You are on page 1of 5

B154 Vol. 56, No.

3 / January 20 2017 / Applied Optics Research Article

Co-precipitation of rare-earth-doped Y2O3 and


MgO nanocomposites for mid-infrared
solid-state lasers
VICTORIA L. BLAIR,1,* ZACKERY D. FLEISCHMAN,1 LARRY D. MERKLE,1
NICHOLAS KU,2 AND CARLI A. MOOREHEAD3
1
U.S. Army Research Laboratory, 2800 Powder Mill Rd, Adelphi, Maryland 20783, USA
2
Oak Ridge Institute for Science Education, 1299 Bethel Valley Rd, Oak Ridge, Tennessee 37830, USA
3
Drexel University, 3141 Chestnut St, Philadelphia, Pennsylvania 19104, USA
*Corresponding author: Victoria.L.Blair3.civ@mail.mil

Received 27 July 2016; revised 20 September 2016; accepted 22 October 2016; posted 4 November 2016 (Doc. ID 272057);
published 6 December 2016

Mid-infrared, solid-state laser materials face three main challenges: (1) need to dissipate heat generated in lasing;
(2) luminescence quenching by multiphonon relaxation; and (3) trade-off in high thermal conductivity and small
maximum phonon energy. We are tackling these challenges by synthesizing a ceramic nanocomposite in which
multiple phases will be incorporated into the same structure. The undoped majority species, MgO, will be the
main carrier of high thermal conductivity, and the minority species, Er:Y2 O3 , will have low maximum phonon
energy. There is also an inherent challenge in attempting to make a translucent part from a mixture of two differ-
ent materials with two different indexes of refraction. A simple, co-precipitation technique has been developed in
which both components are synthesized in situ to obtain intimate mixing. These powders compare well to com-
mercially available ceramics, including their erbium spectroscopy, even when mixed as a composite, and can be
air-fired to ∼96% of theoretical density, yielding translucent parts. As the amount of Er:Y2 O3 increases, the
translucency decreases as the number of scattering sites start to coalesce into large patches. If the amount of
Er:Y2 O3 is sufficiently small and dispersed, the yttria grains will be pinned as individuals in a sea of MgO,
leading to optimal translucency.
OCIS codes: (160.0160) Materials; (160.3380) Laser materials; (160.5690) Rare-earth-doped materials.

https://doi.org/10.1364/AO.56.00B154

1. INTRODUCTION incompatible. This is due to light constituent ions most often


There has been much interest in lasers emitting in the mid- needed for high thermal conductivity resulting in strong ion–
infrared spectral region for use in a wide array of applications, ion interactions, and hence large maximum phonon energies.
including atmospheric sensing [1], wind lidar [2], medical pro- The solution to these challenges could be found in the use of
cedures [3], infrared counter measure technologies, molecular polycrystalline, or ceramic, materials. With well-chosen forma-
spectroscopy, and optical pumping of longer wavelength solid- tion processes, such as performing the sintering procedure with
state lasers. However, a set of thermal management challenges nanosized powders, optical quality can be comparable to that of
arise in pursuit of high-output power lasers operating at mid- single crystals. In fact, rare-earth-doped YAG (yttrium alumi-
infrared wavelengths (λ > 2 μm). First, otherwise-excellent num garnet) ceramics have demonstrated similar or better laser
gain materials may cause luminescence quenching by multi- efficiency compared to their single crystal counterparts [6,7].
phonon relaxation between the relatively closely spaced upper Polycrystalline ceramic materials offer a myriad of other bene-
and lower laser levels. This can be ameliorated by using gain fits as well. These include greater flexibility in size and shape,
materials with small maximum phonon energies. Second, even sintering temperatures considerably lower than melting tem-
the most efficient gain media convert some fraction of the peratures, and potentially lower costs [8].
pump energy into thermal energy, resulting in serious beam Ceramic nanocomposites, in particular, could alleviate the
problems such as thermal lensing and depolarization [4,5]. problem of incompatible host requirements for low quenching
Unfortunately, the material requirements for low maximum and high thermal conductivity [9–12]. Nanocomposites incor-
phonon energy and high thermal conductivity are very often porate multiple phases into the same structure to share the

1559-128X/17/03B154-05 Journal © 2017 Optical Society of America


Research Article Vol. 56, No. 3 / January 20 2017 / Applied Optics B155

overall benefits of the individual constituent materials. of water added was proportional to the solubility limit of mag-
Recently, nanocomposites composed of MgO and Y 2 O3 have nesium nitrate in water. Thermogravimetric analysis was used
been prepared to produce infrared-transmitting windows with to determine the gram of cations per gram of hydrated nitrate.
increased durability and thermal shock resistance [13–19]. To This is imperative to achieve precise concentrations of
reduce optical scattering to laser and optical standards, the grain rare-earth cations in the final material.
size must be substantially smaller than the wavelength of light In addition to the reactant solution, a basic solution was
[20]. This is especially important where there is a significant made in which the chemical reaction would take place. The
difference in the index of refraction between the different basic solution consisted of 2% by weight of ammonium bicar-
constituents. bonate in DI water with ammonium hydroxide added until a
Generally speaking, for a collection of fine particles in con- pH of ∼10 was obtained. Finally, the reactant solution was
tact with one another their size will increase if the system is held added drop-wise to the basic solution in a fume hood while
at a sufficiently high temperature [21,22]. To obtain a fully maintaining the pH at ∼10 using small amounts of ammonium
dense material, high temperature is required to drive densifica- hydroxide. Once the reactant solution was exhausted, the pH
tion, ultimately leading to grain growth. If the grains start out was stabilized at ∼10, and the resulting suspension was allowed
smaller (on the order of nanometers), a larger driving force for to age overnight under constant stirring.
densification (reduction of surface area) leads to a smaller The next day, the suspension was filtered from the salty
overall grain size at the end of the process. solution. The resulting powder was then washed in isopropyl
This paper describes a simple, co-precipitation technique to alcohol and filtered; this was repeated for three total washings.
synthesize nanopowder composite material consisting of a ma- The powder was then dried in an oven at ∼65°C. Once dry,
jority MgO phase and a minority Er:Y 2 O3 phase. The relatively the powder was gently crushed and calcined at a heating rate of
low maximum phonon energy of Y 2 O3 (∼590 cm−1 [23]) 10°C/min and held at 1200°C for 30 min. The calcined
makes it suitable to host the rare-earth fluorescing ion. As powders were loaded into a 1 cm die and uniaxially pressed
long as the volume fraction of the Y 2 O3 phase (TC  into disks. The pellets were then vacuum-sealed in clean
8–13 W∕mK [24] is kept below 15%, the effective thermal bags and further compressed in a cold isostatic pressure
conductivity of the composite will remain close to the higher (CIP) vessel. The pellets were held at a maximum pressure
value of MgO (45–70 W/mK [25] according to the established of 30 k psig for 30 s. The pressed pellets were removed from
Bruggeman model [26]. their bags and sintered at 1550°C for 1 h with a heating rate of
Currently, ceramic composites are made by simple mechani- 10°C/min.
cal mixing of the components, introducing complexities such as The ratio of Er:Y 2 O3 to MgO was varied by around 10%.
impurities from the milling media/mixing method, settling of This is a compromise between having sufficiently little
granules by density during slurry drying, and re-agglomeration Er:Y 2 O3 to inhibit its grain growth and limit optical scattering,
of particles during drying [27]. Recent work on ceramic nano- and having enough to provide adequate volume-averaged
composites has examined transformation processing or other erbium concentration for eventual lasing. With 2% of the Y
wet processing techniques to avoid the difficulties that arise replaced by erbium, the volume-averaged erbium concentration
during mechanical mixing methods [9–11,14,17,27,28]. for 10% Y 2 O3 is ∼5 × 1019 cm−1 . This average is only slightly
However, most of this work examines a 1:1 ratio of the phases lower than the concentrations often used in laser experiments
while we aim to explore smaller additions of a minority phase. on Er:Y 2 O3 and closely related materials, and the 2% atomic in
We aim to use a one-pot synthesis method in which the the Er:Y 2 O3 is within the range for which the upper state life-
materials are intimately mixed as liquids before the product time is reported to remain approximately constant with little
crystallizes and crashes out of the solution. Careful calcination quenching [29–31].
of the precipitated powder leads to nanocrystallites at moder-
ately low temperatures (<1000°C), useful for preventing large B. Characterization of Materials
grains in the final ceramic composite. In addition to the syn- The sintered pellets were measured in a powder diffractometer
thesis and powder characterization, we will report on the (Rigaku MiniFlex II) to determine semi-quantitative composi-
spectroscopic behavior of the nanocomposite in comparison tion (MgO to Y 2 O3 ratio). Calibration specimens were batched
to commercial and in-house synthesized Er:Y 2 O3 . Overall, with known MgO∕Y 2 O3 ratios and measured by x-ray diffrac-
we aim to create a nanocomposite with similar spectroscopic tion (XRD) to generate a calibration curve. Selected pellets
behavior to pure Er:Y 2 O3 while having the high thermal were fractured and imaged in a Hitachi S4700 scanning elec-
conductivity of MgO. tron microscope equipped with a yttrium aluminum garnet
(YAG) backscatter detector to highlight the phase contrast.
ImageJ, Image Processing and Analysis in Java (NIH,
2. EXPERIMENTAL PROCEDURE Bethesda, MD), was used to estimate the amount of Y 2 O3
A. Materials Synthesis in the specimen image. The same pellets were also ion milled
A co-precipitation method was used to synthesize amorphous through the cross-section in a Leica TIC 3X ion mill to
precursor powders in an aqueous environment. Prior to the generate a flat, polished area for further scanning electron
chemical precipitation, a reactant solution was prepared, con- microscopy (SEM) observation.
sisting of yttrium nitrate (high purity 99.99%), magnesium Fluorescence spectra of the ∼1.5 μm 4 I13∕2 → 4 I15∕2 infra-
nitrate (high purity 99.99%), and erbium nitrate (high purity red transition of Er3 were obtained using a Horiba iHR320
99.99%) dissolved in deionized (DI) water. The total amount monochromator with a 1.5 μm blazed grating and equipped
B156 Vol. 56, No. 3 / January 20 2017 / Applied Optics Research Article

with a liquid-nitrogen-cooled, extended-range InGaAs photo-


diode for detection. Excitation at 804.1 nm into the 4 I9∕2
manifold was supplied by a Spectra Physics Tsunami Ti:sap-
phire laser operated in CW mode. Fluorescence lifetime mea-
surements of the 4 I13∕2 manifold were acquired by exciting the
sample with a modulated ∼808 nm laser diode. The decay sig-
nal was detected with an uncooled InGaAs photodiode
equipped with a 1500 nm long-pass filter, and was analyzed
using a Tektronix TDS7104 digital oscilloscope.

3. RESULTS AND DISCUSSION


A. Chemical Composition
XRD patterns of the pellets showed all expected peaks for MgO Fig. 1. Fracture surface of the sintered pellet with 5% by weight
and Y 2 O3 . Er2 O3 is isostructural with Y 2 O3 , so the peaks Er0.04 Y 1.96 O3 , imaged by a YAG backscatter detector. The lighter
phase is the Er:Y 2 O3 .
would overlap if there were any present in the pattern. In
any case, the amount of erbium in the samples is likely too
small to be able to measure by standard-powder XRD
alone. Table 1 shows the targeted chemical compositions in
comparison to the measured composition. The samples with
smaller amounts of Y 2 O3 are closer to the targeted chemical
composition than the material with a target of 15% by weight
Er:Y 2 O3 .
The discrepancy between the target and measured compo-
sitions for the 15% Er:Y 2 O3 sample is possibly due to a change
in the isoelectric point of the solution from which the powder
was precipitated. To precipitate powder from mixing an acid
and a base, the pH must pass through the isoelectric point
of the solution for the reaction to occur. From literature, it
is shown that MgOH2 can be precipitated from a solution
of MgNO3 2 at a pH of ∼10 [32]. Generally, precipitation
of hydroxide precursors to Y 2 O3 occurs below a pH of 8
[33–35], lower than what was used for these nanocomposites. Fig. 2. Polished surface of the sintered pellet with 5% by weight
The increase in yttrium nitrate in the mixed solutions may be Er0.04 Y 1.96 O3 , imaged by a YAG backscatter detector. The lightest
causing a shift in the isoelectric point of the solution, resulting phase is the Er:Y 2 O3 .
in a preferred precipitation of the yttrium precursor over the
magnesium precursor. This would result in a higher magnesium
content in the final calcined material. Higher magnesium con- The discrepancy in the estimated yttria content from frac-
tent than the 15% target was observed in repeated precipitated ture surface to polished surface is likely due to nonuniform frac-
batches, with an average of 25% measured in the final material. ture of the specimen; the cracks propagate through the MgO
grains and around the yttria grains, leading to more yttria grains
B. Microstructure being visible in the part. The same trend was observed in the
The fractured surface of the sintered pellet with 5% by weight other specimens with higher yttria content. However, when a
of Er:Y 2 O3 , corresponding to 3.6% by volume, is shown in clean surface is polished flat through the grains, the yttria
Fig. 1. Pixel counting in ImageJ gave a Y 2 O3 content of content is closer to the amount measured by XRD.
∼14% by area, significantly larger than the targeted volume There is still some discrepancy between the estimated yttria
content of 3.6%. However, a polished surface of the specimen, content of the polished surface in comparison to what was
shown in Fig. 2, gave a Y 2 O3 content of 5.7%. batched, 5.7% for the former and 3.6% for the latter (both
by volume). The difference here may be attributed to process
control and understanding, which still need to be refined and
Table 1. Targeted and Measured Chemical optimized for this particular synthesis process. The solubility of
Compositions of the Sintered Pellets MgO in the precipitant suspension may still be an issue, even
MgO (wt %) Er0.04 Y1.96 O3 (wt %) though a pH of 10 was maintained through the process.
Figure 2 also shows some open porosity which cannot be
Target Measured Target Measured
observed in the fracture surface shown in Fig. 1. Finally, the
95 95 5 4.7 yttria grains are significantly smaller than the magnesia grains,
90 89 10 11 indicating that the magnesia is the dominating growth phase
85 67 15 33
that traps the yttria inside the magnesia. This is an important
Research Article Vol. 56, No. 3 / January 20 2017 / Applied Optics B157

the Y 2 O3 host material [36]. Figure 4 shows that the spectral


signature of the nanocomposite is nearly identical to that of the
Konoshima sample, and both signatures agree well with liter-
ature [37,38]. Close inspection of the two curves shows that the
peaks in the nanocomposite material are slightly broader,
indicative of slightly poorer Y 2 O3 crystal quality.
Fluorescence lifetimes of the 4 I13∕2 level were taken for the
same nanocomposite sample (average erbium concentration of
0.1%) and for a pulverized piece of the Konoshima Er:Y 2 O3
Fig. 3. Backlit macro images of the sintered pellet fragments; ceramic with erbium concentration of 3%. The nanocomposite
(a) 5% by weight Er:Y 2 O3 and (b) 10% by weight Er:Y 2 O3 . An gave a lifetime of 8.4 ms, while the Konoshima material showed
opaque specimen would appear black in this image. a slightly longer 9.4 ms lifetime. These are comparable to recent
measurements by Brown et al. of a 0.5% Er:Y 2 O3 thick
ceramic sample, which showed a fluorescence lifetime of
result to obtain a transparent part, as the secondary phase 8.6 ms [37]. The concentration dependence of these differences
(yttria) must be small and disperse to reduce scattering. in lifetime is consistent with the effects of reabsorption of the
Figure 3 shows a backlit macro image of two pellets consist- fluorescence.
ing of 5% and 10% Er:Y 2 O3 . An opaque specimen would be
black when backlit. There is a visibly observable difference be-
tween the two pellets with the 10% Er:Y 2 O3 specimen being 4. CONCLUSIONS
less translucent. This is expected as the Y 2 O3 is acting as a light Sintered pellets of varying ratios of MgO and Er:Y 2 O3 were
scattering site. synthesized, processed, and characterized. The co-precipitation
process to make these materials has shown some repeatability
C. Spectroscopy for low Y 2 O3 containing compositions (<10%). It is possible
Optical spectroscopy measurements of the 4 I13∕2 → 4 I15∕2 in- that the precipitation of the magnesium precursor is suppressed
frared transition of Er3 were used to further characterize the as the yttria content increases, thus leading to a final compo-
nanocomposite samples. Figure 4 shows the spectrum of the sition that does not match the targeted ratio. Research is on-
5% by weight Er:Y 2 O3 nanocomposite sample along with that going to determine how the isoelectric point varies with varying
from an Er:Y 2 O3 ceramic sample obtained from Konoshima magnesium and yttria ratios in the reactant solutions prior to
Chemical Co, Ltd. (Japan). This Konoshima material repre- precipitation.
sents the pinnacle of rare-earth-doped ceramic sesquioxide The microstructure of the sintered pellets shows large MgO
fabrication and is used in this study as a benchmark to gauge grains with small, dispersed Y 2 O3 grains with a small amount
the quality of the nanocomposite samples. of residual open porosity. The Y 2 O3 grains appear at both triple
The spectral signature of this transition is determined by the points and inside the MgO grains, indicating that the Y 2 O3
crystal-field splitting of Er3 levels involved and is unique to grains are swallowed up by the MgO during densification.
The pellets exhibited some level of translucency after pres-
sure-less sintering, and the translucency appears to decline as
the Y 2 O3 content increases. Research will be ongoing to obtain
fully dense samples by sintering under an applied pressure.
When fully dense material is achieved, thermal conductivity
measurements will be performed to compare to the calculated
values from the Bruggeman model [26].
Fluorescence measurements showed good agreement
between nanocomposite samples and Er:Y 2 O3 ceramic from
Konoshima Chemical Co, Ltd. (Japan). This indicates that
the nanocomposites are comparable in quality to the state-
of-the-art ceramic sesquioxide material on the market today.

Funding. Oak Ridge Institute for Science and Education


(ORISE) (1120-1120-99); Science & Engineering
Apprenticeship Program (SEAP) (W911SR-15-2-0001).

Acknowledgment. C. A. M. was sponsored by an Army


Educational Outreach Program Youth Science Cooperative
Agreement with Battelle Memorial institute. N. K. was sup-
ported by an appointment to the Postgraduate Research
Fig. 4. Fluorescence spectra of the Er34 I13∕2 →4 I15∕2 transition for Participation Program at the U.S. Army Research
a nanocomposite with 5% by weight Er:Y 2 O3 and for a Konoshima Laboratory (USARL) administered by the Oak Ridge
Er:Y 2 O3 ceramic sample. Baselines have been offset for clarity. Institute for Science and Education through an interagency
B158 Vol. 56, No. 3 / January 20 2017 / Applied Optics Research Article

agreement between the U.S. Department of Energy 19. S. Xu, J. Li, C. Li, Y. Pan, and J. Guo, “Infrared-transparent Y2O3–MgO
nanocomposites fabricated by the glucose sol-gel combustion and
and USARL.
hot-pressing technique,” J. Am. Ceram. Soc. 98, 2796–2802
(2015).
REFERENCES 20. R. Apetz and M. P. B. van Bruggen, “Transparent alumina: a light-
1. T. M. Taczak and D. K. Killinger, “Development of a tunable, narrow- scattering model,” J. Am. Ceram. Soc. 86, 480–486 (2003).
linewidth, cw 2.066-μm Ho:YLF laser for remote sensing of atmos- 21. W. D. Kingery, H. K. Bowen, and D. R. Uhlmann, Introduction to
pheric CO2 and H2O,” Appl. Opt. 37, 8460–8476 (1998). Ceramics, 2nd ed. (Wiley, 1976), p. 1032.
2. S. M. Hannon and J. A. Thomson, “Aircraft wake vortex detection and 22. M. N. Rahaman, Ceramic Processing and Sintering, 2nd ed. (CRC
measurement with pulsed solid-state coherent laser radar,” J. Mod. Press, 2003).
Opt. 41, 2175–2196 (1994). 23. Y. Repelin, C. Proust, E. Husson, and J. M. Beny, “Vibrational spec-
3. H. W. Kang, H. Lee, J. Petersen, J. H. Teichman, and A. J. Welch, troscopy of the c-form of yttrium sesquioxide,” J. Solid State Chem.
“Investigation of stone retropulsion as a function of Ho:YAG laser 118, 163–169 (1995).
pulse duration,” Proc. SPIE 6078, 607815 (2006). 24. I. L. Snetkov, D. E. Silin, O. V. Palashov, E. A. Khazanov, H. Yagi, T.
4. W. A. Clarkson, “Thermal effects and their mitigation in end-pumped Yanagitani, H. Yoneda, A. Shirakawa, K.-I. Ueda, and A. A. Kaminskii,
solid-state lasers,” J. Phys. D 34, 2381–2395 (2001). “Study of the thermo-optical constants of Yb doped Y2O3, Lu2O3 and
5. W. Koechner, Solid-State Laser Engineering, 6th ed., Optical Sc2O3 ceramic materials,” Opt. Express 21, 21254–21263 (2013).
Sciences (Springer, 2006), Vol. 1, p. 748. 25. G. A. Slack, “Thermal conductivity of MgO, Al2O3, MgAl2O4, and Fe3O4
6. J. Dong, A. Shirakawa, K.-I. Ueda, H. Yagi, T. Yanagitani, and A. A. crystals from 3 to 300 K,” Phys. Rev. 126, 427–441 (1962).
Kaminskii, “Laser-diode pumped heavy-doped Yb:YAG ceramic 26. J. P. Angle, Z. Wang, C. Dames, and M. L. Mecartney, “Comparison of
lasers,” Opt. Lett. 32, 1890–1892 (2007). two-phase thermal conductivity models with experiments on dilute
7. Y. Qi, X. Zhu, Q. Lou, J. Ji, J. Dong, and Y. Wei, “Nd:YAG ceramic ceramic composites,” J. Am. Ceram. Soc. 96, 2935–2942 (2013).
laser obtained high slope-efficiency of 62% in high power applica- 27. T. Stefanik, R. Gentilman, and P. Hogan, “Nanocomposite optical
tions,” Opt. Express 13, 8725–8729 (2005). ceramics for infrared windows and domes,” Proc. SPIE 6545,
8. J. Kong, D. Y. Tang, C. C. Chan, J. Lu, K. Ueda, H. Yagi, and T. 65450A (2007).
Yanagitani, “High-efficiency 1040 and 1078 nm laser emission of a 28. B. H. Kear, J. Colaizzi, W. E. Mayo, and S. C. Liao, “On the processing
Yb:Y2O3 ceramic laser with 976 nm diode pumping,” Opt. Lett. 32, of nanocrystalline and nanocomposite ceramics,” Scr. Mater. 44,
247–249 (2007). 2065–2068 (2001).
9. L. L. Beecroft and C. K. Ober, “Nanocomposite materials for optical 29. N. Ter-Gabrielyan, L. D. Merkle, G. A. Newburgh, and M. Dubinskii,
applications,” Chem. Mater. 9, 1302-1317 (1997). “Resonantly-pumped Er3+:Y2O3 ceramic laser for remote CO2 monitor-
10. S. Chawla, K. Jayanthi, H. Chander, D. Haranath, S. K. Halder, and M. ing,” Laser Phys. 19, 867–869 (2009).
Kar, “Synthesis and optical properties of ZnO/MgO nanocomposite,” 30. N. Ter-Gabrielyan, L. D. Merkle, A. Ikesue, and M. Dubinskii, “Ultralow
J. Alloys Compd. 459, 457–460 (2008). quantum-defect eye-safe Er:Sc2O3 laser,” Opt. Lett. 33, 1524–1526
11. M. Llusar, V. Royo, J. A. Badenes, M. A. Tena, and G. Monrós, (2008).
“Nanocomposite Fe2O3–SiO2 inclusion pigments from post-functional- 31. T. Li, K. Beil, C. Kränkel, and G. Huber, “Efficient high-power
ized mesoporous silicas,” J. Eur. Ceram. Soc. 29, 3319–3332 (2009). continuous wave Er:Lu2O3 laser at 2.85 μm,” Opt. Lett. 37, 2568–
12. J. Wang, L. Zhang, D. Chen, E. H. Jordan, and M. Gell, “Y2O3–MgO– 2570 (2012).
ZrO2 infrared transparent ceramic nanocomposites,” J. Am. Ceram. 32. B. Vatsha, P. Tetyana, P. M. Shumbula, J. C. Ngila, L. M. Sikhwivhilu,
Soc. 95, 1033–1037 (2012). and R. M. Moutloali, “Effects of precipitation temperature on
13. D. C. Harris, L. R. Cambrea, L. F. Johnson, R. T. Seaver, M. nanoparticle surface area and antibacterial behaviour of Mg(OH)2
Baronowski, R. Gentilman, C. Scott Nordahl, T. Gattuso, S. and MgO nanoparticles,” J. Biomater. Nanobiotechnol. 4, 365–373
Silberstein, P. Rogan, T. Hartnett, B. Zelinski, W. Sunne, E. Fest, (2013).
W. Howard Poisl, C. B. Willingham, G. Turri, C. Warren, M. Bass, 33. L. Muresan, E.-J. Popovici, R. Grecu, and L. B. Tudoran, “Studies on
D. E. Zelmon, and S. M. Goodrich, “Properties of an infrared-transpar- the synthesis of europium activated yttrium oxide by wet-chemical
ent MgO:Y2O3 nanocomposite,” J. Am. Ceram. Soc. 96, 3828–3835 method: 1. Influence of precursor quality on phosphor photolumines-
(2013). cence properties,” J. Alloys Compd. 471, 421–427 (2009).
14. D. Jiang and A. K. Mukherjee, “Spark plasma sintering of an infrared- 34. Z. Huang, X. Sun, Z. Xiu, S. Chen, and C.-T. Tsai, “Precipitation
transparent Y2O3–MgO nanocomposite,” J. Am. Ceram. Soc. 93, 769– synthesis and sintering of yttria nanopowders,” Mater. Lett. 58,
773 (2010). 2137–2142 (2004).
15. C. K. Muoto, E. H. Jordan, M. Gell, and M. Aindow, “Phase homo- 35. L. Wen, X. Sun, Q. Lu, G. Xu, and X. Hu, “Synthesis of yttria nano-
geneity in Y2O3–MgO nanocomposites synthesized by thermal powders for transparent yttria ceramics,” Opt. Mater. 29, 239–245
decomposition of nitrate precursors with ammonium acetate (2006).
additions,” J. Am. Ceram. Soc. 94, 4207–4217 (2011). 36. N. C. Chang, J. B. Gruber, R. P. Leavitt, and C. A. Morrison, “Optical
16. C. K. Muoto, E. H. Jordan, M. Gell, and M. Aindow, “Effects of spectra, energy levels, and crystal-field analysis of tripositive rare
precursor chemistry on the structural characteristics of Y2O3–MgO earth ions in Y2O3. I. Kramers ions in C2 sites,” J. Chem. Phys. 76,
nanocomposites synthesized by a combined sol-gel/thermal decom- 3877–3889 (1982).
position route,” J. Am. Ceram. Soc. 94, 372–381 (2011). 37. E. E. Brown, U. Hömmerich, A. Bluiett, C. Kucera, J. Ballato, and S.
17. J. Wang, D. Chen, E. H. Jordan, and M. Gell, “Infrared-transparent Trivedi, “Near-infrared and upconversion luminescence in Er:Y2O3
Y2O3–MgO nanocomposites using sol-gel combustion synthesized ceramics under 1.5 μm excitation,” J. Am. Ceram. Soc. 97,
powder,” J. Am. Ceram. Soc. 93, 3535–3538 (2010). 2105–2110 (2014).
18. S. Xu, J. Li, C. Li, Y. Pan, and J. Guo, “Hot pressing of infrared- 38. L. D. Merkle and N. Ter-Gabrielyan, “Er-doped sesquioxides for 1.5-
transparent Y2O3–MgO nanocomposites using sol-gel combustion micron lasers—spectroscopic comparisons,” Proc. SPIE 8733,
synthesized powders,” J. Am. Ceram. Soc. 98, 1019–1026 (2015). 87330H (2013).

You might also like