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

Broadband studies of the strong mid-infrared nonlinear optical responses of KPSe6

Open Access Open Access

Abstract

One-dimensional noncentrosymmetric KPSe6 is an excellent mid-infrared (mid-IR) nonlinear optical (NLO) material possessing reversible phase-change behavior. In this special issue we present a broadband NLO study on its crystalline and glassy phases as well as fundamental material characterizations. The compound has a bandgap of ∼ 2.1 eV and exhibits intrinsic second harmonic generation even in the glassy phase. The crystalline compound is type-I phase-matchable and shows strong second- and third-order NLO responses over a broad wavelength range (1.0 – 3.0 μm) with excellent optical transparency. Based on the measured NLO coefficients χ(2) ≃ 142.8 pm/V and χ(3) ≃ 4.7 × 105 pm2/V2, we propose that KPSe6 can be utilized for numerous mid-IR NLO applications.

© 2013 Optical Society of America

1. Introduction

In the emerging age of mid-infrared (mid-IR) science and technology much interest has centered on finding materials that have suitable nonlinear optical (NLO) properties for use in diverse applications [16]. Several classes of materials for this goal include organic- and oxide-based systems [710]. On the other hand, a current consensus points to chalcogenides as the best choice of materials owing to their typical properties such as high refractive indices and optical transparency [11]. Numerous noncentrosymmetric chalcogenide crystals have been investigated for tunable coherent mid-IR generation involving the second-order NLO processes [1216]. Amorphous chalcogenide materials are also important for mid-IR optical switching and signal amplification, which are based on third-order NLO processes [17, 18].

Of particular interest is the polar alkali chalcogenide compound KPSe6 that crystallizes in a noncentrosymmetric one-dimensional (1-D) polymeric structure (see Fig. 1). KPSe6 has a remarkable set of properties in a single material; it i) undergoes a reversible crystal-glass phase transition, ii) exhibits significant second-order NLO responses in both crystalline and amorphous phases, iii) is soluble in organic solvents and thus processible into thin films, and iv) melts congruently to give honey-like viscous melts from which long fibers can be drawn [16, 19, 20]. We recently demonstrated that this multifaceted feature of the material could lead to a new design principle of highly-efficient NLO devices of fiber and film forms [19, 20]. Previous NLO studies on this unique compound, however, were focused on the second-order processes only within a limited fundamental wavelength λ in the range of 1.2 – 1.6 μm. Therefore, the fundamental NLO aspect and the potential of the compound in a broader mid-IR range were far from fully understood. Here we report on strong second- and third-order NLO properties of both crystalline and glassy KPSe6 powders over an extended spectral range (λ = 1.0 – 3.0 μm). We note that broadband THG studies on newly discovered chalcogenides are rare and our wavelength-dependent THG study offers a unique opportunity to evaluate the mid-IR third-order nonlinearity of KPSe6.

 figure: Fig. 1

Fig. 1 Crystal structure of KPSe6 viewed down the b-axis.

Download Full Size | PDF

This paper is organized as follows: Section. 2 describes KPSe6 sample preparation followed by several key characterizations of the compound. The experimental method for the powder NLO measurements is described in Sec. 3. In Sec. 4 we characterize the NLO properties of KPSe6 based on broadband second harmonic generation (SHG) and third harmonic generation (THG). Section 5 gives our summary and conclusions.

2. Sample preparation and characterization

For the synthesis of crystalline KPSe6, stoichiometric amounts of K2Se (144 mg, 0.919 mmol), P (57.0 mg, 1.84 mmol), and Se (799 mg, 10.1 mmol) were placed in a fused silica tube (9 mm in diameter) inside a nitrogen-filled glovebox. The tube was evacuated to ∼ 3×10−4 mbar then flame sealed. The sealed tube was about 15 cm long, and the starting materials occupied about 1/6 of the total tube volume. The tube was placed in a programmable furnace at an angle 30° from the floor then heated to 460 C in 5 hours and dwelled at this temperature for 12 hours to form a homogeneous melt. The sample was cooled to 210 C in 8 hours, heated to 280 C over the course of 2 days to crystallize the product, which annealed at 280 C for 36 hours then cooled to room temperature in 6 hours. This novel heating profile was employed to form highly crystalline product of KPSe6, which uniquely only crystallizes on heating, but the general principles of this synthesis follow already well-established general procedures for crystal growth [2123]. The same stoichiometric mixture in a flame-sealed tube of the same proportions was prepared for glassy KPSe6 but the sample was heated to 800 C over the course of 8 hours and was immediately quenched in ice water. The final product was an ingot for both syntheses. The apparent colors of crystalline and glassy KPSe6 were orange and red, respectively. We confirmed the phase purity of the samples using powder X-ray diffraction (PXRD) [see Fig. 2(a)]. The material is air and moisture sensitive and all samples were stored in an evacuated desiccator to prevent decomposition.

 figure: Fig. 2

Fig. 2 (a) Experimental PXRD patterns of crystalline KPSe6 (blue) and glassy KPSe6 (red) compared to the simulated pattern (black). Each trace is vertically translated for clarity. (b) Solid-state UV-VIS absorption spectra of crystalline (blue) and amorphous (red) KPSe6 superimposed by the theoretical fits (solid lines).

Download Full Size | PDF

KPSe6 crystallizes in the noncentrosymmetric polar orthorhombic space group Pca21 with a = 11.491(12) Å, b = 6.791(7) Å, c = 11.289(12) Å, and Z = 4 at room temperature [24]. It is composed of infinite 1-D chains of 1[PSe6] along the a-axis and separated by K+ cation as schematically illustrated in Fig. 1. Short-range intra- and inter-layer Se-Se interactions play an important role in the crystal structure together with alkali-metal ionic interactions between the chains. The corresponding band structure is shown in Ref. [25] in which the pivotal dependence of the second-order NLO tensor χ(2) on the bandgap energy, optical matrix elements, and joint density of states is also extensively discussed.

Figure 2(a) plots the PXRD patterns of crystalline KPSe6 (blue) and glassy KPSe6 (red), respectively, collected using a computer-controlled INEL CPS 120 powder diffractometer (Cu Kα radiation operating at 40 kV and 20 mA) equipped with a positive-sensitive detector with flat sample geometry. The simulated pattern (black) was generated using the CIF file of the refined structure and the Visualizer program within FindIt. Our PXRD analysis shows that crystalline KPSe6 matches the simulated pattern, although a minor amount of the glassy phase is present (slight background). PXRD measurements of glassy KPSe6 reveal no crystalline peaks, signifying the material is completely amorphous.

Optical diffuse reflective measurements were performed in the range of 0.2–2.0 μm at room temperature using a Shimadzu UV-3101 PC double-beam, double-monochromator spectrophotometer. BaSO4 was used as the standard and set to 100% reflectance. Samples were prepared by pressing well-ground powders on a surface of compressed BaSO4. The collected reflectance data were converted to absorbance via the Kubelka-Munk equation: α/S = (1 − R)2/2R, where R is reflectance, α is the absorption coefficient, and S is the scattering coefficient [26]. The converted data are plotted in Fig. 2(b) and used to estimate the fundamental absorption edges by linearly fitting the absorbance spectra. Analysis of these data shows that crystalline and glassy KPSe6 have bandgaps of 2.11 eV (0.587 μm) and 2.03 eV (0.610 μm), respectively. The tail end in absorbance of crystalline KPSe6 is due to the presence of some of the glassy phase.

The thermal properties of crystalline KPSe6 were investigated by differential thermal analysis (DTA) performed on a Shimadzu DTA-50 thermal analyzer. Ground crystalline samples (∼ 90 mg) were sealed under vacuum in a fused silica ampoule. A similar mass of α-Al2O3 sealed in a separate ampoule was used as the standard. The sample and reference were heated and cooled at rate of 5 C/min, with a maximum temperature of 400 C. The crystallization and melting points were measured at the maximum of the exothermic peak and the minimum of the endothermic peak, respectively. Figure 3(a) shows that KPSe6 only crystallizes on heating at 261 C and melts at 300 C. No exothermic peak was observed in the first DTA cycle because the material was already crystalline. The reproducible exotherm in the second and third cycles demonstrates that the recrystallization of KPSe6 is reversible. The DTA residue was examined by PXRD and reveals amorphous KPSe6 because this compound only crystallizes on heating.

 figure: Fig. 3

Fig. 3 (a) DTA of crystalline KPSe6 from 50 C to 400 C, displaying three cycles to demonstrate its reversible crystal-glass phase-change properties; first cycle (black), second cycle (red), and third cycle (blue). (b) Raman spectra of crystalline (blue) and glassy (red) KPSe6. Each trace is vertically translated for clarity.

Download Full Size | PDF

Raman spectra were recorded from 100 – 1000 cm−1 on a DeltaNU Advantage NIR spectrometer equipped with a charge-coupled device (CCD) camera using 0.785 μm radiation from a diode laser [see Fig. 3(b)]. The samples were ground into powders and loaded into borosilicate glass capillaries for measurement. The spectra were collected using an integration time of 5 s. The Raman spectrum of crystalline KPSe6 (blue) shows major shifts at 220 cm−1, 228 cm−1, and 245 cm−1. The shift at 220 cm−1 corresponds to the PSe4 stretching mode determined by comparison with the Ag stretching mode of the Td symmetry of the [PSe4]3− ligand. The shifts at 228 cm−1 and 243 cm−1 are assigned to the antisymmetric and symmetric Se-Se stretching modes of the diselenide group, respectively. The Raman spectrum of glassy KPSe6 (red) shows major shifts at 219 cm−1 and 252 cm−1. The overall pattern of the glassy Raman spectrum is similar to crystalline KPSe6 although slightly broader. Generally, the Raman spectra of chalcogenide glasses are broad and featureless because of a larger variation in the structural building units in the glass [27, 28]. The fact that the overall Raman pattern of the glass is similar to the crystalline phase suggests that the local structure and the [PSe4] building block and Se-Se bonds are mostly preserved in the glassy phase [19, 29]. This essentially leads to an “intrinsic” SHG response with no poling treatment as described in Sec. 4.

3. Wavelength-dependent NLO experimental methods

Both crystalline and amorphous samples for NLO measurements in powder form were prepared under dry nitrogen conditions. The powders were sieved with size ranges of 20 – 32 μm, 32 – 45 μm, 45–53 μm, 53–63 μm, 63–90 μm, 90–106 μm, 106–125 μm, and 125–150 μm to examine the phase-matching behavior [30] of KPSe6. The specific particle sizes were then placed in borosilicate capillary tubes. Each tube was loaded into a homemade sample holder that was mounted on a Z-scan translation stage. All capillary tubes were capped in dry nitrogen environment then flame sealed to keep all materials from being exposed to air and moisture during the measurement.

Broadband SHG and THG experiments were conducted at room temperature. In order to generate tunable pulses, coherent light of wavelength 1.064 μm was first produced using an EKSPLA PL-2250 series diode-pumped picosecond Nd:YAG laser with a pulse width of 30 ps and a repetition rate of 50 Hz. The Nd:YAG laser pumped an EKSPLA Harmonics Unit (HU) H400 where the input beam can be frequency tripled by successive NLO wave mixing. The 1.064 μm and frequency-tripled radiation from the HU entered an EKSPLA PG403-SH-DFG Optical Parametric Oscillator (OPO) composed of four main parts; a double-pass parametric generator, a single-pass parametric amplifier, a second harmonic generator (SH), and a difference frequency generation (DFG) scheme. The output wavelengths of the OPO used in our experiments were from 1.0 μm to 3.0 μm at increments of 0.1 μm.

The incident pulse energy was tuned to 20 μJ before being mildly focused onto samples with a spot size of 0.5 mm in diameter by a convex lens or a concave gold mirror, far away from the Z-scan focus. Here we deliberately determined the beam spot size in order to i) properly average the NLO signals from powders of random orientations and ii) to minimize the change in the spot size when we swept λ over a broad range of 1.0–3.0 μm; the beam waist w0 at the Z-scan focus undergoes a significant λ-dependent variation via w0 = (λ/π)(f/σ), where f and σ are the focal length and the Gaussian width of the incident beam, respectively. For the fundamental wavelengths inaccessible with 20 μJ, the NLO counts were properly scaled in accordance with the measured SHG and THG power dependence. However, the data at λ = 2.2 μm were omitted because the pulse energy was too low to be measured by a Gentec energy meter. The NLO signals from the samples were collected using reflection geometry by a fiber-optic bundle, which was coupled to a selective-grating (1800, 600, and 300 grooves/mm) spectrometer equipped with a CCD camera (Synapse) as well as an extended InGaAs (Symphony) detector. The overall detection range obtained from the combination of the two detectors was 0.3–2.0 μm. We confirmed that any surface-induced effect as well as NLO signals from other optical components were negligible. Any thermal load on the samples by the laser pulses tuned below the bandgap was negligible due to its slow repetition rate of 50 Hz. The relative NLO signals recorded in a broad wavelength range were precisely calibrated with the known and measured efficiencies of all optical components.

4. Broadband NLO properties of KPSe6

Prior to observing SHG and THG from KPSe6, the NLO responses were measured from our reference materials, AgGaS2 and AgGaSe2, prepared by a similar method as described in Sec. 2. Figure 4(a) plots the series of SHG spectra as a function of SHG wavelength, λSHG = λ/2, from AgGaS2 for the particle size d in the range of 125 – 150 μm, when λ was varied from 1.0 μm to 3.0 μm. AgGaS2 is type-I phase-matchable in our experimental range as we also confirmed based on its particle-size dependence [19,31]. The SHG signals decrease when λSHG approaches the bandgap wavelength of AgGaS2 (λg ∼ 0.5 μm at room temperature) because of strong absorption of the SHG light and/or two-photon absorption of the fundamental light near the band edge. However, the mid-IR SHG response is essentially flat as λ approaches the static limit (λ → ∞) [25]. AgGaS2 was used to estimate the absolute value of χ(2) of crystalline KPSe6 as explained later.

 figure: Fig. 4

Fig. 4 Wavelength-dependent (a) SHG from AgGaS2 and (b) THG from AgGaSe2.

Download Full Size | PDF

Figure 4(b) corresponds to the broadband THG spectra from AgGaSe2 (d = 20 – 32 μm) obtained under the same experimental conditions. Bandgap absorption of THG is significant for λTHG = λ/3 < λg ∼ 0.73 μm (bandgap wavelength of AgGaSe2) but the THG response also converges into the static limit at mid-IR. The THG counts as a function of d indicated a non-phase-matching case, which is typical due to a large phase-mismatch between λ and λTHG[33]. Although AgGaSe2 is a good phase-matching SHG material for λ > 3 μm, the compound is not phase-matchable for λ used in our experiments [31, 32, 34]. Therefore, we used AgGaSe2 as a standard for the non-phase-matching SHG and THG cases.

Unlike conventional glassy materials, the amorphous phase of KPSe6 exhibits a substantial SHG response due to a partial preservation of the local crystallographic order as evidenced by Fig. 3(b) as well as the pair distribution function analysis [19]. Figure 5(a) plots the SHG counts as a function of d from the glassy compound at λ = 1.8 μm. The observed non-phase-matching behavior is indeed predicted for the macroscopically isotropic glassy phase due to lack of birefringence. The red dots in Fig. 5(b) correspond to the wavelength-dependent SHG counts from glassy KPSe6 (d = 20 – 32 μm) in our experimental range. Those from AgGaSe2 of the same particle size are also plotted for comparison (circles), yielding characteristic of bandgap absorption for shorter wavelengths and saturation to a static value for longer wavelengths except for a pronounced dip around λSHG ≃ 1.15 μm. The origin for the dip is twofold; i) linear absorption of the fundamental beam by the borosilicate tube [35] and ii) three-photon absorption of the fundamental beam by AgGaSe2, which all in turn decrease the SHG efficiency observed. The latter process is viable considering that the corresponding λ is approximately resonant with the three-photon edge of AgGaSe2 (3λg ≃ 2.2 μm). The former effect was seen from all of our SHG and THG measurements [see Figs. 6(b) and 6(c) near λ = 2.1 – 2.4 μm], but this does not affect our estimation of the NLO coefficients as explained later.

 figure: Fig. 5

Fig. 5 (a) Non-phase-matching SHG from glassy KPSe6 based on particle-size dependence at λ = 1.8 μm. Broadband (b) SHG counts and (c) THG counts from glassy KPSe6 (red) and AgGaSe2 (circles), respectively, plotted on a semi-log scale.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 (a) Phase-matching SHG from crystalline KPSe6 based on particle-size dependence at λ = 1.8 μm. (b) Broadband SHG counts from crystalline KPSe6 (blue) and AgGaS2 (black) at d = 125 – 150 μm. (c) Broadband THG counts from crystalline KPSe6 (blue) and AgGaSe2 (circles) at d = 20 – 32 μm.

Download Full Size | PDF

It is noteworthy that even glassy KPSe6 shows a good SHG efficiency compared with the benchmark NLO material of AgGaSe2 for λ < 2.0 μm. However, this innate SHG is highly suppressed for λ > 2.0 μm since the short-range noncentrosymmetric order partially intact in the glassy phase becomes averaged out for longer wavelengths. Though, we note that this unusual SHG occurs only in a few phase-change chalcogenide materials [29, 33, 34, 36]. More importantly, the SHG efficiency of the compound can be boosted by several orders of magnitude by simply exploiting its phase-change property as demonstrated by the blue dots in Fig. 6(b).

Figure 5(c) shows the broadband THG response (×100 magnified) from glassy KPSe6 (red) under the same experimental conditions. Note that the measured THG counts are almost comparable with those from AgGaSe2 (circles) and the response becomes most efficient at the mid-IR range. As the leading NLO process in an isotropic medium, THG becomes more efficient than the lower-order SHG process for λ > 2.0 μm due to the reason explained above. A considerable disparity between SHG and THG responses, therefore, proves that any contribution to frequency tripling by sum frequency generation of SHG and fundamental radiation (3ω = 2ω + ω) is negligible; this cascade process is unlikely in a powdered sample in which random scattering by powders prevents coherent overlap of both beams. As shown in Fig. 6(c), the THG response can be further enhanced by several times with simple annealing.

The crystalline KPSe6 exhibits strong phase-matching SHG within our entire observation range. For example, Fig. 6(a) illustrates particle-size dependence of the SHG counts at λ = 1.8 μm, clearly showing a growing trend with increasing d (type-I phase-matchable). Figure 6(b) displays the semi-logarithmic (semi-log) plots of relative SHG counts from crystalline KPSe6 (blue) and AgGaS2 (black) at d = 125 – 150 μm as a function of λ for direct comparison. Although the SHG responses from both sample and reference show a similar tendency at mid-IR, they are noticeably different for lower λ due to distinct degree of bandgap absorption. From Fig. 6(b) we can estimate the absolute value of χ(2) of the sample because AgGaS2 is also phase-matchable with a well-established value of χref(2)=36 pm/V [31]. Here the relative SHG counts must be compared in the static limit (plateau region), where bandgap absorption and multiphoton absorption are minimal. The χ(2) value of crystalline KPSe6 was calculated for the phase-matching case [30]:

χ(2)=χref(2)(ISHG/ISHGref)1/2142.8±10.5pm/V,
where ISHG and ISHGref are the averaged SHG counts from the sample and reference in the static mid-IR range (λ = 2.7 – 3.0 μm), respectively. Our experimental value of χ(2) is in good agreement with the first-principle calculation, which yields χ(2) = 151.3 pm/V [25]. This χ(2) value is the highest among phase-matching NLO materials having bandgaps over 1.0 eV.

The broadband THG response (×100 magnified) from crystalline KPSe6 (d = 20 – 32 μm) is plotted as the blue dots in Fig. 6(c), which is noticeably higher than that from AgGaSe2 (circles). We confirmed that THG is not phase-matchable. However, the majority of important third-order NLO applications do not involve beam mixing, and therefore, this stringent condition is not required. Estimation of the third-order NLO tensor χ(3) based on the powder method is less straightforward because the corresponding THG coherence length lc is much shorter than experimentally accessible particle sizes due to a large phase-mismatch. By simply assuming similar THG coherence lengths ( lc~lcref) and comparing with χref(3)=1.6×105 pm2/V2 of AgGaSe2[37], we estimated χ(3) ≃ (4.7 ± 0.6) × 105 pm2/V2 using

χ(3)=χref(3)(ITHG/ITHGref)1/2(lcref/lc)χref(3)(ITHG/ITHGref)1/2,
where ITHG and ITHGref are the averaged THG counts from the sample and reference at λ = 2.7–3.0 μm, respectively. More precise estimations of the third-order NLO parameters will be provided elsewhere by performing a traditional Z-scan technique [38] on the film form of the compound.

Although KPSe6 has a wide linear transparency window ranging from the fundamental bandgap in the visible regime to P-Se vibration modes in the far-IR (∼ 19.0 μm) [19, 20, 24], it is important to check the absence of multiphoton absorption for high-power NLO applications. The effect of multiphoton absorption can be probed by observing the power dependence of NLO harmonic signals [33, 34] that can noticeably deviate from the predicted power laws if multiphoton processes are significant. We found that this detrimental effect is insignificant in KPSe6 up to the 30-ps pulse of 30 μJ, corresponding to a high irradiance of 1 GW/cm2. For example, Figs. 7(a) and 7(b) display the measured SHG and THG spectra as a function of input pulse energies from 10 μJ to 30 μJ at λ = 1.55 μm, which corresponds to the major telecommunication wavelength. The corresponding power dependence is plotted by dots (SHG) and circles (THG) in Fig. 7(c) on a log-log scale. The solid lines are best fits to the data generated by simple square and cubic dependences, clearly indicating that multiphoton effects are negligible. Together with highly-efficient broadband NLO responses, this shows that KPSe6 is a promising candidate for NLO applications especially in the mid-IR spectral range.

 figure: Fig. 7

Fig. 7 (a) SHG spectra and (b) THG spectra at λ = 1.55 μm for several incident pulse energies (10 – 30 μJ), respectively. (c) Corresponding power dependence of SHG (dots) and THG (circles) superimposed by theoretical square and cubic fits (solid lines).

Download Full Size | PDF

5. Conclusions

We have investigated the NLO properties of a potential mid-IR NLO chalcogenide compound KPSe6 in which a crystal-glass phase transition is reproducible as confirmed by DTA analysis. The phase-change aspects of the dual-phase material have been characterized in terms of PXRD and Raman experiments. The compound exhibits apparent optical contrast between the two phases as confirmed by optical absorption measurements. A series of NLO experiments has been conducted on both crystalline and armophous KPSe6 over a broad wavelength range based on powder SHG and THG measurements. The observed NLO responses strongly depend on the fundamental wavelength but reach static ones in the mid-IR range without any signature of multiphoton absorption effects. Remarkably, the glassy phase of the compound yields an intrinsic but significant SHG response, which basically arises due to a partial retention of the local noncentrosymmetric order. The NLO efficiencies of the glassy compound can be further enhanced by converting into the crystalline counterpart with simple heat treatment. We have estimated the key NLO parameters of the crystalline compound; χ(2) ≃ 142.8 ± 10.5 pm/V and χ(3) ≃ (4.7 ± 0.6) × 105 pm2/V2. The observed strong optical nonlinearities are consistent with the theoretical calculation, which emphasizes the role of the low-dimensional structural anisotropy coupled with a strong covalent character in the KPSe6 compound. The results of our study demonstrate that KPSe6 can be utilized for both second- and third-order mid-IR NLO applications. Furthermore, due to its unique physical and chemical properties, optical fibers and thin films having strong NLO properties can be readily engineered. Further studies will focus on these forms of the material.

Acknowledgments

This work is supported by the National Science Foundation Grant DMR-1104965. A.S.H. gratefully acknowledges support by the National Science Foundation through a Graduate Research Fellowship under Grant No. DGE-1324585.

References and links

1. D. Cotter, R. J. Manning, K. J. Blow, A. D. Ellis, A. E. Kelly, D. Nesset, I. D. Phillips, A. J. Poustie, and D. C. Rogers, “Nonlinear optics for high-speed digital information processing,” Science 286, 1523–1528 (1999). [CrossRef]   [PubMed]  

2. G. A. Thomas, D. A. Ackerman, P. R. Prucnal, and S. L. Cooper, “Physics in the whirlwind of optical communications,” Phys. Today 53, 30–36 (2000). [CrossRef]  

3. W. R. Zipfel, R. M. Williams, and W. W. Webb, “Nonlinear magic: multiphoton microscopy in the biosciences,” Nat. Biotechnol. 21, 1369–1377 (2003). [CrossRef]   [PubMed]  

4. M. Pushkarsky, A. Tsekoun, I. G. Dunayevskiy, R. Go, and C. K. N. Patel, “Sub-parts-per-billion level dectection of NO2using room-temperature quantum cascade lasers,” Proc. Natl. Acad. Sci. USA 103, 10846–10849 (2006). [CrossRef]  

5. D. Pestov, X. Wang, G. O. Ariunbold, R. K. Murawski, V. A. Sautenkov, A. Dogariu, A. V. Sokolov, and M. O. Scully, “Single-shot detection of bacterial endospores via coherent Raman spectroscopy,” Proc. Natl. Acad. Sci. USA 105, 422–427 (2008). [CrossRef]   [PubMed]  

6. V. A. Serebryakov, E. V. Boiko, N. N. Petrishchev, and A. V. Yan, “Medical applications of mid-IR lasers. Problems and prospects,” J. Opt. Technol. 77, 6–17 (2010). [CrossRef]  

7. J.D. Luo, S. Huang, Z.W. Shi, B.M. Polishak, X.H. Zhou, and A.K.Y. Jen, “Tailored organic electro-optic materials and their hybrid systems for device applications,” Chem. Mater. 23, 544–553 (2011). [CrossRef]  

8. C. Conti, A.J. Agranat, and E. DelRe, “Subwavelength optical spatial solitons and three-dimensional localization in disordered ferroelectrics: Toward metamaterials of nonlinear origin,” Phys. Rev. A 84, 043809 (2011). [CrossRef]  

9. J. Parravicini, C. Conti, A.J. Agranat, and E. DelRe, “Programming scale-free optics in disordered ferroelectrics,” Opt. Lett. 37, 2355–2357 (2012). [CrossRef]   [PubMed]  

10. J. Parravicini, A.J. Agranat, C. Conti, and E. DelRe, “Equalizing disordered ferroelectrics for diffraction cancellation,” Appl. Phys. Lett. 101, 111104 (2012). [CrossRef]  

11. A. B. Seddon, “Chalcogenide glasses: a review of their preparation, properties and applications,” J. Non-Cryst. Solids 184, 44–50 (1995). [CrossRef]  

12. G.C. Bhar, S. Das, U. Chatterjee, P. K. Datta, and Y. N. Andreev, “Noncritical second harmonic generation of CO2 laser radiation in mixed chalcopyrite crystal,” Appl. Phys. Lett. 63, 1316–1318 (1993). [CrossRef]  

13. K. Kato, “Second-harmonic generation and sum-frequency generation in ZnGeP2,” Appl. Opt. 36, 2506–2510 (1997). [CrossRef]   [PubMed]  

14. W. Shi, Y. J. Ding, X. Mu, and N. Fernelius, “Tunable and coherent nanosecond radiation in the range of 2.7 – 28.7 μm based on difference-frequency generation in gallium selenide,” Appl. Phys. Lett. 80, 3889–3891 (2002). [CrossRef]  

15. J. H. Liao, G. M. Marking, K. F. Hsu, Y. Matsushita, M. D. Ewbank, R. Borwick, P. Cunningham, M. J. Rosker, and M. G. Kanatzidis, “α- and β-A2Hg3M2S8(A = K, Rb; M = Ge, Sn): Polar quaternary chalcogenides with strong nonlinear optical response,” J. Am. Chem. Soc. 125, 9484–9493 (2003). [CrossRef]   [PubMed]  

16. J. I. Jang, I. Chung, J. B. Ketterson, and M. G. Kanatzidis, ” “Nonlinear optical chalcogenide fibers and films of APSe6(A = K, Rb),” in New Developments in Photon and Materials Research, J. I. Jang, eds. (NOVA Scientific Publishers, 2013), pp. 195–218 and references therein.

17. M. Asobe, “Nonlinear optical properties of chalcogenide glass fibers and their application to all-optical switching,” Opt. Fiber Technol. 3, 142–148 (1997). [CrossRef]  

18. A. Zakery and S. R. Elliott, Optical Nonlinearities in Chalcogenide Glasses and Their Applications (Springer, 2007).

19. I. Chung, J. I. Jang, C. D. Malliakas, J. B. Ketterson, and M. G. Kanatzidis, “Strongly nonlinear optical glass fibers from noncentrosymmetric phase-change chalcogenide materials,” J. Am. Chem. Soc. 132, 384–389 (2010). [CrossRef]  

20. I. Chung, M. -G. Kim, J. I. Jang, J. He, J. B. Ketterson, and M. G. Kanatzidis, “Strongly nonlinear optical chalcogenide thin films of APSe6(A = K, Rb) form spin-coating,” Angew. Chem. Int. Ed. 50, 10867–10870 (2011). [CrossRef]  

21. T. K. Bera, J. I. Jang, J.-H. Song, C. D. Malliakas, A. J. Freeman, J. B. Ketterson, and M. G. Kanatzidis, “Soluble semiconductors AAsSe2(A = Li, Na) with a direct-band-gap and strong second harmonic generation: A combined theoretical and experimental study,” J. Am. Chem. Soc. 132, 3484–3495 (2010). [CrossRef]   [PubMed]  

22. C. D. Morris, C. D. Malliakas, and M. G. Kanatzidis, “Germanium selenophosphates: The incommensurately modulated 1[Ge4xPxSe124]and the molecular [Ge2P2Se14]6−,” Inorg. Chem. 50, 10241–10248 (2011). [CrossRef]   [PubMed]  

23. S. Banerjee, C. D. Malliakas, and M. G. Kanatzidis, “New layered tin(2) thiophosphates ASnPS4(A = K, Rb, Cs): Synthesis, structure, glass formation, and the modulated CsSnPS4,” 51, 11562–11573 (2012).

24. I. Chung, J. Do, C. G. Canlas, D. P. Weliky, and M. G. Kanatzidis, “APSe6(A = K, Rb, and Cs): Polymeric selenophosphates with reversible phase-change properties,” Inorg. Chem. 43, 2762–2764 (2004). [CrossRef]   [PubMed]  

25. J. -H. Song, A. J. Freeman, I. Chung, T. K. Bera, and M. G. Kanatzidis, “First-principles prediction of an enhanced optical second-harmonic susceptibility of low-dimensional alkali-metal chalcogenides,” Phys. Rev. B 79, 245203 (2009). [CrossRef]  

26. T. J. McCarthy, S. P. Ngeyi, J. H. Liao, D. C. DeGroot, T. Hogan, C. R. Kannewurf, and M. G. Kanatzidis, “Molten salt synthesis and properties of three new solid-state ternary bismuth chalcogenides, β-CsBiS2, γ-CsBiS2, and K2Bi8Se13,” Chem. Mater. 5331–340 (1993). [CrossRef]  

27. B. Frumarová, P. Nĕmec, M. Frumar, and J. Oswald, “Synthesis and properties of Ge-Sn-S: NdCl3 glasses,” Semiconductors 32, 812–816 (1998). [CrossRef]  

28. C. Bischoff, K. Schuller, M. Haynes, and S. W. Martin, “Structural investigations of y Na2S+(1 − y)PS5/2glasses using Raman and infrared spectroscopies,” J. Non-Cryst. Solids 358, 3216–3222 (2012). [CrossRef]  

29. I. Chung, C. D. Malliakas, J. I. Jang, C. G. Canlas, D. P. Weliky, and M. G. Kanatzidis, “Helical polymer 1[P2Se62]: Strong second harmonic generation response and phase-change properties of its K and Rb Salts,” J. Am. Chem. Soc. 129, 14996–15006 (2007). [CrossRef]   [PubMed]  

30. S. K. Kurtz and T. T. Perry, “A powder technique for the evaluation of nonlinear optical materials,” J. Appl. Phys. 39, 3798–3813 (1968). [CrossRef]  

31. G. C. Bhar, “Silver thiogallate and related diamond-like semiconductors for nonlinear-optical laser devices,” Jpn. J. Appl. Phys. Part I, Supplement 32, 653–659 (1993).

32. D. N. Nikogosyan, Nonlinear Optical Crystals: A Complete Survey, 1st edition (Springer, 2005).

33. J. I. Jang, S. Park, C. M. Harrison, D. J. Clark, C. D. Morris, I. Chung, and M. G. Kanatzidis, “K4GeP4Se12: A case for phase-change nonlinear optical chalcogenide,” Opt. Lett. 38, 1316–1318 (2013). [CrossRef]   [PubMed]  

34. C. D. Morris, I. Chung, S. Park, C. M. Harrison, D. J. Clark, J. I. Jang, and M. G. Kanatzidis, “Molecular germanium selenophosphate salts: Phase-change properties and strong second harmonic generation,” J. Am. Chem. Soc. 134, 20733–20744 (2012). [CrossRef]   [PubMed]  

35. See, for example, http://www.uqgoptics.com/materials_commercial_corning_pyrexBorosilicate.aspx.

36. I. Chung, J. I. Jang, M. A. Gave, D. P. Weliky, and M. G. Kanatzidis, “Low valent phosphorus in the molecular anions [P5Se12]5−and β-[P6Se12]4−: phase change behavior and near infrared second harmonic generation,” Chem. Commun. 47, 4998–5000 (2007). [CrossRef]  

37. http://marzenell.de/Research/Z-Scan/z-scan.html.

38. M. Sheik-Bahae, A. A. Said, T. Wei, D. J. Hagan, and E. W. Van Stryland, “Sensitive measurement of optical nonlinearities using a single beam,” IEEE J. Quantum Electron. 26, 760–769 (1990). [CrossRef]  

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 (7)

Fig. 1
Fig. 1 Crystal structure of KPSe6 viewed down the b-axis.
Fig. 2
Fig. 2 (a) Experimental PXRD patterns of crystalline KPSe6 (blue) and glassy KPSe6 (red) compared to the simulated pattern (black). Each trace is vertically translated for clarity. (b) Solid-state UV-VIS absorption spectra of crystalline (blue) and amorphous (red) KPSe6 superimposed by the theoretical fits (solid lines).
Fig. 3
Fig. 3 (a) DTA of crystalline KPSe6 from 50 C to 400 C, displaying three cycles to demonstrate its reversible crystal-glass phase-change properties; first cycle (black), second cycle (red), and third cycle (blue). (b) Raman spectra of crystalline (blue) and glassy (red) KPSe6. Each trace is vertically translated for clarity.
Fig. 4
Fig. 4 Wavelength-dependent (a) SHG from AgGaS2 and (b) THG from AgGaSe2.
Fig. 5
Fig. 5 (a) Non-phase-matching SHG from glassy KPSe6 based on particle-size dependence at λ = 1.8 μm. Broadband (b) SHG counts and (c) THG counts from glassy KPSe6 (red) and AgGaSe2 (circles), respectively, plotted on a semi-log scale.
Fig. 6
Fig. 6 (a) Phase-matching SHG from crystalline KPSe6 based on particle-size dependence at λ = 1.8 μm. (b) Broadband SHG counts from crystalline KPSe6 (blue) and AgGaS2 (black) at d = 125 – 150 μm. (c) Broadband THG counts from crystalline KPSe6 (blue) and AgGaSe2 (circles) at d = 20 – 32 μm.
Fig. 7
Fig. 7 (a) SHG spectra and (b) THG spectra at λ = 1.55 μm for several incident pulse energies (10 – 30 μJ), respectively. (c) Corresponding power dependence of SHG (dots) and THG (circles) superimposed by theoretical square and cubic fits (solid lines).

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

χ ( 2 ) = χ ref ( 2 ) ( I S H G / I S H G ref ) 1 / 2 142.8 ± 10.5 pm / V ,
χ ( 3 ) = χ ref ( 3 ) ( I T H G / I T H G ref ) 1 / 2 ( l c ref / l c ) χ ref ( 3 ) ( I T H G / I T H G ref ) 1 / 2 ,
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.