Control of Melt Decomposition for the Growth of High Quality

Sep 26, 2014 - Control of Melt Decomposition for the Growth of High Quality. AgGaGeS4 Single Crystals for Mid-IR Laser Applications. J. Rame,. †. B...
0 downloads 0 Views 9MB Size
Article pubs.acs.org/crystal

Control of Melt Decomposition for the Growth of High Quality AgGaGeS4 Single Crystals for Mid-IR Laser Applications J. Rame,† B. Viana,‡ Q. Clement,† J. M. Melkonian,† and J. Petit*,† †

ONERA The French Aerospace Lab, 92320 Chatillon, France IRCP Chimie ParisTech, 11 rue Pierre et Marie Curie, 75231 Paris Cedex 05, France



ABSTRACT: AgGaGeS4 (AGGS) is a promising nonlinear crystal for mid-IR laser applications which could satisfy the lack of materials able to convert a 1.064 μm pump signal (Nd:YAG laser) to wavelengths higher than 4 μm, up to 11 μm. The processing steps of this material are presented in this study. The key issue of AGGS crystal processing is the control of decomposition at high temperature due to the high volatility of GeS2. This study presents solutions to obtain high quality single crystals. AGGS ingots with 28 mm diameter and 70 mm length were grown by the Bridgman−Stockbarger method. The crystals have good homogeneity and high transparency in the 0.5−11.5 μm spectral range, making them suitable for optical experiments. The influence of GeS2 volatility on melt stoichiometry during the AgGaGeS4 processing is outlined, and solutions to improve the crystals quality are presented.



(0.5−11.5 μm) and a residual absorption coefficient as low as 0.05 cm−1 at 1.064 μm, making it compatible with a Nd:YAG laser pumping.6 AgGaGeS4 belongs to the orthorhombic C19 2υFdd2 space group.7 This material demonstrated a high laser damage threshold, 50 MW·cm−2 at 1.064 μm (τ = 10 ns),8 1.1 J·cm−2 (73 MW·cm−2) at 2.05 μm (τ = 15 ns; 10 kHz),6 and 230 MW·cm−2 (τ = 30 ns) at 9.55 μm,9 which makes it promising for high power applications. The nonlinear optical properties of AGGS were first reported by Petrov et al.10 in 2004. Since then, several teams have reported encouraging results concerning the utilization of AGGS in frequency mixing processes such as second harmonic generation (SHG) or difference frequency generation (DFG).11 However, difficulties relative to the growth of good optical quality AGGS single crystals were pointed out,12 and dissimilar results were reported regarding the phase stability of the AgGaS2−GeS2 system. Indeed, Badikov et al.8 indicated that AGGS has a stability range included between 46 and 83 mol % of GeS2 with a 845 °C congruent melting point. On the contrary, Chbani et al.13 reported that the AGGS phase has a fixed composition and an incongruent melting point at 840 °C. Lastly, Olekseyuk et al.14 indicated that the AGGS phase, with variable compositions, exists in the 48−55 mol % GeS2, and they assumed that previously published results were incorrect due to measurements performed on nonequilibrated alloys. These contradictory results are due to the volatility of AGGS derivative compounds and particularly to the germanium sulfide (GeS2)

INTRODUCTION Mid-infrared laser sources have attracted particular attention due to their potential applications in different fields, such as infrared counter-measures and remote chemical sensing.1 One way to produce such sources is the down conversion process in optical parametric oscillators based on birefringence phasematching in nonlinear crystals. The large majority of infrared laser sources based on the OPO technology are complex and bulky systems such as tandem OPO schemes where a first OPO is used to pump the mid-IR OPO.2 Consequently, it would be interesting to convert efficiently the radiation of a commercial 1.064 μm Nd:YAG laser source, with good beam quality and high power, to develop compact and tunable high power midIR laser sources. The development of such systems requires the use of nonlinear crystals that can directly be pumped at 1.064 μm. Nonlinear materials such as periodically poled LiNbO3 (PPLN)3 or periodically poled KTiOPO4 (PPKTP),4 which have been extensively studied the last decades, are compatible with that kind of pumping, but their transparency is limited to 4−5 μm due to intrinsic multiphonon absorption. Consequently, non-oxide materials, phosphides, or sulfides, for example, have been used in order to go beyond this limit. In this regard, CdSiP2 has been recently developed and has exhibited interesting properties such as a high thermal conductivity (13.6 W·mK−1) and a high nonlinear coefficient (84.5 pm·V−1).5 However, although this material is compatible with a 1 or 2 μm pumping, intrinsic multiphonon absorption limits its transparency beyond 6.5 μm. Thus, we have been interested in AgGaGeS4, which is also compatible with a 1 μm pumping but has a broader transparency in the mid-IR range. Indeed, this material has a wide transmission range © 2014 American Chemical Society

Received: June 4, 2014 Revised: September 10, 2014 Published: September 26, 2014 5554

dx.doi.org/10.1021/cg500813q | Cryst. Growth Des. 2014, 14, 5554−5560

Crystal Growth & Design

Article

vaporization, which can induce stoichiometry deviations in the melt. Thus, the control of AGGS processing steps, taking into account the vaporization of GeS2, appears to be a key issue for the growth of good optical quality single crystals. After presenting the experimental conditions to obtain high quality AGGS single crystals, we are going to discuss all the different steps explaining the impact of the volatile compounds on the crystal growth and the optimization of the processing steps.



EXPERIMENTAL SECTION

Chemical synthesis of AgGaGeS4. AgGaGeS4 polycrystals were synthesized from high purity elements Ag, Ga, Ge, and S (5 or 6 N). A stoichiometric amount of the starting elements, for a total amount of 120 g, was introduced into a quartz ampule under argon atmosphere. The quartz reactor was evacuated to 10−4 mbar and sealed. The ampule was placed into the synthesis furnace, tilted by 15° relative to the horizontal, which was confined into a stainless steal autoclave. The synthesis process was carried out using the two zones temperature method. This method allows performing the reaction between the different components at high temperature while limiting the explosion hazards related to the high vapor pressure of sulfur components. The measured temperatures in the lower and upper zones, during the synthesis process, are presented in Figure 1. In the first step,

Figure 2. AgGaGeS4 polycrystalline ingot after the first step: chemical synthesis.

Figure 3. Schemes of quartz ampules designed for growing crystal from seed with (a) and without (b) crucible. Figure 1. Plot of measured temperatures in the lower and upper zones during the chemical synthesis process.

pyrolytic boron nitride crucible (Figure 3(a)). This kind of ampule was previously successfully used in our laboratory for the crystal growth of others materials, such as ZnGeP2 and AgGaS2.1,15 Experiments were also performed using a quartz ampule with a specially designed bottom, which allowed crystal growth without a crucible in order to minimize the free volume in the ampule (Figure 3(b)). A single crystal seed, extracted from a large grain of the same polycrystal, was placed at the bottom of the quartz growth ampules. The AGGS polycrystals were ground into powder and introduced above the single crystal seed. A quartz cap was used in order to reduce, as much as possible, the free volume above the ground sample. The ampule was evacuated to 10−4 mbar for 1 h, sealed, and loaded into the furnace. The single crystal growth was performed using the Bridgman−Stockbarger method. The Bridgman−Stockbarger oven is a vertical tube furnace (Cyberstar Inc.). Temperatures of each zone were increased at the rate of 30 °C· h−1. The upper and lower zones were respectively raised up to 895 and 835 °C, leading to a temperature gradient of 12 °C·cm−1 at the AGGS melting point (845 °C).8 At the beginning, the ampule was positioned to start the crystallization from the seed. After a 5 h homogenization step, the ampule was pulled down at a rate of 0.4 mm·h−1. At the end, the temperature of the furnace was cooled at 10 °C·h−1 down to 785 °C.h−1 and then at a 30 °C·h−1 rate down to room temperature. The

temperature was raised to 950 °C in the lower zone, where the heterogeneous chemical reaction takes place between liquid (Ag, Ga, Ge, GeS2) and gas (S, GeS2) components. Volatile compounds condense in the upper zone, where the temperature is around 450 °C. This reduces the rising pressure. Once the major part of volatiles have reacted (∼15 h), the temperature was slowly raised in the upper zone, up to 1000 °C, in order to consume the remaining condensed sulfide compounds. It is interesting to notice that the measured temperature in the upper zone decreases during the consumption of the condensed compounds (Figure 1). This is due to endothermic melting and vaporization reactions. After a 15 h homogenization step, the entire ampule was cooled down to 600 °C at 0.2 °C·min−1 and then to room temperature at 5 °C·min−1. Polycrystalline ingot, visible in Figure 2, was collected, breaking the quartz ampule. Single crystal growth of AgGaGeS4. The crystal growth was performed from previously synthesized AGGS polycrystals using two types of ampules. Experiments were carried out with a round-bottom shaped ampule, where the AGGS polycrystals were introduced into a 5555

dx.doi.org/10.1021/cg500813q | Cryst. Growth Des. 2014, 14, 5554−5560

Crystal Growth & Design

Article

as-grown single crystal, orange in color, was 28 mm in diameter and 70 mm in length. It was transparent with the presence of color inhomogeneities. Samples extracted from the ingot are shown in Figure 4.

Then, about 100 mg of samples was introduced into small quartz crucibles, which were evacuated and sealed. These sealed crucibles avoid the vaporization losses of volatile compounds during the experiments. The optical properties were studied using a Nicolet 5700 FT-IR spectrometer in the 1.6−14 μm range and a cary 6000 (Varian Inc.) spectrometer for the visible and near-infrared range (0.4−1.7 μm). The samples were polished before measurements. Absorption coefficients were calculated referring to the Sellmeier coefficients provided by Das et al.16 Thermal conductivity measurements were carried out at room temperature using a C-Term TCI (Setaram Inc.) thermal conductivity analyzer. The apparatus is based on the modified transient plane source (MPTS) technique, which employs a one-sided, interfacial heat reflectance sensor that applies a momentary constant heat source to the sample, which induces a small increase of the temperature on the crystal (typically less than 2 °C). Thermal conductivity and effusivity were measured directly, providing a detailed overview of the thermal characteristics of the sample. Calibrated samples were also tested with the same setup. Study of GeS2 vapor pressure variation over the AgGaGeS4 melt was simulated using the HSC Chemistry software.

Figure 4. Picture of single crystals extracted from the AgGaGeS4 ingot.



Postgrowth treatments. A post-growth treatment step was performed in order to improve the quality of the as-grown crystal. Thermal annealing was performed using the apparatus previously described.15 Extracted samples from as-grown ingot (Figure 5(a)) were introduced into a quartz ampule which was evacuated to ∼1 mbar and capped using a ground glass stopper. The contact between the stopper and the ampule was efficient enough to be pressure-tight. The annealing was carried out under static vacuum, which was the vacuum initially imposed in the ampule, at 600 °C for 250 h. After this step, the samples had an homogeneous yellow color (Figure 5(b)) and were transparent with no visible presence of defects. Characterizations. X-ray powder diffraction was performed using a Panalytical X’Pert Pro diffractometer operating at Cu Kα (1.54182 Å) wavelength and equipped with a front beam Ge (1 1 1) monochromator and a linear PixCell detector. The samples were ground into powder before measurements, and the patterns were collected at room temperature in the 2θ range of 10−80° with a 0.05°· s−1 scanning rate. Sample compositions and microstructures were determined using a ZEISS DSM962 scanning electron microscope (SEM), and energy dispersive X-ray spectroscopy (EDS) was performed at 15 kV without reference samples. Differential thermal analysis (DTA) measurements were carried out using a Setsys evolution 16/18 (Setaram Inc.) apparatus. Pure silver, aluminum, and copper were used for the DTA temperature calibration.

RESULTS AND DISCUSSION Chemical synthesis. The AgGaGeS4 phase is formed in the solid solution AgGaS2−GeS2. This compound has a narrow stability range comprising between 48 and 55 GeS2 mol % compositions according to the latest published phase diagram,14 to which we will refer in the following discussion. Indeed, compositions below 48% GeS2 mol % or above 55% GeS2 mol %, respectively, lead to the formation of the mixed AgGaGeS4 + AgGaS2 or AgGaGeS4 + GeS2 phases. Otherwise, AGGS decomposes at high temperature to form AgGaS2 and GeS2 phases according to the following reaction:17 AgGaGeS4(S,L) ⇌ AgGaS2(S,L) + GeS2(L,V) (1) Moreover, above 627 °C, GeS2 dissociates in the vapor phase to form GeS and Sx. However, to simplify the discussion, we will refer to GeS2 rather than GeS and Sx species.17 In order to study the GeS2 vapor pressure behavior with increasing temperature, thermodynamic calculations were performed. The results of these simulations are reported in Figure 6 and are compared to experimental data extrapolated from Vasil’eva et al.17 They are in good agreement with experimental results and indicated that GeS2 vapor pressure increases quickly above

Figure 5. Picture of a polished AgGaGeS4 single crystal before (a) and after post-treatment (b). 5556

dx.doi.org/10.1021/cg500813q | Cryst. Growth Des. 2014, 14, 5554−5560

Crystal Growth & Design

Article

GeS2. Consequently, our work focused on reducing the effect of the GeS2 vaporization on melt stoichiometry in order to prepare high quality AGGS polycrystals. To do that, parameters such as ampule volume, starting elements amount, and synthesis temperature have to be optimized. First, in this case, since vapor pressure only depends on temperature, the ratio between the ampule volume and the amount of starting elements can be optimized within two directions: (i) either to limit the volume of the ampule or (ii) to increase the reactants quantity. Thus, for a given ampule volume, a larger quantity of starting elements reduces the impact of volatility on melt composition. Indeed, the calculated vapor pressure of GeS2 is 0.7 bar at 845 °C. If we consider 120 g of starting elements and a 190 cm3 ampule, the GeS2 existing in the vapor phase corresponds to 0.2 g of GeS2, representing 0.5 mol % of the total GeS2 mass. Thus, under these conditions, the melting composition is maintained in the stability range of the AgGaGeS4 phase.14 However, if we decrease the amount of starting elements, the mass of GeS2 in the vapor phase will represent a higher ratio and the melt composition can be shifted outside the stability zone of AgGaGeS4. This is a reason why large nonlinear crystals of volatile compounds usually present higher quality than smaller crystals.18 For example, under the same conditions, considering 25 g of starting elements, the GeS2 existing in the vapor phase represents 2.2 mol % of the total GeS2 mass and the composition is shifted outside the AgGaGeS4 stability zone. This was indeed observed at the beginning of our work and could occur when working with small quantities on new materials processing. However, the ratio between the ampule volume and the starting elements amount has to be optimized taking into account the explosions hazards related to the amount of sulfur. For the moment, we have tested 120 g of compounds inside a 190 cm3 ampule, which was sufficient to maintain the vaporization of GeS2 at a low level. Otherwise, the cooling rate is another important parameter to take into consideration during the process. Indeed, at 950 and 845 °C, the vapor phase contains respectively 2.0 and 0.5 mol % GeS2. Thus, a fast cooling rate can induced an important stoichiometric deviation in the final crystal if the GeS2 vapor does not have enough time to react with the melt before crystallization. Consequently, the kinetics of the reaction has an important role here. For example, our experiments, carried out with a cooling rate of 5 °C·min−1 at the end of the synthesis, led to ingots composed by AgGaGeS4 + AgGaS2 phases and unreacted GeS2 at their top, as shown in Figure 8. A cooling rate of 0.2 °C·min−1 was sufficient to promote vapor/melt

Figure 6. Plot of simulated GeS2 vapor pressure variation (solid red line) versus temperature and comparison with experimental data from ref 17 (dots).

700 °C. Thus, the vapor pressure of GeS2 is about 0.7 bar around 845 °C, at the melting point temperature of AgGaGeS4 reported by Badikov et al.,8 and about 5 bar at 1000 °C, which is the highest chemical synthesis temperature in our process. This vaporization can induce a deviation from the initial composition depending on the synthesis conditions, the amount of starting elements, and the reactor volume. Thus, if the GeS2 vaporization leads to a melt composition below 48 mol % GeS2, inclusion of AgGaS2 appears in the materials, which became opaque and unusable for our applications, as shown in Figure 7. Indeed, this SEM picture of our first synthesis shows an eutectic-like microstructure composed of gray and white phases. These phases were respectively identified as AgGaGeS4 and AgGaS2, and the overall calculated composition was about 47 mol % GeS2 according to EDS analysis. Thus, the synthesis of a pure AGGS phase is a tricky process which requires taking into account the vaporization of

Figure 7. SEM picture of a crystal’s cut showing the eutectic-like microstructure composed by a mixture of AgGaGeS4 and AgGaS2 phases. The overall composition is around 47 mol % GeS2 according to EDS analysis.

Figure 8. Picture of the ingot after chemical synthesis performed with a 5 °C·min−1 cooling rate. 5557

dx.doi.org/10.1021/cg500813q | Cryst. Growth Des. 2014, 14, 5554−5560

Crystal Growth & Design

Article

reaction during cooling and totally consume the GeS2 content in the vapor phase. Thus, the optimization of these parameters (cooling rate, ratio ampule volume/mass of charge) allowed better control of GeS2 vaporization and the processing of polycrystals (Figure 2) with a pure AGGS orthorhombic phase (JCPDS 72-19127) according to XRD analysis (Figure 9).

Figure 10. DTA analysis showing the melting and crystallization temperatures of AgGeGeS4.

responsible for a multinuclei formation observed on samples grown from the melt without seed.19 We supposed that the quasi-entire charge was cooled below the supercooling crystallization point during growth, which could explain the obtained results. Moreover, this supercooling can induce a decomposition of the initial AGGS phase, as was recently reported by Nikolaev et al.20 The use of single crystal seed avoided the supercooling effect by providing a starting nucleation point and allowed growing AGGS single crystals. An extracted sample from a AgGaGeS4 single crystal is shown in Figure 5(a). SEM-EDS analyses were performed on different spots of that crystal after annealing. The results indicated the good homogeneity of the single crystal with an estimated composition similar to the expected composition in the range of 3 at %, which is the precision limit of the EDS analysis (according to the previously described conditions). Optical measurements were performed on polished and annealed AGGS samples (10 × 10 × 5 mm3). The optical transmission was measured at 2.05 μm using a thulium doped fiber cw laser21 (20 Khz, 100−400 ns, peak power 1 kW) and at 1.064 μm using a Nd:YAG laser. These measurements indicated an absorption coefficient of 0.04 cm−1 at 2.05 μm and 0.12 cm−1 at 1.064 μm. The absorption spectra recorded in the 0.5−12.5 μm range (Figure 11) shows that AGGS samples have high transparency from about 0.5 to 11.5 cm−1. The absorption coefficient of the annealed sample was below ∼0.1 cm−1 in the 0.9−8 μm spectral range, which indicates the good optical quality of our AGGS crystal. The cutoff wavelength is 0.45 μm, which perfectly corresponds to the 2.78 eV band edge reported by Badikov et al.8 Otherwise, the transmission of the as-grown sample indicates a short cutoff wavelength of 0.49 μm (2.53 eV). Thus, the as-grown crystal cutoff is shifted to longer wavelengths due to absorption losses, which clearly explains the crystal’s orange color. The as-grown sample has sharp absorption peaks located at 2.9, 4, and 10 μm which may be attributed to water contaminations (H−S and −OH bonds).226 These absorptions peaks were eliminated by the annealing treatment as shown in Figure 11. The measured thermal conductivity value of the unoriented as-grown sample was 0.4 W·m−1·K−1, which is in good agreement with the 0.399 ± 0.002 W·m−1·K−1 value reported by Schunemann et al.6 Then, we measured thermal conductivity on the same sample after annealing. The thermal conductivity value was 10% higher than in the as-grown sample (0.44 W·

Figure 9. XRD pattern of a polycrystalline sample matching the AgGaGeS4 phase according to JCPDS 72-1912.7

Crystal growth. The crystal growth step is also sensitive to GeS2 vaporization, which has a strong impact on the final single crystal quality. Indeed, the ratio between starting elements and ampule volume has to be optimized in order to avoid stoichiometry deviation during the crystallization process. For example, use of a specially shaped ampule (Figure 3(a)) is preferable than using a round-bottom shaped ampule and a pyrolytic boron nitride or graphite crucible (Figure 3(b)). Indeed, experiments performed with this latter led to condensation of GeS2 in the coldest part, at the ampule bottom. This condensation occurred between the crucible and the ampule walls, which led to stoichiometry deviation in the melt and the growth of poor quality crystals. Another important parameter to take into consideration is the supercooling of AGGS. Indeed, growths from the melt without seed were performed with the two types of ampules (Figure 3). These experiments led to ingots composed of a mixture of AgGaGeS4 and AgGaS2 phases in the major part. Only the top of the crystals were composed of a pure AgGaGeS4 polycrystalline phase. Thus, DTA analysis was carried out in order to explain these results and to study the supercooling amplitude. The experiments were performed at different heating and cooling rates, 1 °C·min−1, 3 °C·min−1, 5 °C·min−1, and 10 °C·min−1. The melting and crystallization points are respectively shown on the heating and cooling curves presented in Figure 10 (experiment performed with a 5 °C· min−1 heating/cooling rate). The melting point, related to the endothermic peak on the heating curve, is at about 838 °C. The crystallization temperature, related to the exothermic peak on the cooling curve, is at about 752 °C. Thus, the supercooling temperature value reaches 86 °C under these conditions and was similar in the experiments performed with other heating/ cooling rates. Even if this value strongly depends on different parameters, such as the container quality, and could be slightly different under real crystallization conditions, one can say that the supercooling is relatively high. Consequently, it should be 5558

dx.doi.org/10.1021/cg500813q | Cryst. Growth Des. 2014, 14, 5554−5560

Crystal Growth & Design

Article

2010; pp 783811−783811-6. (b) Antoine, G. C. R. Phys. 2007, 8, 1100−1128. (c) Petrov, V. Opt. Mater. 2012, 34, 536−554. (2) (a) Raffy, J.; Debuisschert, T.; Pocholle, J.-P.; Papuchon, M. Appl. Opt. 1994, 33, 985. (b) Vaicikauskas, V.; Kaucikas, M.; Svedas, V.; Kuprionis, Z. Rev. Sci. Instrum. 2007, 78, 023106 (PMID: 17578102). (3) (a) Myers, L. E.; Eckardt, R. C.; Fejer, M. M.; Byer, R. L.; Bosenberg, W. R.; Pierce, J. W. J. Opt. Soc. Am. B 1995, 12, 2102. (b) Yamada, M.; Nada, N.; Saitoh, M.; Watanabe, K. Appl. Phys. Lett. 1993, 62, 435. (c) Burns, W.; McElhanon, W.; Goldberg, L. IEEE Photonics Technol. Lett. 1994, 6, 252−254. (4) (a) Fradkin, K.; Arie, A.; Skliar, A.; Rosenman, G. Appl. Phys. Lett. 1999, 74, 914. (b) Fiorentino, M.; Spillane, S. M.; Beausoleil, R. G.; Roberts, T. D.; Battle, P.; Munro, M. W. Opt. Express 2007, 15, 7479. (5) (a) Zawilski, K. T.; Schunemann, P. G.; Pollak, T. C.; Zelmon, D. E.; Fernelius, N. C.; Kenneth Hopkins, F. J. Cryst. Growth 2010, 312, 1127−1132. (b) Buehler, E.; Wernick, J. J. Cryst. Growth 1971, 8, 324−332. (c) Mughal, S. A.; Payne, A. J.; Ray, B. J. Mater. Sci. 1969, 4, 895−901. (d) Marchev, G.; Tyazhev, A.; Petrov, V.; Schunemann, P. G.; Zawilski, K. T.; Stoppler, G.; Eichhorn, M. Opt. Lett. 2012, 37, 740−742. (e) Kumar, S. C.; Jelı ́nek, M.; Baudisch, M.; Zawilski, K. T.; Schunemann, P. G.; Kubecek, V.; Biegert, J.; Ebrahim-Zadeh, M. Opt. Express 2012, 20, 15703−15709. (f) Hildenbrand, A.; Kieleck, C.; Tyazhev, A.; Marchev, G.; Stoppler, G.; Eichhorn, M.; Schunemann, P.; Petrov, V. Laser damage studies of CdSiP2 and ZnGeP2 nonlinear crystals with nanosecond pulses at 1064 and 2090 nm. Proc. SPIE 8964, Nonlinear Frequency Generation and Conversion: Materials, Devices, and Applications XIII, 896417 (February 20, 2014), p 896417; doi: 10.1117/12.2039050. (6) (a) Schunemann, P. G.; Zawilski, K. T.; Pollak, T. M. J. Cryst. Growth 2006, 287, 248−251. (b) Petrov, V.; Badikov, V.; Panyutin, V. In Mid-Infrared Coherent Sources and Applications; Ebrahim-Zadeh, M., Sorokina, I. T., Eds.; Springer Netherlands: Dordrecht, 2008; pp 105− 147. (7) Pobedimskaya, E. Sov. Phys. Dokl. 1981, 26, 259. (8) Badikov, V.; Tyulyupa, A.; Shevyrdyaeva, G.; Sheina, S. Inorg. Mater. 1991, 27, 177−180. (9) Andreev, Y.; Geiko, P. P.; Badikov, V. V.; Bhar, G. C.; Das, S.; Chaudhury, A. K. Nonlinear Opt. (Mclc) Sect. B 2002, 29, 19−27. (10) Petrov, V.; Badikov, V.; Shevyrdyaeva, G.; Panyutin, V.; Chizhikov, V. Opt. Mater. 2004, 26, 217−222. (11) (a) De-Ming, R.; Jin-Zhe, H.; Yan-Chen, Q.; Xiao-Yong, H.; Yuri, A.; Pavel, G.; Valerii, B.; Anna, S. Chin. Phys. 2004, 13, 1468− 1473. (b) Badikov, V.; Shevyrdyaeva, G.; Chizhikov, V.; Panyutin, V.; Xu, G.; Petrov, V.; Noack, F. Appl. Phys. Lett. 2005, 87, 24111− 241113-3. (c) Miyata, K.; Petrov, V.; Kato, K. Appl. Opt. 2007, 46, 6848−6848. (d) Ni, Y.; Wu, H.; Wang, Z.; Mao, M.; Cheng, G.; Fei, H. J. Cryst. Growth 2009, 311, 1404−1406. (12) Wu, H.; Ni, Y.; Lin, C.; Mao, M.; Cheng, G.; Wang, Z. Front. Optoelectron. China 2011, 4, 137−140. (13) Chbani, N.; Loireau-Lozac’h, A.-M.; Rivet, J.; Dugue, J. J. Solid State Chem. 1995, 117, 189−200. (14) Olekseyuk, I.; Gorgut, G.; Shevtchuk, M. Polym. J. Chem. 2002, 76, 915−919 (WOS:000176390600002). (15) Petit, J.; Bejet, M.; Daux, J.-C. Mater. Chem. Phys. 2010, 119, 1− 3. (16) Das, S.; Ghosh, C.; Gangopadhyay, S.; Andreev, Y. M.; Badikov, V. V. Jpn. J. Appl. Phys. 2006, 45, 5795−5797. (17) Vasil’eva, I. G.; Nikolaev, R. E. Inorg. Mater. 2006, 42, 1299− 1301. (18) (a) Zawilski, K. T.; Schunemann, P. G.; Setzler, S. D.; Pollak, T. M. J. Cryst. Growth 2008, 310, 1891−1896. (b) Verozubova, G.; Okunev, A.; Gribenyukov, A.; Trofimiv, A.; Trukhanov, E.; Kolesnikov, A. J. Cryst. Growth 2010, 312, 1122−1126. (c) Lei, Z.; Zhu, C.; Xu, C.; Yao, B.; Yang, C. J. Cryst. Growth 2014, 389, 23−29. (19) Chen, B.; Zhu, S.; Zhao, B.; Lei, Y.; Wu, X.; Yuan, Z.; He, Z. J. Cryst. Growth 2008, 310, 635−638. (20) Nikolaev, R.; Vasilyeva, I. J. Solid State Chem. 2013, 203, 340− 344.

Figure 11. Optical absorption coefficient of an AGGS sample (thickness: 6 mm) in the 0.5−12 μm range from spectrophotometric measurements (solid curves) and data measured with a thulium doped fiber laser (▲) and a Nd:YAG laser (★).

m−1·K−1), which is related to decreasing of the crystal defect amount.



CONCLUSIONS The strong influence of GeS2 vaporization on AgGaGeS4 melt stoichiometry and crystal quality was outlined in this study. Solutions to reduce the impact of this compound’s volatility on stoichiometry were presented. With optimized experimental conditions, we have synthesized high quality AgGaGeS4 polycrystals and have grown good optical quality single crystals by the Bridgman−Stockbarger method. These crystals had good homogeneity, and the absorption loss coefficient was below ∼0.1 cm−1 in the 0.9−8 μm spectral range, making them suitable for nonlinear optical applications. Otherwise, experiments are currently in progress in order to improve the quality of these crystals by optimizing growth rate, seed orientation, and annealing parameters. We believe that the present work will be helpful for the processing of such nonlinear infrared materials which may experience the same volatility problems. Laser experiments are currently in progress.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors thank Loiseau P. from IRCP Chimie ParisTech for allowing us to collect X-ray data and for his valuable advice. We also want to thank Ritt M. H. and Jankowiak A. from the Departement of composite materials of the ONERA for their guidances on DTA measurements.



REFERENCES

(1) (a) Petit, J.; Godard, A.; Raybaut, M.; Melkonian, J.-M.; Lefebvre, M. Mid-IR nonlinear materials: chemical synthesis, crystal growth, and difference frequency generation in ZnGeP2. Proc. SPIE 7838, Optics and Photonics for Counterterrorism and Crime Fighting VI and Optical Materials in Defence Systems Technology VII, Toulouse, France, 5559

dx.doi.org/10.1021/cg500813q | Cryst. Growth Des. 2014, 14, 5554−5560

Crystal Growth & Design

Article

(21) Lucas, E. M. D.; Canat, G.; Lombard, L.; Jaouën, Y.; Bordais, S. Pulsed kW-peak power and integrated all fiber MOPA singlefrequency source at 2050nm. Proc. SPIE 9135, Laser Sources and Applications II. Brussels, Belgium, 2014; pp 91350M−91350M-7. (22) Young, P. A. J. Phys. C: Solid State Phys. 1971, 4, 93−106.

5560

dx.doi.org/10.1021/cg500813q | Cryst. Growth Des. 2014, 14, 5554−5560