B 3 P 4 units changes. Equation (4) does not change the numbers of P−O and B−O bonds, as illustrated in Figure S7. That allows us to calculate the accurate ONB/Zn without knowing the accurate B 3 P 4 fraction and this ONB/Zn ratio largely determines the Zn−O environments.
Now, the question of how one can divide the oxygen fractions to coordinate the Zn and B is discussed. Simply knowing the sample composition of the ternary phosphate glasses is not sufficient for addressing this. The available number of oxygens is derived by subtracting the number of P−O−P bonds per PO4 and that is calculated from the total O/P ratio. In analogy to the structural behavior of the Zn phosphate glasses (cf. Introduction), all oxygens will find two neighbors with decreasing P2O5 content (P, B, Zn) before threefold coordinated oxygen sites occur. Sophisticated considerations have to take into account the different properties of the two sorts of cations besides the P5+ such as charge balance, field strength, preference of definite oxygen polyhedra, or threefold coordinated oxygen sites.
A general model (model 1) takes into account the different requirements for the charge compensation of the Zn2+ and B3+ cations. It is assumed that the two cationic species act independently of each other and are linked to the different Qn with equal probability, which makes sense for a simple approach. The corresponding ONB/Zn and OB/B values are obtained with the following:
O N B / Z n = 4 ( c O 2   c P ) / ( 2 c Z n + 3 c B )
and
O B / B = 6 ( c O 2   c P ) / ( 2 c Z n + 3 c B )
where the ci are the concentrations of the four sorts of atoms. This approach produces OB/B ratios that are too large (>4) for all samples, as shown in Figure 9, where the ratios are considered for the series with constant P2O5 contents at 50 mol% (a) and constant O/P ratios of 3.25 (b). The boron cannot meet this coordination behavior. In other words, the Zn−O environments and BO4 units cannot form independently of each other. Preferences for special Qn neighbors, as shown in Figure 8c,d, must be effective. In the range of the glasses studied, the situation is quite simple because 11B MAS NMR detected only BO4 (no BO3) [13]. When NBO is fixed to the number four for BO4 units, a well-defined fraction of oxygens is used as bridging oxygens by the borons. Then, the oxygens formed as non-bridging in the PO4 units are used for the coordination of the zinc (model 2) with the following ratio:
O N B / Z n = 2 c O 2   c P N B O   c B / c Z n .
For glasses with P2O5 contents of 50 mol%, model 2 predicts that ONB/Zn = 4. This value is exactly what is necessary for isolated ZnO4 units and fully tetrahedral networks (filled squares in Figure 9a). Differently, the fixed O/P ratios of 3.25 produce a continuous increase in ONB/Zn with increasing B2O3 content (cf. Figure 9b). For completeness, the ONB/Zn ratios according to model 2 for the zbp12 and zbp13 samples result in 3.10 and 3.27, respectively.
The ONB/Zn ratio becomes less than four for glasses with P2O5 contents below 50 mol%. Then, two ZnO4 units must share some of their ONB neighbors. The obtained ONB/Zn ratios are still larger than three, which means a few connected ZnO4 units are present, but not any interconnected ZnO4 substructures. If one looks at the binary ZnO-B2O3 glasses obtained with 54 to 70 mol% ZnO [19], theONB/Zn ratios reach only 1.5 to 1.9 (Table S3). For the ZnO4 units existing in these glasses, each ONB is shared by at least two Zn and extended substructures of interconnected ZnO4 exist. Octahedral ZnO6 units were proposed for these glasses [19]. However, EXAFS and X-ray scattering suggest that ZnO4 units are the dominant moiety [43]. The NZnO value of 4.3 for our Zn borate glass is similar to those of the borophosphate glasses. Its mean bond length of 0.200 nm is the largest among those of the other glasses (cf. Figure 7c). The Zn−O distances of zbp14 show large variations due to the massive need for shared ONB. Highly distorted ZnO5 polyhedra co-exist beside ZnO4.
Figure 9b shows strongly increasing ONB/Zn ratios > 4 with increasing B2O3 contents. Thus, a significant increase in NZnO is forced. According to the Zn−O distances (Figure 5b and Figure 7b), the NZnO of the zbp11 sample is significantly larger than four but it is expected to be less than five. Otherwise, in the case of NZnO < ONB/Zn, some terminal P=O double bonds must occur. According to the other limitations of the zbp11 sample (NPO = 3.6), its P2O5 content should be a little less than the nominal value. The formation of homogeneous glasses reaches the limit for this sample as reported for such compositions [22].
Model 1 was introduced assuming equal preferences of B and Zn for the oxygens of the different Qn. The B and Zn would form their environments independently. However, that is not possible with the limit NBO ≤ 4. For the ZnO-Al2O3-P2O5 glasses, the Al−O coordination number can increase up to six and the changes of NAlO according to model 1 are possible in a large concentration range. The NAlO values were calculated from the fractions of the AlO4, AlO5, and AlO6 units, which were determined with 27Al MAS NMR [24]. One of the glasses is a compositional analog to that of the zbp11 sample. That Al−O coordination number was 5.2 and close to 5.0, which resulted from model 1 (Figure 9b). The structural analysis of Na2O-Al2O3-P2O5 glasses has shown that all oxygen is used to break P−O−P bridges [44]. The compositional dependence of NAlO was found to follow other rules than simply a continuous increase with the ONB fractions. Abrupt changes from NAlO = 4 to 6 were found [44], which indicates the greater stability of the AlO4 and AlO6 polyhedra if compared with the AlO5. The Na+ coordinating the oxygens in Al−O−P bridges is essential for this behavior. Certainly, special preferences with the different Qn groups exist.
In this work, the different preferences for the oxygens of the Q2, Q1, and Q0 units with the Zn and B polyhedra are due to the restrictions of the B to BO4 units (model 2). The different field strengths of Zn2+ and B3+ are not essential in this context. The maximum limit NBO = 4 made the model considerations comparably simple for the borophosphate glasses of this study. As long as the NBO value = 4 is less than the OB/B ratio of model 1, the ONB/Zn ratio can be calculated and it determines the Zn−O environments. The success of model 2 became obvious in the predicted behavior of the Zn−O distances with the narrow peaks along the line of 50 mol% P2O5.

4.2. The Stabilization of the BO4 Units and the Boron Anomaly

BO4 tetrahedra and planar BO3 triangles are the units in the crystalline forms of B2O3 [45,46]. The BO4 unit is the variant with the densest packing and can be formed provided that its bond valences can be balanced. In the high-pressure B2O3 crystal [45], charge compensation is realized with oxygens which are threefold-linked to borons. Usually, glasses are only obtained when there is no or little threefold connected oxygen in the network. Accordingly, vitreous B2O3 forms BO3 units. Other mechanisms can deplete electron density from the B−O bonds in the BO4 unit resulting in four bonds with the necessary valences of ~0.75 vu. The BO4 units in binary borate glasses were identified as the origin of non-continuous property changes known as the boron anomaly [12].
For the borophosphate glasses, the BPO4 crystals [14,15] show the mutual benefit for the B and P atoms with a deficit and excess valence electron density in the B−O−P bridges of the tetrahedral networks (cf. Figure 8a). This type of valence transfer is effective throughout the Zn borophosphate glasses studied, whereby the BO4 units do not form B−O−Zn bridges. The typical PO4–Zn interactions are shown in Figure 10a,b, which were discussed above. The interaction of the Zn with the oxygen sites in P−O−B bridges is rather improbable. Each Zn−O bond of a ZnO4 requires a bond valence of ~0.50 vu, which is too much to be shared with these oxygens. For the same reasons, the oxygen corner of a ZnO4 unit cannot be the oxygen between two BO3 units or a pair of BO3 and BO4 (analogously to Figure 10g). These circumstances suggest a relation to the lack of glass formation in the B2O3-rich and P2O5-poor regions of the Zn borophosphate system. There is one exception; the oxygen sites in B−O−B between BO4 pairs can coordinate a Zn (cf. Figure 10d), as found in several Zn borate crystals [47,48,49,50]. However, this feature is not sufficient to stabilize any glasses poor in P2O5, but it can occur in the glasses of moderate P2O5 content until just under 50 mol%.
Binary Zn borate glasses are obtained in a small range rich in ZnO [19], where a significant fraction of ONB in BO3-triangles already exists (ONB/Zn ≥ 1.5). Two Zn2+ cations share each ONB (cf. Figure 10c), which means that at least three such ONB corners are present in each Zn-O environment. The scattering results of the Zn borate glass (zbp14) suggest a fraction of ZnO5 units in addition to the ZnO4. The ZnO5 have bonds of unequal lengths whose two more distant corners can coordinate the oxygens in any B−O−B bridge. The corner of a ZnO4 can only coordinate an OB in a bridge between two BO4 units (cf. Figure 10d), as explained in the preceding paragraph.
Similar to the ZnO4 units, the Na+ cations have total coordination numbers close to four (Table S2). The cation oxygen distance is larger for Na+ and the width (fwhm) of the Na−O peak is ~0.042 nm (cf. Figure S8), much more than the ~0.017 nm of the Zn−O peak (Figure 6). These parameters express the large flexibility of the Na+ cations to form distorted oxygen environments. Figure 10e,h show the interaction of the Na+ with the ONB of the P or B atoms, which is similar to the Zn−O bonds in Figure 10a–c. Here, the Na-O bonds are drawn with bond valences of ~0.25 vu as belonging to a NaO4 tetrahedron. This value can vary according to the distortions of the NaOm polyhedron, including the variations in m. The Na−O bonds are dominantly ionic, but it is better not to mix the bond valences and electronic charges in the schematic presentations. The weak Na−O interaction allows the Na+ to approach the OB sites in the P−O−B and B−O−B bridges (cf. Figure 10f,g). The corresponding oxygens are still quasi-twofold linked. The third partner Na+ forms a flexible bond that maintains sufficient flexibility to the disordered network as necessary for glass formation.
It was emphasized that the BO4 units in the Zn borophosphate glasses with P2O5 contents of ≥50 mol% are charge-balanced by PO4 units (Figure 8). The Zn2+ cations can approach oxygens in B−O−B bridges only for glasses with a P2O5 content < 50 mol% and then also contribute to the BO4 stabilization. Of course, the BO4 fractions in the binary Zn borate glasses [19] are charge-balanced by the Zn2+ cations. What about the Na borophosphate glasses? One can compare them with the known crystal structures. The Na2B8O13 crystal, which is free of phosphate [51], shows two mechanisms for BO4 stabilization through Na+. The lengths of the B−O bonds suggest that the surplus negative charge of the BO4 is not only balanced by the Na+ coordinating the BO4 corners, but it is also transferred to the neighboring BO3 triangles. The bond lengths in the BO3 units are strengthened in B−O−B bridges to the BO4 (~0.134 nm), whereas they are elongated in bridges to other BO3 (~0.138 nm). Thus, the Na+ cations also act across the BO3 triangles as neighbors of the BO4. It is to be presumed that there is no serious difference between the depletion effects of electron density from the bonds in the BO4 in the direction of the PO4 units or the Zn2+ and Na+ cations.

4.3. Different Effects of Zn and Na on the P−O Bond Lengths

The lengths of the P−O bonds in the zbp13 and nbp13 glasses with compositions with an O/P ratio of 3.5 reveal different effects of the Zn2+ and Na+ cations. Binary phosphate glasses with the same O/P ratio (pyrophosphates with ONB/OB = 6:1) possess two different lengths of P−O bonds, the P−ONB and P−OB bonds, with a frequency ratio of 3:1 [23,36]. The obtained distribution of bond lengths of a corresponding glass (ZnPbP2O7 [36]) is shown in Figure 11, and its structure is illustrated in Figure S9a.
The neutron diffraction results of zbp13 and nbp13 also show two types of P−O bonds (cf. Figures S3 and S4). The frequency numbers of the two P−O bonds are obtained with 2.8 and 1.0 for zbp13 and 2.5 and 1.3 for nbp13 (cf. Table S2). The corresponding ratios of bonds are similar to that of the ZnPbP2O7 glass, though the P−ONB/P−OB ratios of 9:11 for zbp13 and nbp13 are formally quite different from 3:1 (cf. Figure S9). This difference is explained by short P−OB bonds in the P−O−B bridges if compared with those in the P−O−P bridges.
For the zbp13 sample, the ratio is close to 3:1, and the P−O peak agrees with that of the ZnPbP2O7 glass (cf. Figure 11). The nbp13 sample has a slightly larger fraction of P−O bonds with lengths of ~0.160 nm. An effect that could increase this P−O fraction is the formation of B−O−B bridges between two BO4 according to Equation (4), which is accompanied by additional P−O−P bridges. A borophosphate glass with a composition close to that of zbp13 has 27.7% boron in B 3 P 4 units [13]. According to the glass composition of zbp13 and excluding B−O−B bridges, the P−O bonds can be divided into 25% in P−O−P bridges, 30% in P−O−B bridges, and 45% in P−ONB. A total of 27.7% of the B in B 3 P 4 units means that there is an increase in the bonds in P−O−P bridges from 25% to 27%. Therefore, the corresponding increase in the number of longer P−OB bonds is insignificant (<0.1).
The Na borophosphate glass nbp13 is expected to be almost free of B−O−B bridges, as has been reported for a similar glass (sample x = 0.125 in [10]). The structures of the Na3BP2O8 [16] and Na5B2P3O13 [17] crystals help to understand the larger fraction of the longer P−O bonds, although the crystal’s compositions differ somewhat from that of nbp13 (cf. Figure 1). The Na borophosphate crystals have isolated PO4 units (Q0). From the point of view of binary phosphates [23], only a single P−O distance should occur. However, two distances become obvious (cf. Figure 11). Half of the oxygens of the PO4 units participate in P−O−B bridges and these oxygens have a Na+ cation in their close vicinity, as illustrated in Figure 10f. These Na+ stabilize the BO4 but also reduce the bond valence in the adjacent PO4 (elongation of the P−O bond). The other two PO4 corners coordinate only Na+ cations, as shown in Figure 10e, and the corresponding short P−O bonds carry the surplus bond valence taken over from the others. The difference in both P−O bond strengths in these crystals is a little smaller than that in the zbp13 and ZnPbP2O7 glasses, the latter with bond valences of 1.33 and 1.0 vu (cf. Figure 11). For the nbp13 glass, the changes in bond lengths appear less pronounced than in the crystals. Again, 25% of the P−O bonds are in P−O−P bridges, 30% in P−O−B bridges, and 45% in P−ONB. To achieve the right ratio of the short and long P−O bonds of the nbp13 sample (2.5:1.3), the fraction of the P−O−B bridges is split into two parts, those with a Na+ close nearby and those free of any Na+ neighbor. Then, 30% of the P−O bonds in the P−O−B bridges are elongated, indicating the effect of a Na+ cation. The other 70% of P−O bonds in P−O−B bridges are not affected by any Na+ cations and have lengths similar to those of typical P−ONB bonds. This difference from zbp13 means that the BO4 units in the nbp13 glass undergo charge compensation equally by Na+ cations and PO4 units. The ZnO4 units in zbp13 contribute less because their corners do not participate in P−O−B bridges.
Two species of P−O−B bridges are suggested for nbp13 that seem to be easily distinguishable. One is coordinated to a Na+ cation, and the other is not. On the other hand, a clear separation of the P−O−B bridges into those with and without a Na+ neighbor might be questionable due to the large width of the Na−O distance peak. Both P−O peak components overlay with each other. The analyses of the O 1s spectra (XPS) did not distinguish two different types of P−O−B bridges in the Na and Zn borophosphate glasses [8,20]. Nevertheless, the effect of the Na+ on the P−O−B is real and it is not effective for Zn2+. Another approach supports the fact that only part of the P−O−B bridges of nbp13 can have Na+ neighbors. The total Na−O coordination number is calculated assuming all oxygens in the P−ONB bonds or P−O−B bridges are present with three or one Na+ neighbors, as shown in Figure 10e,f, respectively. Then, for nbp13, a NNaO value of ~6.0 is obtained, which noticeably exceeds the value of 4.0 obtained from scattering. Accordingly, only part of the P−O−B can have a Na+ neighbor (30% as estimated for nbp13). Also, the ONB shown in Figure 10e can only have 2.3 Na+ neighbors on average instead of 3 Na+ neighbors. These conditions mean that only a third of the Na+ can coordinate the oxygen in a P−O−B bridge. All other oxygens in the Na−O bonds belong to the P−ONB bonds. Here, only the behavior of a single sample is discussed. The corresponding relations will strongly change with the glass compositions.
The ionic forces of Na+ are weak compared to strong P−O and B−O bonds. Therefore, Na+ does not form dense oxygen environments, e.g., NaO6 octahedra. The broad Na−O distance peaks (cf. Figure S8) correspond to NNaO values of 4.5 and 4.0 in the glasses and are thus significantly smaller than those (5 and 6) in the related crystals [16,17], accompanied by shorter bonds. Likewise, the number and bond valences of the P−ONB bonds in zbp13 (cf. Figure 10b) would allow the Zn2+ to form ZnO6 octahedra, which is not the case. The strong bonds in the PO4 and BO4 units and the comparatively high cross-linking density of the borophosphate networks prevent the modifier cations from forming compact oxygen polyhedra.
The relation between the network structure and the modifier environments was also used to interpret the mixed network-former effect. The average number of OB per glass-forming unit, BO, calculated for the yNa2O-xB2O3-(1−x)P2O5 glasses, characterizes the stiffness of the network [6,10]. This value’s behavior correlates well with the Tg values. A larger BO means an increasing network stiffness, which is accompanied by more disordered modifier environments; in this case, it is the NaOm polyhedron. In addition, the ONB/Na ratio decreases (Equation (7)). These changes explain the reduced activation energies for ionic transport when the B2O3 content increases while the Na2O content is constant [6,10]. The average number of bridging corners per network-forming unit, BCNFU, is a more plausible value with BCNFU = 2 BO. For the borophosphate glasses, it is calculated with the following:
B C N F U = 2 4 c P + N B O   c B c O / c P + c B     .
The BCNFU value is two for the P 0 B 2 chain units in the NaPO3 glass. It increases to 2.6 for the nbp13 sample. Maximum values of ≥3 are reached for a larger B2O3 content until the potential of the PO4 units to stabilize the BO4 is exhausted and BO3 triangles occur. The Na−O distances of the two glass compositions are shown in Figure S8. A significant change in the Na−O peaks is not observed. The small scattering power of Na, if compared with Zn and the overlap with the O−O distances at 0.25 nm, impede an accurate distance analysis.
Previous diffraction work on Na2O-B2O3-P2O5 glasses [52,53] did not reach such a large Qmax, which is needed to resolve different P−O bonds. The structural analysis was made using the Reverse Monte Carlo (RMC) technique, which aimed to elucidate the mixed network-former effect. A main point was the description of the medium-range order which includes the migration pathways for the Na+ cations. The analysis of the Na−O environments in the model configurations gave total NNaO values of ~4.5 with very broad distributions of the distinct Na sites ranging from a two- to eightfold coordination. The two distances of 0.152 and 0.156 nm given in [53] for the P−ONB and P−OB bonds are simply calculated from the RMC configurations, but are not resolved in the measuring results and thus cannot be compared with our data.

4.4. The Evolution of the Properties of the Zn Borophosphate Glasses

The figures with the mass densities ρ of the ZnO-B2O3-P2O5 glasses in [22] show a continuous increase with fixed ZnO and increasing B2O3 contents. Glass transition temperatures Tg show similar trends while the expansion coefficients α change in opposite directions. The values of ρ and Tg in [13] show similar trends, although the comparisons are made along constant O/P ratios. It needs to be remembered that the Zn phosphate glasses have been classified as showing anomalous behavior [40]. The minima of mass density at 50 mol% ZnO [25,40,54] or Tg values [13,54] at ~60 mol% ZnO are characteristic of the binary Zn phosphate glasses. Below, the mass densities that were published for the full range of glass formation in the ZnO-B2O3-P2O5 system are discussed [22]. Packing densities are better suited for comparisons whereby the influence of atomic size and mass is widely eliminated. Such comparisons were made for the ZnO-P2O5 glasses recently [27] by using the ionic radii from [55].
A packing density is the filled volume fraction, assuming the atoms are spheres of known ionic radii. Since the BO4 tetrahedra have small O−O distances, there is some overlap of the oxygen spheres with an ionic radius of 0.135 nm; to a lesser extent, there is also some overlap for the PO4 tetrahedra. A correction was made for this overlap. The packing densities of the Zn borophosphate glasses obtained from the mass densities given in [22] are shown in Figure 12 as numbers given close to the sample positions in the concentration triangle. The absolute minimum at 50 mol% ZnO shows a packing density of 46.4%. This behavior was attributed to being caused by the network of corner-connected PO4 and ZnO4 tetrahedra with all oxygens in bridges (Figure 8b), a network that fills the space quite inefficiently [25].
Above, it was shown that the tetrahedral character of the network does not change when B2O3 replaces the ZnO while 50 mol% P2O5 is fixed. A plateau of ~51% packing density is reached with 10% B2O3 (Figure 12). This increase is due to the exchange of the rather open ZnO4 unit for the more compact BO4 unit. There exists a further subtle increase in the packing density up to ~52.5% in the direction of decreasing P2O5 contents. This fact is interpreted with the change from P−O−Zn bridges (Figure 10a) to the first ONB shared by two Zn (Figure 10b) and the exchange of PO4 for BO4 units. Finally, when only isolated PO4 units exist, then BO3 units with ONB shared by two Zn are also formed (Figure 10c), and the packing densities still increase. The (ZnO)x(B2O3)1−x glasses with 0.54 ≤ x ≤ 0.67 [19] with few BO4 connected with BO3 have packing efficiencies of 55% to 53.5%, where the ONB of the BO3 are shared by two Zn. For the reachable glass compositions, the BO3 units of the Zn borate glasses have one or two ONB.

5. Conclusions

The atomic structures of Zn and Na borophosphate glasses were investigated using scattering methods. Neutron scattering was successfully performed on glasses with up to 8% boron atoms, which contains the strongly absorbing 10B isotope in natural abundance. The X-ray scattering of high-energy photons using the image plate technique was crucial for determining the Zn−O distances. Neutron scattering at the spallation source offered a large measuring range (large Qmax), which is a prerequisite for the high resolving power of pair distances. Detailed analyses of different lengths of P−O bonds became possible.
The Zn−O coordination numbers were found to be a little larger than four and did not change significantly when changing the glass compositions. What is striking, however, is that the glasses along a line connecting the BPO4 and Zn(PO3)2 compositions (50 mol% P2O5) showed the narrowest peaks of Zn−O distances. This indicates the formation of tetrahedral networks of PO4, BO4, and ZnO4 with mostly isolated ZnO4 units and no B−O−Zn bridges. The calculated amount of the ONB that is available for the Zn−O coordination supports this interpretation. For smaller P2O5 contents, the Zn share ONB which leads to distances of ~0.212 nm and thus, a broadening of the Zn−O peaks. The one sample of a P2O5 content larger than 50 mol% shows a subtle increase in NZnO that is accompanied by a broadened Zn−O peak as well. The two investigated Na borophosphate glasses have Na−O distances similar to those of a NaPO3 glass with coordination numbers of 4 to 4.5, while NaO5 and NaO6 polyhedra exist in the related crystals with Na−O distances that are a little larger. The first-neighbor distances of the B−O, P−O, and O−O pairs in the real-space correlations confirm that only BO4 units co-exist with the PO4 in the borophosphate networks.
Two peaks of P−O bonds are visible for the binary phosphate glasses, which has been commonly observed. They are assigned to the short P−ONB bonds and the longer P−OB bonds. This behavior also exists for the ZnO-B2O3-P2O5 glasses whereby, however, the first peak component includes the ordinary P−ONB bonds together with the bridging bonds of the P−O−B linkages, according to their similar lengths. Though some B−O−B bridges exist, the concomitant increase in the fraction of the longer P−OB bonds is small. On the other hand, significantly larger fractions of the long P−OB bonds are found for the Na2O-B2O3-P2O5 glasses. Some Na+ cations coordinate oxygens in a part of the P−O−B bridges which elongates the corresponding P−O bonds. Thus, BO4 units formed in the Zn borophosphate glasses are charge-balanced by mainly PO4 units while the charge balance of the BO4 in the Na borophosphate glasses is realized by the neighboring PO4 and the Na+ cations equally.
The property changes of the ZnO-B2O3-P2O5 glasses seem to be complex. The changes with the B2O3 additions interfere with the known anomalies of the ZnO-P2O5 glasses that have a minimum packing density at 50 mol% and a minimum glass transition temperature at ~40 mol% P2O5.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/solids5030024/s1, Figure S1: X-ray structure factors; Figure S2: X-ray correlation functions zbp14; Figure S3: Neutron and X-ray correlation functions zbp07, zbp11, and zbp13; Figure S4: Neutron and X-ray correlation functions nbp13; Figure S5: Zn−O coordination numbers; Figure S6: Fractions of the long Zn−O bonds; Figure S7: Rearrangements of network groups; Figure S8: Na−O first-neighbor distances; Figure S9: Structural illustrations; Table S1: Batch compositions and mass densities; Table S2: Parameters of the model Gaussian peaks; Table S3: Parameters of the Zn borate glasses [19].

Author Contributions

Conceptualization, U.H. and R.K.B.; methodology, P.T.F., J.B., U.H. and A.C.H.; formal analysis, U.H.; investigation, P.T.F. and U.H.; resources, R.K.B., J.B. and A.C.H.; writing—original draft preparation, U.H.; writing—review and editing, R.K.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data are available upon request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ray, N.H. Inorganic Polymers; Academic Press: London, UK, 1978; pp. 79–90. [Google Scholar]
  2. Sharmin, N.; Hasan, M.S.; Parsons, A.J.; Furniss, D.; Scotchford, C.A.; Ahmed, I.; Rudd, C.D. Effect of boron addition on the thermal, degradation, and cytocompatibility properties of phosphate-based glasses. BioMed Res. Int. 2013, 2013, 902427. [Google Scholar] [CrossRef] [PubMed]
  3. Petit, L.; Cardinal, T.; Videau, J.J.; Smektala, F.; Jouan, T.; Richardson, K.; Schulte, A. Fabrication and characterization of new Er3+ doped niobium borophosphate glass fiber. Mater. Sci. Eng. B 2005, 117, 283–286. [Google Scholar] [CrossRef]
  4. Chiodelli, G.; Magistris, A.; Villa, M. Ionic conductivity and glass transition of borophosphate glasses. Solid State Ion. 1986, 18–19, 356–361. [Google Scholar] [CrossRef]
  5. Christensen, G.; Olson, G.; Martin, S.W. Ionic conductivity of mixed glass former 0.35Na2O + 0.65[xB2O3 + (1 x)P2O5] glasses. J. Phys. Chem. B 2013, 117, 16577–16586. [Google Scholar] [CrossRef]
  6. Zielnok, D.; Cramer, C.; Eckert, H. Structure/property correlations in ion-conducting mixed-network former glass: Solid-state NMR studies of the system Na2O-B2O3-P2O5. Chem. Mater. 2007, 19, 3162–3170. [Google Scholar] [CrossRef]
  7. Carta, D.; Qiu, D.; Guerry, P.; Ahmed, I.; Abou Neel, E.A.; Knowles, J.C.; Smith, M.E.; Newport, R.J. The effect of composition on the structure of sodium borophosphate glasses. J. Non-Cryst. Solids 2008, 354, 3671–3677. [Google Scholar] [CrossRef]
  8. Raskar, D.; Rinke, M.T.; Eckert, H. The mixed-network former effect in phosphate glasses: NMR and XPS studies of the connectivity distribution in the glass system (NaPO3)1−x(B2O3)x. J. Phys. Chem. C 2008, 112, 12530–12539. [Google Scholar] [CrossRef]
  9. Muñoz, F.; Montagne, L.; Pascual, L.; Durán, A. Composition and structure dependence of the properties of lithium borophosphate glasses showing boron anomaly. J. Non-Cryst. Solids 2009, 355, 2571–2577. [Google Scholar] [CrossRef]
  10. Larink, D.; Eckert, H.; Reichert, M.; Martin, S.W. Mixed network former effect in ion-conducting alkali borophosphate glasses: Structure/property correlations in the system [M2O]1/3[(B2O3)x(P2O5)1−x]2/3 (M = Li, K, Cs). J. Phys. Chem. C 2012, 116, 26162–26176. [Google Scholar] [CrossRef]
  11. Christensen, R.; Olson, G.; Martin, S.W. Structural studies of mixed glass former 0.35Na2O + 0.65[xB2O3 + (1 − x)P2O5] glasses by Raman and 11B and 31P Magic Angle Spinning Nuclear Magnetic Resonance spectroscopies. J. Phys. Chem. B 2013, 117, 2169–2179. [Google Scholar] [CrossRef]
  12. Krogh-Moe, J. On the Structure of Boron Oxide and Alkali Borate Glasses; Doktorsavhandlingar vid Chalmers Tekniska Högskola Nr. 22: Göteborg, Sweden, 1959. [Google Scholar]
  13. Freudenberger, P.T.; Brow, R.K. Spectroscopic and chromatographic analyses of zinc borophosphate glasses. Phys. Chem. Glas. B 2018, 59, 227–284. [Google Scholar]
  14. Haines, J.; Cambon, O.; Astier, R.; Fertey, P.; Chateau, C. Crystal structures of α-quartz homeotypes boron phosphate and boron arsenate: Structure-property relationships. Z. Krist. 2004, 219, 32–37. [Google Scholar] [CrossRef]
  15. Haines, J.; Chateau, C.; Léger, J.M.; Bogicevic, C.; Hull, S.; Klug, D.D.; Tse, J.S. Collapsing cristobalitelike structures in silica analogues at high pressure. Phys. Rev. Lett. 2003, 91, 015503. [Google Scholar] [CrossRef]
  16. Bang, X.D.; Hong, C.H.; Xin, Y.X.; Thai, Z.J. Low-temperature flux syntheses and characterizations of two 1-D anhydrous borophosphates: Na3B6PO13 and Na3BP2O8. J. Solid State Chem. 2007, 180, 233–239. [Google Scholar]
  17. Hauf, C.; Friedrich, T.; Kniep, R. Crystal structure of pentasodium catena-(diborato-triphosphate), Na5B2P3O13. Z. Kristallogr. 1995, 210, 446. [Google Scholar] [CrossRef]
  18. Zhang, E.; Zhao, S.; Zhang, J.; Fu, P.; Yao, J. The β-modification of trizinc borate phosphate, Zn3(BO3)(PO4). Acta Crystallogr. E 2001, 67, i3. [Google Scholar] [CrossRef] [PubMed]
  19. Topper, B.; Möncke, D.; Youngman, R.E.; Valvi, C.; Kamitsos, E.I.; Versamis, C.P.E. Zinc borate glasses: Properties, structure and modelling of the composition-dependence of borate speciation. Phys. Chem. Chem. Phys. 2023, 25, 5967–5988. [Google Scholar] [CrossRef]
  20. Brow, R.K. An XPS study of oxygen bonding in zinc phosphate and zinc borophosphate glasses. J. Non-Cryst. Solids 1996, 194, 267–273. [Google Scholar] [CrossRef]
  21. Brow, R.K.; Tallant, D.R. Structural design of sealing glasses. J. Non-Cryst. Solids 1997, 222, 396–406. [Google Scholar] [CrossRef]
  22. Koudelka, L.; Mošner, P. Borophosphate glasses of the ZnO−B2O3−P2O5 system. Mater. Lett. 2000, 42, 194–199. [Google Scholar] [CrossRef]
  23. Brow, R.K. Review: The structure of simple phosphate glasses. J. Non-Cryst. Solids 2000, 263–264, 1–28. [Google Scholar] [CrossRef]
  24. Smith, C.E.; Brow, R.K.; Montagne, L.; Revel, B. The structure and properties of zinc aluminophosphate glasses. J. Non-Cryst. Solids 2014, 386, 105–114. [Google Scholar] [CrossRef]
  25. Hoppe, U.; Walter, G.; Kranold, R.; Stachel, D.; Barz, A. The dependence of structural peculiarities in binary phosphate glasses on their network modifier content. J. Non-Cryst. Solids 1995, 192–193, 28–31. [Google Scholar] [CrossRef]
  26. Hoppe, U.; Walter, G.; Stachel, D.; Barz, A.; Hannon, A.C. Neutron and X-ray diffraction study on the structure of ultraphosphate glasses. Z. Naturforsch. A 1997, 52, 259–269. [Google Scholar] [CrossRef]
  27. Hoppe, U. The oxygen environments of divalent cations in phosphate glasses and crystals—The change with composition and ionic radii. J. Non-Cryst. Solids 2023, 619, 122599. [Google Scholar] [CrossRef]
  28. Bionducci, M.; Licheri, G.; Musinu, A.; Navarra, G.; Piccaluga, G.; Pinna, G. The structure of Zn(II) metaphosphate glass. Z. Naturforsch. A 1996, 51, 1209–1215. [Google Scholar] [CrossRef]
  29. Xia, Y.; Chen, H.; Hung, I.; Gan, Z.; Sen, S. Structure and fragility of Zn-phosphate glasses: Results from multinuclear NMR spectroscopy and calorimetry. J. Non-Cryst. Solids 2022, 580, 121395. [Google Scholar] [CrossRef]
  30. Dippel, A.-C.; Liermann, H.-P.; Delitz, J.T.; Walter, P.; Schulte-Schrepping, H.; Seeck, O.H.; Franz, H. Beamline P02.1 at PETRA III for high-resolution and high-energy powder diffraction. J. Synchrotron Radiat. 2015, 22, 675–687. [Google Scholar] [CrossRef]
  31. Waasmaier, D.; Kirfel, A. New analytical scattering-factor functions for free atoms and ions. Acta Crystallogr. A 1995, 51, 416–431. [Google Scholar] [CrossRef]
  32. Hubbell, J.H.; Veigele, W.J.; Briggs, E.A.; Brown, R.T.; Cromer, D.T.; Howerton, R.J. Atomic form factors, incoherent scattering functions, and photon scattering cross sections. J. Phys. Chem. Ref. Data 1975, 4, 471–538. [Google Scholar] [CrossRef]
  33. Hannon, A.C. Results on disordered materials from the GEneral Materials diffractometer, GEM, at ISIS. Nucl. Instrum. Methods Phys. Res. Sect. A 2005, 551, 88–107. [Google Scholar] [CrossRef]
  34. Leadbetter, A.J.; Wright, A.C. Diffraction studies of glass structure: I. Theory and quasi-crystalline model. J. Non-Cryst. Solids 1972, 7, 23–26. [Google Scholar] [CrossRef]
  35. Hannon, A.C. Neutron diffraction techniques for structural studies of glasses. In Modern Glass Characterization; Affatigato, M., Ed.; John Wiley & Sons: Hoboken, NJ, USA, 2015; p. 190ff. [Google Scholar]
  36. Hoppe, U.; Kranold, R.; Stachel, D.; Barz, A.; Hannon, A.C. Variation in P-O bonding in phosphate glasses—A neutron diffraction study. Z. Naturforsch. A 2000, 55, 369–380. [Google Scholar] [CrossRef]
  37. Weil, M. The high-temperature modification of zinc catena-polyphosphate, β-Zn(PO3)2. Acta Crystallogr. C 2004, 60, i20–i22. [Google Scholar] [CrossRef] [PubMed]
  38. Robertson, B.E.; Calvo, C. Crystal structure of α-Zn2P2O7. J. Solid State Chem. 1970, 1, 120–133. [Google Scholar] [CrossRef]
  39. Bataille, T.; Benard-Rocherulle, P.; Louer, D. Thermal behavior of zinc phenylphosphonate and structure determination of g-Zn2P2O7 from X-ray powder diffraction data. J. Solid State Chem. 1998, 140, 62–70. [Google Scholar] [CrossRef]
  40. Kordes, E.; Vogel, W.; Feterowsky, R. Physikalisch-chemische Untersuchungen über die Eigenschaften und den Feinbau von Phosphatgläsern. Z. Elektrochem. 1953, 57, 282–289. [Google Scholar] [CrossRef]
  41. Krause, J.T.; Kurkjian, C.R. Vibrational anomalies in inorganic glass formers. J. Am. Ceram. Soc. 1968, 51, 226–227. [Google Scholar] [CrossRef]
  42. Hoppe, U. A structural model for phosphate glasses. J. Non-Cryst. Solids 1995, 195, 138–147. [Google Scholar] [CrossRef]
  43. Kajinami, A.; Harada, Y.; Inoue, S.; Deki, S.; Umesaki, N. The structural analysis of zinc borate glass by laboratory EXAFS and X-ray diffraction measurements. Jpn. J. Appl. Phys. 1999, 38 (Suppl. 1), 132–135. [Google Scholar] [CrossRef]
  44. Brow, R.K.; Kirkpatrick, R.J.; Turner, G.L. Nature of alumina in phosphate glasses: II, structure of sodium aluminophosphate glass. J. Am. Ceram. Soc. 1993, 76, 919–928. [Google Scholar] [CrossRef]
  45. Prewitt, C.T.; Shannon, R.D. Crystal structure of a high-pressure form of B2O3. Acta Crystallogr. B 1968, 24, 869–874. [Google Scholar] [CrossRef]
  46. Gurr, G.E.; Montgomery, P.W.; Knutson, C.D.; Gorres, B.T. The crystal structure of trigonal diboron trioxide. Acta Crystallogr. B 1970, 26, 906–915. [Google Scholar] [CrossRef]
  47. Martinez-Ripoli, M.; Martinez-Carrera, S.; Garcia-Blanco, S. he crystal structure of zinc diborate, ZnB4O7. Acta Crystallogr. B 1971, 27, 672–677. [Google Scholar] [CrossRef]
  48. Bondareva, O.S.; Egorov-Tismenko, Y.K.; Simonov, M.A.; Belov, N.V. Refined crystal structure of cubic skeletov zinc borate Zn4O[B6O12]. Dokl. Akad. Nauk. SSSR 1978, 241, 815–817. [Google Scholar]
  49. Ziegler, R.; Purtscher, F.R.S.; Bayarjargal, L.; Hofer, T.S.; Huppertz, H. The new non-centrosymmetric zinc borate Zn2B10O17 with a unique crystal structure. Z. Anorg. Allg. Chem. 2023, 649, e202300093. [Google Scholar] [CrossRef]
  50. Ziegler, R.; Seibald, M.; Stoll, C.; Huppertz, H. Synthesis and crystal structure of the zinc borate Zn3B4O9. Eur. J. Inorg. Chem. 2023, 26, e2023000174. [Google Scholar] [CrossRef]
  51. Hyman, A.; Perloff, A.; Mauer, F.; Block, S. The crystal structure of sodium tetraborate. Acta Crystallogr. 1967, 22, 815–821. [Google Scholar] [CrossRef]
  52. Le Roux, S.; Martin, S.; Christensen, R.; Ren, Y.; Petkov, V. Three-dimensional structure of multicomponent (Na2O)0.35[(P2O5)1−x(B2O3)x]0.65 glasses by high-energy X-ray diffraction and constrained reverse Monte Carlo simulations. J. Phys. Condens. Matter 2011, 23, 035403. [Google Scholar] [CrossRef]
  53. Karlsson, M.; Schuch, M.; Christensen, R.; Maass, P.; Martin, S.W.; Imberti, S.; Matic, A. Structural origin of the mixed glass former effect in sodium borophosphate glasses investigated with neutron diffraction and reverse Monte Carlo modeling. J. Phys. Chem. C 2015, 119, 27275–27283. [Google Scholar] [CrossRef]
  54. Brow, R.K.; Tallant, D.R.; Myers, S.T.; Phifer, C.C. The short-range structure of zinc polyphosphate glass. J. Non-Cryst. Solids 1995, 191, 45–55. [Google Scholar] [CrossRef]
  55. Shannon, R.D.; Prewitt, C.T. Effective ionic radii in oxides and fluorides. Acta Crystallogr. B 1969, 25, 925–946. [Google Scholar] [CrossRef]
Figure 1. Compositions of the ternary zinc and alkali borophosphate glasses as reported in the literature (a) and measured in this work (b). The upper three citations [13,20,21,22] in (a) belong to ZnO-B2O3-P2O5 glasses, and the others [6,7,10,11] to alkali (A2O) borophosphate glasses. The red dots with numbers in (b) mark the Zn borophosphate glasses of the present work. In addition to the ZnO and P2O5 contents, the ONB/Zn and OB/B ratios for the binary samples are given. Borophosphate glasses of compositions 8 and 13 with Na2O instead of ZnO were also measured.
Figure 1. Compositions of the ternary zinc and alkali borophosphate glasses as reported in the literature (a) and measured in this work (b). The upper three citations [13,20,21,22] in (a) belong to ZnO-B2O3-P2O5 glasses, and the others [6,7,10,11] to alkali (A2O) borophosphate glasses. The red dots with numbers in (b) mark the Zn borophosphate glasses of the present work. In addition to the ZnO and P2O5 contents, the ONB/Zn and OB/B ratios for the binary samples are given. Borophosphate glasses of compositions 8 and 13 with Na2O instead of ZnO were also measured.
Solids 05 00024 g001
Figure 2. Weighted neutron interference functions, Q·[S(Q)−1]: the experimental data (black dots) are compared with model functions (orange solid lines) that are calculated by the parameters given in Table S2 for the model first-neighbor peaks. Except for zbp13, the baselines of the data are shifted up or down for visual clarity.
Figure 2. Weighted neutron interference functions, Q·[S(Q)−1]: the experimental data (black dots) are compared with model functions (orange solid lines) that are calculated by the parameters given in Table S2 for the model first-neighbor peaks. Except for zbp13, the baselines of the data are shifted up or down for visual clarity.
Solids 05 00024 g002
Figure 3. Neutron and X-ray correlation functions T(r) of the zbp05 and zbp09 samples as obtained by Equation (2) without damping (M(Q) = 1). The baselines of the neutron data are shifted up for visual clarity. The experimental curves (black dots) are compared with the total model functions (thick red lines) and the partial Zn−O peaks (thick olive lines). The other partial model peaks are given with thin lines (B−O—magenta; P−O—blue; O−O—orange and purple). The two model O−O peaks correspond to different edge lengths of the BO4 and PO4 tetrahedra.
Figure 3. Neutron and X-ray correlation functions T(r) of the zbp05 and zbp09 samples as obtained by Equation (2) without damping (M(Q) = 1). The baselines of the neutron data are shifted up for visual clarity. The experimental curves (black dots) are compared with the total model functions (thick red lines) and the partial Zn−O peaks (thick olive lines). The other partial model peaks are given with thin lines (B−O—magenta; P−O—blue; O−O—orange and purple). The two model O−O peaks correspond to different edge lengths of the BO4 and PO4 tetrahedra.
Solids 05 00024 g003
Figure 4. X-ray correlation functions T(r) as obtained without damping (M(Q) = 1) for the series with fixed P2O5 content (a) and fixed O/P ratio (b). Except for zbp01 and zbp05 samples, the baselines of the data are shifted up for visual clarity. The experimental curves (black dots) are compared with the total model functions (thick red lines) and the partial Zn−O peaks (thick olive lines). The other partial model peaks are given with thin lines (B−O—magenta; P−O—blue; O−O—orange and purple). The two model O−O peaks correspond to different edge lengths of the BO4 and PO4 units.
Figure 4. X-ray correlation functions T(r) as obtained without damping (M(Q) = 1) for the series with fixed P2O5 content (a) and fixed O/P ratio (b). Except for zbp01 and zbp05 samples, the baselines of the data are shifted up for visual clarity. The experimental curves (black dots) are compared with the total model functions (thick red lines) and the partial Zn−O peaks (thick olive lines). The other partial model peaks are given with thin lines (B−O—magenta; P−O—blue; O−O—orange and purple). The two model O−O peaks correspond to different edge lengths of the BO4 and PO4 units.
Solids 05 00024 g004
Figure 5. The Zn−O first-neighbor distances for two series of the Zn borophosphate glasses: (a) constant P2O5 content, (b) constant O/P ratio. The black squares denote the values corresponding to typical bond lengths. The red circles are values that include the contributions being a little longer with ~0.23 nm. To assign sample numbers, see Figure 1b. The equal values at 12.5 mol% B2O3 in (a) and (b) belong to sample zbp09.
Figure 5. The Zn−O first-neighbor distances for two series of the Zn borophosphate glasses: (a) constant P2O5 content, (b) constant O/P ratio. The black squares denote the values corresponding to typical bond lengths. The red circles are values that include the contributions being a little longer with ~0.23 nm. To assign sample numbers, see Figure 1b. The equal values at 12.5 mol% B2O3 in (a) and (b) belong to sample zbp09.
Solids 05 00024 g005
Figure 6. Full widths at half maximum (fwhm) values of the model Zn−O peaks as shown in Figure 7 for two series of the Zn borophosphate glasses: (a) constant P2O5 content, (b) constant O/P ratio. To assign sample numbers, see Figure 1b. The equal values at 12.5 mol% B2O3 in (a,b) belong to the zbp09 sample.
Figure 6. Full widths at half maximum (fwhm) values of the model Zn−O peaks as shown in Figure 7 for two series of the Zn borophosphate glasses: (a) constant P2O5 content, (b) constant O/P ratio. To assign sample numbers, see Figure 1b. The equal values at 12.5 mol% B2O3 in (a,b) belong to the zbp09 sample.
Solids 05 00024 g006
Figure 7. Model peaks of the Zn−O first-neighbor distances of the series with (a) constant P2O5 content, (b) constant O/P ratio of 3.25, and (c) constant O/P ratio of 3.5 as obtained from Gaussian fitting (solid lines). It is compared with peaks that are calculated for the ZnO4, ZnO5, and ZnO6 units known of related crystal structures [37,38,39] (lines with symbols) and the peak of a binary Zn borate glass (zbp14). The model peaks of the glasses are given with identical heights and the changes in their widths become obvious.
Figure 7. Model peaks of the Zn−O first-neighbor distances of the series with (a) constant P2O5 content, (b) constant O/P ratio of 3.25, and (c) constant O/P ratio of 3.5 as obtained from Gaussian fitting (solid lines). It is compared with peaks that are calculated for the ZnO4, ZnO5, and ZnO6 units known of related crystal structures [37,38,39] (lines with symbols) and the peak of a binary Zn borate glass (zbp14). The model peaks of the glasses are given with identical heights and the changes in their widths become obvious.
Solids 05 00024 g007
Figure 8. Schemes of the bond valence distributions in the PO4, BO4, and ZnO4 units of fully tetrahedral networks: (a) BPO4 crystal, (b) Zn(PO3)2 glass, (c) Zn borophosphate glass with P 2 B 3 and B 4 P 4 units, and (d) Zn borophosphate glass with P 3 B 4 and B 3 P 4  units. The numbers indicate the bond valences given in valence units (vu). Bond valences of 1.0 represent P−O−P or B−O−B bridges. The values of bond valences are not repeated for equivalent bonds.
Figure 8. Schemes of the bond valence distributions in the PO4, BO4, and ZnO4 units of fully tetrahedral networks: (a) BPO4 crystal, (b) Zn(PO3)2 glass, (c) Zn borophosphate glass with P 2 B 3 and B 4 P 4 units, and (d) Zn borophosphate glass with P 3 B 4 and B 3 P 4  units. The numbers indicate the bond valences given in valence units (vu). Bond valences of 1.0 represent P−O−P or B−O−B bridges. The values of bond valences are not repeated for equivalent bonds.
Solids 05 00024 g008
Figure 9. Ratios of the available oxygens for coordination of the Zn and B atoms for two series of the Zn borophosphate glasses with (a) constant P2O5 content of 50 mol% and (b) constant O/P ratio of 3.25. The models are explained in the text. To assign sample numbers, see Figure 1b. The values at 12.5 mol% B2O3 in (a,b) belong to the zbp09 sample.
Figure 9. Ratios of the available oxygens for coordination of the Zn and B atoms for two series of the Zn borophosphate glasses with (a) constant P2O5 content of 50 mol% and (b) constant O/P ratio of 3.25. The models are explained in the text. To assign sample numbers, see Figure 1b. The values at 12.5 mol% B2O3 in (a,b) belong to the zbp09 sample.
Solids 05 00024 g009
Figure 10. Characteristic modifier oxygen interactions in the Zn and Na borophosphate glasses and crystals: (a,b) one or two Zn bound to an ONB (P−ONB), (c) two Zn bound to an ONB (B−ONB), (d) Zn bound to an OB in a B−O−B bridge (Zn2B10O17 crystal [49]), (e) three Na bound to an ONB (P−ONB), (f) Na bound to an OB in a P−O−B bridge (Na5B2P3O13 crystal [17]), (g) Na bound to an OB in a B−O−B bridge (Na2B8O13 crystal [51]), and (h) three Na bound to an ONB (B−ONB). The numbers indicate the bond valences given in valence units (vu).
Figure 10. Characteristic modifier oxygen interactions in the Zn and Na borophosphate glasses and crystals: (a,b) one or two Zn bound to an ONB (P−ONB), (c) two Zn bound to an ONB (B−ONB), (d) Zn bound to an OB in a B−O−B bridge (Zn2B10O17 crystal [49]), (e) three Na bound to an ONB (P−ONB), (f) Na bound to an OB in a P−O−B bridge (Na5B2P3O13 crystal [17]), (g) Na bound to an OB in a B−O−B bridge (Na2B8O13 crystal [51]), and (h) three Na bound to an ONB (B−ONB). The numbers indicate the bond valences given in valence units (vu).
Solids 05 00024 g010
Figure 11. The model peaks of the P−O bond lengths of the zbp13 and nbp13 samples are compared to the P−O distances of a (ZnO)0.10(PbO)0.56(P2O5)0.34 glass [36] with the same O/P ratio = 3.5. A numeral ratio of 3:1 is effective for the P−ONB and P−OB bonds of a phosphate glass with this O/P. The bonds are indicated in the figure. The Zn borophosphate glass shows a similar behavior in its bond lengths. Regarding the Na borophosphate glass and the crystals Na3BP2O8 and Na5B2P3O13 [16,17], some P−O bonds in the P−O−B bridges reach lengths typical for the P−OB bonds in P−O−P bridges. The five curves have been given equal areas.
Figure 11. The model peaks of the P−O bond lengths of the zbp13 and nbp13 samples are compared to the P−O distances of a (ZnO)0.10(PbO)0.56(P2O5)0.34 glass [36] with the same O/P ratio = 3.5. A numeral ratio of 3:1 is effective for the P−ONB and P−OB bonds of a phosphate glass with this O/P. The bonds are indicated in the figure. The Zn borophosphate glass shows a similar behavior in its bond lengths. Regarding the Na borophosphate glass and the crystals Na3BP2O8 and Na5B2P3O13 [16,17], some P−O bonds in the P−O−B bridges reach lengths typical for the P−OB bonds in P−O−P bridges. The five curves have been given equal areas.
Solids 05 00024 g011
Figure 12. The packing densities of the ZnO-B2O3-P2O5 glasses obtained from the mass densities given in [22] are given as numbers near the sample positions in the concentration triangle of the three components. The values of the packing densities are given in percent.
Figure 12. The packing densities of the ZnO-B2O3-P2O5 glasses obtained from the mass densities given in [22] are given as numbers near the sample positions in the concentration triangle of the three components. The values of the packing densities are given in percent.
Solids 05 00024 g012
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hoppe, U.; Freudenberger, P.T.; Brow, R.K.; Bednarčik, J.; Hannon, A.C. Study of the Structure of Zn and Na Borophosphate Glasses Using X-ray and Neutron Scattering Techniques. Solids 2024, 5, 355-374. https://doi.org/10.3390/solids5030024

AMA Style

Hoppe U, Freudenberger PT, Brow RK, Bednarčik J, Hannon AC. Study of the Structure of Zn and Na Borophosphate Glasses Using X-ray and Neutron Scattering Techniques. Solids. 2024; 5(3):355-374. https://doi.org/10.3390/solids5030024

Chicago/Turabian Style

Hoppe, Uwe, Parker T. Freudenberger, Richard K. Brow, Jozef Bednarčik, and Alex C. Hannon. 2024. "Study of the Structure of Zn and Na Borophosphate Glasses Using X-ray and Neutron Scattering Techniques" Solids 5, no. 3: 355-374. https://doi.org/10.3390/solids5030024

APA Style

Hoppe, U., Freudenberger, P. T., Brow, R. K., Bednarčik, J., & Hannon, A. C. (2024). Study of the Structure of Zn and Na Borophosphate Glasses Using X-ray and Neutron Scattering Techniques. Solids, 5(3), 355-374. https://doi.org/10.3390/solids5030024

Article Metrics

Back to TopTop