Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Reducing the refractive index by replacing an [AlPO4]° unit with [BPO4]° in fused silica: comment

Open Access Open Access

Abstract

Recent computational models led to the conclusion that replacing AlPO4 with BPO4 when doped into SiO­2 results in a lower refractive index of the subsequent glass [Opt. Mater. Express 13(4), 935 (2023) [CrossRef]  ]. This finding is conceptually important since there is always value in having a greater diversity of dopant systems by which to tailor the refractive index profile of silica fibers, not to mention their spectroscopic and other properties. This is especially the case for high energy laser fibers, which are quite highly modified from a compositional perspective relative to telecom fibers. However, as this Comment shows, experimental results on the BPO4-SiO2 join, dating back nearly 75 years, clearly refute the theoretical predictions. Also refuted are the computed values and trends along the AlPO4-SiO2 join, a very important material system for laser fibers.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

The growing demand for optical fibers with enhanced performance, especially as relates to high power laser systems, has led to a renaissance in fiber materials [1] and waveguide designs [2]. These two principal design considerations are not mutually exclusive and are, indeed, symbiotic as there are long established connections between the impact of the fiber’s core glass composition and the optical and acoustic properties that lead to an optimum design. The most canonical connection between fiber materials and waveguide design is whether a given dopant raises or lowers the effective refractive index of the resultant glass when added to the predominantly silica host; see, for example Fig. 4.2 in Ref. [2]. Secondary considerations, though no less important in high energy fiber applications, are how said dopants influence nonlinearities such as Brillouin and Raman scattering, and transverse mode instabilities [3,4].

There are a greater number of index-raising additives that can be doped into silica than index-lowering ones. This is further compounded by the processes used to fabricate silica optical fiber preforms, which further limits the available index modifying compounds and their doping concentrations into silica due to the very high processing temperatures [1]. Index-lowering compounds are important because, as noted, most additives raise the index. Thus, having species to counter this trend is critical to controlling the refractive index profile in the fiber, hence related fiber characteristics, such as the number of propagating modes and mode field diameter.

Commonly employed index-lowering species are fluorine (F) and boria, B2O3 [58]. Another index-lowering compound that has grown in importance is AlPO4 which results during preform fabrication from the reaction of P2O5 and Al2O3 [9]. AlPO4 is isostructural with SiO2 in the glass matrix and, in addition to lowering the effective refractive index, also decreases the glasses’ thermo-optic coefficient and Brillouin gain coefficient [10].

Recently reported in Xiang, et al., is molecular modeling of silica glasses doped with AlPO4 and BPO4 [11]. Their theoretical results suggest that doping SiO2 with BPO4 results in a lower refractive index than when silica is doped with AlPO4. Indeed, having another index-lowering dopant to silica would be noteworthy. Unfortunately, this computational result is not consistent with experimental literature dating back at least to 1950, nor does it agree with straight-forward additivity models. This Comment summarizes the literature reports on glass in the BPO4-SiO2 system and corroborates those historical experimental results with simple additivity relations.

2. Results and discussion

Xiang, et al., employ ab initio molecular dynamic simulations to construct SiO2-AlPO4 and SiO2-BPO4 glass structural models. Density functional theory was then used to estimate various properties, including refractive indices. Major findings therein include that “The refractive index of the SiO2, SiO2-AlPO4, and SiO2-BPO4 are 1.4752, 1.4856, and 1.4766 at 760 nm” (page 942) and “It can be found that the refractive index of the SiO2-BPO4 is lower than that of the SiO2-AlPO4, as shown in Figures S8 and S9, which is consistent with the above calculation results (96 atoms)” (supplemental information). For the purposes of this Comment, the BPO4 and AlPO4 concentrations in Xiang, et al., are interpreted as being about 3.2 mole percent, though the value is never explicitly stated. This is inferred from the simulations starting with 96 atoms (32 SiO2 units), from which two SiO2 units are replaced by an XPO4, yielding a presumed concentration of 100/(32-1) mole% XPO4. Arguably, this is relatively light doping.

The refractive index along the BPO4-SiO2 join will be discussed in greater detail below. However, it is worth noting that the stated index trends cannot be right on their face because the addition of AlPO4 into SiO2 has been known for almost 40 years to result in a lower refractive index [9]. Xiang, et al., compute the refractive index of AlPO4 doped SiO2 to be higher than SiO2 across the entirety of the visible spectral range (see Fig. 6 in Ref. [11]).

2.1 Established results in the literature

The refractive index of SiO2 glass is well established and best fit by the Sellmeier relationship given in Malitson [12]. At a wavelength of 760 nm, the refractive index should be about 1.455, measurably lower than the 1.475 modeled in Xiang, et al. The refractive indices for AlPO4 and BPO4 are slightly more difficult to define since they are not classical glass forming compounds. As crystals, like quartz (crystalline SiO2 at equilibrium ambient conditions), AlPO4 (Berlinite) and BPO4 are uniaxial, possessing two refractive indices. The ordinary and extraordinary refractive indices for BPO4 were measured on high quality single crystals and provided as raw data and in Sellmeier form by Li, et al. [13]. Based on this experimentally determined data, the Sellmeier forms yield refractive indices of 1.594 (ordinary) and 1.588 (extraordinary) at a wavelength of 760 nm. Corroboration of the BPO4 refractive index dates back to 1950 [14], which places the refractive indices between 1.595 and 1.601. Though the exact wavelength of measurement was not specified in those works, given the publication date, the values would almost certainly have been in the visible.

More importantly for this Comment, glasses along the BPO4-SiO2 join were fabricated over a nearly complete composition range and their refractive indices reported in 1955 [15]. Clearly shown there, and included in Fig. 1 here, is that the refractive index along the BPO4-SiO2 join increases markedly with the addition of BPO4 into SiO2, in opposition to the assertion in Xiang, et al., that “the refractive index of the SiO2-BPO4 is close to the pure silica, thus not altering the refractive index profile of the silica-based fiber core” [11]. Additionally, as noted, the refractive index of AlPO4 doped SiO2 decreases with increasing AlPO4 concentration. Calculations found in Xiang, et al., suggest instead that the index increases.

 figure: Fig. 1.

Fig. 1. (a) Refractive index along the BPO4-SiO2 join including experimental data on glasses [15], calculated values based on additivity of the end-member indices, and theorized results [11]. Also included for reference are the deduced values for the AlPO4-SiO2 join based on modeled and experimental measurements [10]. (b) Deduced refractive index along the BPO4-SiO2 join overlaid by the experimental data [15]. Connecting lines are guides-to-the-eye.

Download Full Size | PDF

2.2 Material additivity results

The stated bulk indices of BPO4 and SiO2 would naturally suggest that the refractive index of a composition within the binary (BPO4-SiO2) system would be somewhere between the end-member values. Since there are no known or presumed reactions between the BPO4 and SiO2, unlike the case between Al2O3 and P2O5 yielding AlPO4, material property additivity should apply and provide a reasonable estimation of refractive index, among other properties, including the density [16]. To illustrate this, linear additivity is applied to the Ref. [15] data shown in Fig. 1(a). The additivity of the refractive index is quite simple. The aggregate glass takes on a net refractive index that is the weighted sum of those of the constituents (BPO4 and SiO2). The weighting is against the total volume fraction occupied by each constituent. Converting to mole% from volume fraction requires knowledge of the molar volume, or alternatively the mass density. As it turns out, from a reasonably complete set of compositional data, the latter can be estimated. The procedure is as follows: The base SiO2 refractive index was set to that of Ref. [15] (1.456, note that a wavelength was not provided) and its density was assumed to be 2200 kg/m3 as in [11]. Then, both the refractive index and density for glassy BPO4 can become fit parameters in the model. The former value sets the endpoint of the fitting (where the BPO4 concentration is 100%) and the latter sets the curvature of the graph. Due to some variation in the data, a range of densities 2640 ± 300 kg/m3 around nBPO4 = 1.545 best fits the data (the extremes are shown plotted in Fig. 1(b)). In the modeling, a single BPO4 unit has a molar mass of 105.78 g/mol and the reaction P2O5 + B2O3 → 2BPO4 is assumed to take place with 100% efficiency. In comment, and to a large extent as expected, both the refractive index and density of glassy BPO4 are somewhat lower than what is observed in the crystalline phase (density ∼2800 kg/m3 [17]). However, both are significantly lower in pure SiO2. Using the silica values from [11] and 2640 kg/m3, nBPO4 = 1.504 is deduced from [11]. Similarly, using a density of 2233 kg/m3 for AlPO4 [10], nAlPO4 = 1.640 is deduced from [11]. In short, the former index is lower and the latter much higher than reported literature values.

A similar set of refractive index data can be found in a more recent paper [18]. That data cuts across the compositional range (30-x)P2O5-xB2O3-70SiO2 and index measurements were performed in the visible. That data exhibits a discontinuity at the BPO4 join and the system is well modeled as a quaternary glass (SiO2, P2O5, B2O3, BPO4) with index and density values being (1.462, 1.549, 1.452, 1.540) and (2200, 2390, 1820, 2640) kg/m3, respectively. Again, it is assumed that the reaction P2O5 + B2O3 → 2BPO4 is 100% efficient. Due to limited space, that fitting will not be shown here. However, it is worth noting that the two data sets, published 50 years apart, corroborate extremely well indicating that the impact of doping BPO4 into SiO2 has been known for some time, but so far perhaps overlooked by the broader optical fiber community. Being experimentalists, the authors of this Comment are not positioned to know the source of uncertainties that can be introduced in molecular dynamics modeling. That said, this is a cautionary note of the possible impact of purely theoretical and modeling efforts absent thorough literature analysis or simple additivity calculations.

3. Conclusions

This Comment aims to correct statements and conclusions made in the recent work by Xiang et al. [11]. Based both on direct experimental evidence and as illustrated with additivity relationships on selected experimental results, the refractive index along the BPO4-SiO2 join is higher than that for AlPO4-SiO2 and is expected to significantly raise, not lower, and the addition of AlPO4 into SiO2 reduces, not raises, the refractive index.

Funding

US DoD Joint Directed Energy Transition Office through the Office of Naval Research (N00014-17-1-2546).

Disclosures

The authors declare no conflict of interest.

Data availability

Data underlying the results are available in Ref. [15].

References

1. P. D. Dragic, M. Cavillon, and J. Ballato, “Materials for optical fiber lasers: A review,” Appl. Phys. Rev. 5(4), 041301 (2018). [CrossRef]  

2. L. Dong and B. Samson, Fiber Lasers: Basics, Technology, and Applications (CRC Press, 2016).

3. J. Ballato, M. Cavillon, and P. D. Dragic, “A unified materials approach to mitigating optical nonlinearities in optical fiber. I. Thermodynamics of optical scattering,” Int. J. Appl. Glass Sci. 9(2), 263–277 (2018). [CrossRef]  

4. T. W. Hawkins, P. D. Dragic, N. Yu, A. Flores, M. Engholm, and J. Ballato, “Kilowatt power scaling of an intrinsically low Brillouin and thermo-optic Yb-doped silica fiber,” J. Opt. Soc. Am. 38(12), F38–49 (2021). [CrossRef]  

5. J. W. Fleming and D. L. Wood, “Refractive index dispersion and related properties in fluorine doped silica,” Appl. Opt. 22(19), 3102–3104 (1983). [CrossRef]  

6. H. Wehr and D. Wiechert, “Refractive index and density of fluorine doped silica prepared by the PCVD process,” Mater. Res. Bull. 21(5), 559–566 (1986). [CrossRef]  

7. L. G. Van Uitert, D. A. Pinnow, J. C. Williams, T. C. Rich, R. E. Jaeger, and W. H. Grodkiewicz, “Borosilicate glasses for fiber optical waveguides,” Mater. Res. Bull. 8(4), 469–476 (1973). [CrossRef]  

8. S. H. Wemple, D. A. Pinnow, T. C. Rich, R. E. Jaeger, and L. G. Van Uitert, “Binary SiO2–B2O3 glass system: refractive index behavior and energy gap considerations,” J. Appl. Phys. 44(12), 5432–5437 (1973). [CrossRef]  

9. D. J. DiGiovanni, J. B. MacChesney, and T. Y. Kometani, “Structure and properties of silica containing aluminum and phosphorus near the AlPO4 join,” J. Non-Cryst. Solids 113(1), 58–64 (1989). [CrossRef]  

10. N. Yu, T. W. Hawkins, T. V. Bui, M. Cavillon, J. Ballato, and P. D. Dragic, “AlPO4 in silica glass optical fibers: Deduction of additional material properties,” IEEE Photonics J. 11(5), 1–13 (2019). [CrossRef]  

11. G. Xiang, J. Zhang, Z. Xing, L. Huang, Z. Pan, H. Zhang, X. Wang, and W. Hua, “Reducing the refractive index by replacing an [AlPO4]° unit with [BPO4]° in fused silica,” Opt. Mater. Express 13(4), 935–944 (2023). [CrossRef]  

12. I. H. Malitson, “Interspecimen comparison of the refractive index of fused silica,” J. Opt. Soc. Am. 55(10), 1205–1208 (1965). [CrossRef]  

13. Z. Li, Z. Lin, Y. Wu, P. Fu, Z. Wang, and C. Chen, “Crystal growth, optical properties measurement, and theoretical calculation of BPO4,” Chem. Mater. 16(15), 2906–2908 (2004). [CrossRef]  

14. F. A. Hummel and T. A. Kupinski, “Thermal properties of the compound BPO4,” J. Am. Chem. Soc. 72(11), 5318–5319 (1950). [CrossRef]  

15. W. F. Horn and F. A. Hummel, “Notes on the system B2O3-SiO2-P2O5. Part I. The BPO4-SiO2 join,” J. Soc. Glass Technol. 39(187), 113–120 (1955).

16. P. D. Dragic, M. Cavillon, A. Ballato, and J. Ballato, “A unified materials approach to mitigating optical nonlinearities in optical fiber. II. A. Material additivity models and basic glass properties,” Int. J. Appl. Glass Sci. 9(2), 278–287 (2018). [CrossRef]  

17. https://materialsproject.org/materials/mp-3589.

18. B. G. Aitken and R. E. Youngman, “Borophosphosilicate glasses: properties and structure,” Phys. Chem. Glasses: Eur. J. Glass Sci. Technol. B 47(4), 381–387 (2006).

Data availability

Data underlying the results are available in Ref. [15].

15. W. F. Horn and F. A. Hummel, “Notes on the system B2O3-SiO2-P2O5. Part I. The BPO4-SiO2 join,” J. Soc. Glass Technol. 39(187), 113–120 (1955).

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (1)

Fig. 1.
Fig. 1. (a) Refractive index along the BPO4-SiO2 join including experimental data on glasses [15], calculated values based on additivity of the end-member indices, and theorized results [11]. Also included for reference are the deduced values for the AlPO4-SiO2 join based on modeled and experimental measurements [10]. (b) Deduced refractive index along the BPO4-SiO2 join overlaid by the experimental data [15]. Connecting lines are guides-to-the-eye.
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.