Next Article in Journal
Assessing the Sustainability and Acceptance Rate of Cost-Effective Household Water Treatment Systems in Rural Communities of Makwane Village, South Africa
Next Article in Special Issue
Tunable Stokes Laser Based on KTiOPO4 Crystal
Previous Article in Journal
Electro-Optic Intensity Modulation in Fe-Doped KTa0.65Nb0.35O3 Crystals
Previous Article in Special Issue
Ablation of BaWO4 Crystal by Ultrashort Laser Pulses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stimulated Raman Scattering in Yttrium, Gadolinium, and Calcium Orthovanadate Crystals with Single and Combined Frequency Shifts under Synchronous Picosecond Pumping for Sub-Picosecond or Multi-Wavelength Generation around 1.2 µm

by
Milan Frank
1,
Sergei N. Smetanin
2,*,
Michal Jelínek
1,
David Vyhlídal
1,
Lyudmila I. Ivleva
2,
Elizaveta E. Dunaeva
2,
Irina S. Voronina
2,
Dmitry P. Tereshchenko
2,
Vladislav E. Shukshin
2,
Petr G. Zverev
2 and
Václav Kubeček
1
1
Department of Physical Electronics, Faculty of Nuclear Sciences and Physical Engineering, Czech Technical University in Prague, Břehová 7, 11519 Prague 1, Czech Republic
2
Research Center for Laser Materials and Technologies, Prokhorov General Physics Institute of the Russian Academy of Sciences, Vavilova 38, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Crystals 2020, 10(10), 871; https://doi.org/10.3390/cryst10100871
Submission received: 24 August 2020 / Revised: 16 September 2020 / Accepted: 24 September 2020 / Published: 26 September 2020
(This article belongs to the Special Issue Multifunctional Optical Crystals for Raman Lasers)

Abstract

:
Comparative investigation of stimulated Raman scattering (SRS) characteristics in the YVO4, GdVO4, and Ca3(VO4)2 orthovanadate crystals at both low and high frequency anionic group vibrations is presented. It was found that GdVO4 is the most perspective for SRS generation on both the ν1 stretching and ν2 bending modes of internal anionic group vibrations with the strongest SRS pulse shortening under synchronous picosecond pumping. It is as a result of GdVO4‘s widest linewidth (17cm−1) of the homogeneously broadened scheelite-type component of the bending ν2 Raman line that led to the strongest SRS pulse shortening down to the dephasing time of the widest (scheelite-type) Raman mode at the secondary intracavity short-shifted SRS conversion. It allowed us to achieve SRS pulses with sub-picosecond duration under tens-of-picoseconds pumping due to the strongest 42-fold pulse shortening. Using the Ca3(VO4)2 crystal with essentially wider Raman lines (~50cm−1) did not allow us to generate SRS pulses shorter than 1 ps. It can be explained by inhomogeneous broadening of the Raman lines in Ca3(VO4)2 because of its structural disordering. Using the measured SRS pulse duration, the homogeneous broadening of the inhomogeneously broadened bending Raman line of Ca3(VO4)2 was estimated to be ~9cm−1. Among the orthovanadate crystals, the YVO4 crystal with the highest Raman gain and with homogeneously broadened Raman lines allowed us to realize the most efficient SRS lasing and SRS pulse shortening truly down to inverse half-width of the bending Raman line.

Graphical Abstract

1. Introduction

Yttrium, gadolinium, and calcium orthovanadate crystals are well-known as the laser host materials [1,2]. These crystals can also be used as the active materials for stimulated Raman scattering (SRS) and even for self-Raman lasing at the most intense Raman mode with the highest frequency of ν1~854–889 cm−1. This Raman mode corresponds to totally symmetric V-O stretching internal vibration in the tetrahedral (VO4)3– anionic group [3,4,5,6]. Recently, self-Raman lasing was also obtained on secondary Raman modes in Nd:YVO4 and Nd:GdVO4 with frequencies of ~380 cm−1 (O-V-O bending mode) [7,8] and ~260 cm−1 (Y-O stretching mode in YVO4) [9,10]. One more interesting property of the orthovanadate crystals is the presence of two intense vibrational lines in one of the polarized spontaneous Raman scattering configurations: the most intense one corresponds to the above mentioned high-frequency (ν1) stretching mode, but the second line with a low frequency of ν2~354–382 cm−1 corresponds to O-V-O bending internal vibration of the crystal anionic group [11,12]. This is similar to polarized Raman spectra of scheelite-type crystals [13], but the crystalline structure is different leading to wider Raman lines in the orthovanadate crystals. This feature is interesting for the method of SRS pulse shortening down to an inverse width of the widest Raman line under synchronously pumped SRS with combined frequency shift [14]. This method comprises a cascade process of primary extracavity Raman conversion from pump radiation into a ν1-shifted Stokes component and secondary intracavity Raman conversion from the ν1-shifted Stokes component into the (ν1 + ν2)-shifted Stokes component with significant pulse shortening down to the inverse width of the widest ν2 line. It was realized in a set of alkali-earth molybdate and tungstate crystals having a scheelite-type structure [14,15,16,17] where the ν2 line is the widest because of an overlap of two symmetric bending (scissoring and twisting) modes (Ag + Bg) of internal vibrations of the crystal anionic group [13]. In the SrMoO4 and SrWO4 crystals, the ν2 line width is the widest among the scheelite-type crystals amounting Δν2~10cm−1 that allowed to realize SRS pulse shortening down to 1/(π∙Δν2∙c)~1ps [15,16]. Recently, using this method, the strongest 42-fold SRS pulse shortening down to 860fs [18] has been achieved in a zircon-type GdVO4 crystal because the ν2 line is extremely wide (Δν2 = 24 cm−1), but the nature of its widening differs from that in scheelite-type crystals, and the minimal SRS pulse duration (860 fs) is longer than 1/(π∙Δν2∙c).
In the present work, comparative investigation of SRS in YVO4, GdVO4, and Ca3(VO4)2 crystals with a single and combined frequency shift on both Raman modes (stretching and bending) under synchronous picosecond pumping was carried out taking into account the crystalline and vibrational structure to answer the question how the nature of the bending Raman mode line widening affects the SRS pulse shortening.

2. Raman Crystals Characterization

Yttrium and gadolinium orthovanadate crystals (MVO4 where M = Y or Gd) crystallize in the tetragonal zircon-type structure (I41/amd space group). Similarly to tetragonal scheelite-type crystals, they have vibrational structure with the most intense, high-frequency (stretching mode), narrow (Δν1~3 cm−1) Raman line (ν1 = 889 cm–1 for YVO4 and ν1= 882 cm–1 for GdVO4) and the widest low-frequency (bending mode) Raman line (ν2 = 376 cm–1 for YVO4 and ν2 = 382 cm–1 for GdVO4). The zircon and scheelite structures containing tetrahedral anionic groups with strong covalent bond are similar in many respects. However, unlike the scheelite-type crystals, the ideal zircon structure (ZrSiO4) has no overlap of two bending modes (Ag and Bg) in the low-frequency ν2 Raman line, but the ν2 Raman line in YVO4 and GdVO4 has even stronger widening depending on temperature [19]. Figure 1 demonstrates decomposition of the ν2 Raman line in the Raman spectra of YVO4 and GdVO4 crystals into two components, one of which (red dashed lines) is strongly dependent on temperature.
It must be noted that this additional temperature-dependent component in the ν2 Raman line (red dashed lines in Figure 1) is absent in the ideal zircon (ZrSiO4) structure, and so it is caused by structural disordering in YVO4 and GdVO4. In work [19], it is explained by a thermally activated rotation of some tetrahedral anionic groups through 45° around their fourfold axis giving a partial zircon–scheelite phase transition. Thus, the additional component in the ν2 Raman line can be identified as the scheelite-type internal vibration in these 45°-rotated anionic groups with an overlap of two bending modes (Ag + Bg), and so, this scheelite-type component (red dashed lines in Figure 1) is wider than the ordinary zircon-type component (blue dotted lines in Figure 1) including only one (Ag) bending mode. This process is stronger in the GdVO4 crystal with a larger cation, where the scheelite-type and zircon-type component intensities are comparable already at room temperature. In GdVO4 at 300 K (Figure 1e), we have the zircon-type component linewidth of Δν21 = 10.4 cm−1 (the blue dotted line) and the scheelite-type component linewidth of Δν22 = 17 cm−1 (the red dashed line) giving the full ν2 Raman line width of Δν2 = 24 cm−1 (black solid line) [18]. In YVO4 at 300 K (Figure 1b), the additional scheelite-type component is low-intense yet, and therefore, the full ν2 Raman line width is only Δν2 = 11 cm–1 defined mainly by the zircon-type component. However, at increased temperature up to 600 K (Figure 1c), the scheelite-type component intensity becomes comparable with the zircon-type component intensity giving strong widening of the full ν2 Raman line up to ~40 cm−1. In GdVO4 at 600 K (Figure 1f) the scheelite-type component becomes significantly more intense than the zircon-type component, and therefore, the full ν2 Raman line width of ~50 cm−1 is mainly defined by the scheelite-type component.
Both the stretching ν1 and bending ν2 Raman lines in YVO4 and GdVO4 are intense under excitation with polarization of light parallel to the crystal optical axis with scattering configuration y(zz)y. In this case, values of Raman scattering peak cross-section of theν1 line differ each other only by 15% among these crystals at 300 K. For instance, the GdVO4 crystal has Raman gain of 4.5 cm/GW under 1.06-μm excitation [3], while YVO4 has a slightly higher gain of ~5 cm/GW.
The Ca3(VO4)2 crystal crystallizes in the whitlockite-type structure (R3c space group), which is a distorted variant (displacement of calcium and oxygen atoms from their ideal positions with random distribution of vacancies) of the ideal palmierite (K2Pb(SO4)2) structure (R 3 ¯ m space group) [20,21]. Disordering of the nonequivalent oxygen atoms with different local coordination spheres leads to an overlap of many vibrational modes that would otherwise be split in the case when the vacancies were ordered (in the ideal palmierite structure) [21]. It results in the partial coupling of the internal modes of the tetrahedral (VO4)3– anion vibrations with the strongest broadening of the stretching and bending Raman lines at room temperature demonstrated in Figure 2. It must be noted that in the polarized Raman spectra (Figure 2), both the stretching (ν1 = 854 cm−1) and bending (ν2 = 354 cm–1) Raman lines in the Ca3(VO4)2 crystal are intense in scattering configuration y(xx)y in contrast to the YVO4 and GdVO4 crystals, where it was in scattering configuration y(zz)y. Both values of the line width amount ~50 cm−1. The Ca3(VO4)2 crystal has Raman gain of about 1.6cm/GW under 1.06-μm excitation, which is 2.4 times lower than that in the KGd(WO4)2 [4] and in YVO4 and GdVO4 crystals.
For comparative study, we used a-cut YVO4 and GdVO4 crystals both with lengths of 16 mm and with anti-reflection coatings (T > 99% at 1000–1400 nm). Because of lower Raman gain, we used a longer Brewster-cut Ca3(VO4)2 crystal with a length of 30 mm.

3. Experimental Setup

The experimental setup of the synchronously pumped crystalline Raman laser system is presented in Figure 3. The YVO4, GdVO4, or Ca3(VO4)2 crystal was placed in an external ring cavity and synchronously pumped by the 36-ps, 150-MHz, 330-nJ Nd:GdVO4 laser at a wavelength of λ0 = 1063 nm. This pump laser was the same as used by us earlier for the scheelite-type Raman crystals [14,15,16,17]. The YVO4 and GdVO4 crystal optical axis (c) was oriented horizontally for pumping by Ec enabling to access the maximum intensities of the ν1 and ν2 Raman lines [11]. The Ca3(VO4)2 crystal optical axis (c) was oriented vertically for pumping by E c enabling to access the maximum intensities of the ν1 and ν2 Raman lines (see Figure 2). For optimal mode matching between the pump beam and the Raman laser cavity mode, the pump beam was focused into the crystal by a spherical lens (f = 100 mm). The external bow-tie ring cavity of the Raman laser system was compensated for astigmatism and consisted of two concave high-reflective (HR) mirrors PM and M1 having identical radius of curvature of 100 mm, a flat HR mirror M2 (HR@1169–1174 nm), and a flat output coupler OC. We tested two output couplers noted as OC1 and OC2. OC1 (R = 87–88%@1169–1174 nm, R = 80%@1220–1228 nm) was used for oscillation at the ν1-shifted first Stokes SRS wavelength of λ1 = (λ0−1ν1)−1. OC2 (R = HR@1169–1174 nm, R = 95%@1220 nm, and R = 90%@1228 nm) was used for oscillation at the unusual cascade SRS wavelength of λ12 = [λ0−1 − (ν1 + ν2)]−1 with a combined (ν1 + ν2) Raman shift. Both output couplers had high transmittance at the second (ν1 + ν1)-shifted Stokes component (HT@1299–1311 nm).
The generated Stokes components were separated by long-pass filter with the cut-on wavelength of 1200 nm (Thorlabs FEL 1200). The radiation spectra were measured by OceanOptics NIR512 spectrometer (wavelength range 850–1700 nm, FWHM resolution ~3 nm). The average power was measured with a Standa 11PMK-15SH5power meter; pulse energy was determined by calculation of average power, repetition rate, and duty factor of QCW pumping. The output pulses were measured by a laboratory designed non-collinear second harmonic generation (SHG) autocorrelator based on a LiIO3 crystal. For pulse duration calculation, we assumed a Gaussian shape of the measured autocorrelation curves.

4. Perfect Synchronization Case

The output characteristics of SRS radiation were significantly dependent on the synchronization condition that was adjusted by precise translation of the mirror M2. In the case of perfect synchronization of the pump pulses repetition period with the Raman laser cavity round-trip time, the lowest SRS oscillation thresholds and the highest slope efficiencies were achieved. Table 1 summarizes the SRS generation characteristics in the perfect synchronization case.
Measured output spectra of these lasers with the output coupler OC2 are presented in Figure 4 (in the case of OC1, the (ν1 + ν2)-shifted Stokes component in the output spectra was absent). The measured wavelengths of the ν1-shifted and (ν1 + ν2)-shifted Stokes SRS components (see Table 1 and Figure 4) are in a good agreement with the calculated values of λ1 = (λ0–1ν1)−1 and λ12 = [λ0–1 − (ν1 + ν2)]−1, respectively.
Figure 5 demonstrates the output–input energy characteristics in the perfect synchronization case for the YVO4, GdVO4, and Ca3(VO4)2 Raman lasers with the output couplers OC1 and OC2.
In the case of OC1, as is shown in Figure 5a,c,e, a single ν1-shifted Stokes component was generated with high slope efficiency for all the active crystals. The single-Stokes SRS generation threshold was achieved at the input pulse energies of 128, 133, and 188 nJ for the YVO4, GdVO4, and Ca3(VO4)2 active crystals, respectively. These values correlate with the row of their Raman gains. The slope efficiencies were 27.4, 24.7, and 14.3%, respectively. At the maximal input pump pulse energy of 330 nJ, the output pulse energies reached 54, 49, and 21 nJ (optical-to-optical efficiencies of 16.4, 14.8, and 6.4%) for YVO4, GdVO4, and Ca3(VO4)2, respectively. The SRS oscillation in the YVO4 Raman laser was more efficient due to a higher Raman gain.
In the case of OC2, as is shown in Figure 5b,d,f, the Raman laser generated not only the first ν1-shifted Stokes component but also the component with the combined (ν1 + ν2) Raman shift. The SRS generation threshold of this additional Stokes component was achieved at the input pulse energies of 141, 146, and 201 nJ for the YVO4, GdVO4, and Ca3(VO4)2 active crystals, respectively. It can be seen that slope efficiency for the first Stokes component (ν1-shifted) was low (1.2, 1.4, and 1.3% for YVO4, GdVO4, and Ca3(VO4)2) in comparison with 2.2–4.8 times higher slope efficiency for the (ν1 + ν2)-shifted Stokes component (5.8, 5.0, and 2.8% for YVO4, GdVO4, and Ca3(VO4)2). This is a result of using OC2 that has higher transmission (5%) around 1220 nm for realization of nonlinear cavity dumping of the Raman laser [22]. As a result, at the maximal input pump pulse energy of 330 nJ, we achieved the top output energies of 11, 10, and 4 nJ in the (ν1 + ν2)-shifted Stokes component for YVO4, GdVO4, and Ca3(VO4)2, respectively.

5. Cavity Length Detuning for the Strongest Pulse Shortening

For the purpose of obtaining the maximal SRS pulse shortening, we realized the Raman laser cavity length detuning relative to the perfect synchronization case (zero detuning). Figure 6 presents the measured dependencies of the output pulse duration and energy on the cavity length detuning for the YVO4, GdVO4, and Ca3(VO4)2 Raman lasers with the output coupler OC2 under maximal pump pulse energy of 330 nJ.
It is evident from Figure 6a,c,e that the detuning range, where the SRS generation was observed, was short (from 0 up to +50 μm) at the positive cavity length detuning, while the detuning range was significantly wider at the negative detuning (from 0 down to −200 μm). There is a similarity with other synchronously pumped Raman lasers [14,15,16,17,18,23,24,25,26,27,28], and it is explained by non-efficient interaction between the SRS and pump pulses where only the front edge of the SRS pulse is amplified at positive detuning. Nevertheless, positive cavity length detuning has allowed us to achieve the strongest SRS pulse shortening, as shown in Figure 6b,d,f.
It can also be observed from Figure 6b,d,f that the ν1-shifted Stokes pulses (at 1174, 1173, and 1169 nm) have been shortened only slightly from pump pulse tp = 36 ps to tS = 21–22 ps at the positive cavity length detuning of +50 μm. However, the (ν1 + ν2)-shifted Stokes pulses (at 1228, 1228, and 1220 nm) went through strong self-shortening from negative to positive cavity length detuning and pulse duration remained below 7ps. The strongest pulse shortening down to tS = 1.19 ± 0.06 ps for YVO4, tS = 0.86 ± 0.03 ps for GdVO4, and tS = 1.18 ± 0.24 ps for Ca3(VO4)2 was achieved at the positive cavity length detuning of +50 μm.
Strong pulse shortening can be explained by the theory of ultra-short SRS pulse formation at intracavity pumping [29] that predicted the shortened SRS pulse duration close to a dephasing time of a Raman mode equal to an inverse half-width of the homogeneously broadened Raman line. We should take into account that generation of the strongly shortened (ν1 + ν2)-shifted SRS pulses in the extracavity Raman laser took place on the bending (ν2) Raman line under intracavity pumping by the ν1-shifted SRS radiation.
For the YVO4 crystal with the homogeneously broadened bending Raman line with a linewidth of Δν2 = 11 cm−1 at room temperature (Figure 1b), the measured SRS pulse duration (tS = 1.2 ps) was really close to the inverse half-width of the bending Raman line (1/(πΔν2c) = 0.96ps). Similar results of the strongest pulse shortening were recently achieved for synchronously pumped SRS in scheelite-type crystals with homogeneously broadened Raman lines [14,15,16,17,18].
However, for the GdVO4 crystal having wider bending line width of Δν2 = 24 cm−1 at room temperature (Figure 1e), the measured SRS pulse duration (tS=0.86ps) was longer than 1/(πΔν2c) = 0.44 ps. However, the bending Raman line in GdVO4 consists of two homogeneously broadened spectral components (Figure 1e) corresponding to the zircon-type and scheelite-type structures with linewidths of Δν21 = 10.4 cm−1 and Δν22 = 17 cm−1, respectively. We can conclude that, the pulse duration is determined not by the total spectral width of the bending Raman line, but by that of the homogeneously broadened component. It can be both the zircon-type and scheelite-type components, because the SRS pulse duration is almost in the middle between the corresponding values of dephasing time of 1/(πΔν21c) = 1.02 ps and 1/(πΔν22c) = 0.62 ps, respectively. Taking into account all the previous results [14,15,16,17], we have never observed SRS pulse duration shorter than dephasing time of the homogeneously broadened bending Raman mode, and so the present result of 860 fs is more likely caused by the scheelite-type bending mode having a shorter dephasing time (0.62 ps).
Despite the widest (~50 cm−1) Raman lines in Ca3(VO4)2, the shortened SRS pulse (tS = 1.2 ps) was significantly longer than the inverse half-width of the bending Raman line of 1/(πΔν2c) = 0.21 ps. It can be explained by inhomogeneous broadening of Raman lines in Ca3(VO4)2 because of its structural disordering. Using the shortest generated SRS pulse duration, we can estimate the homogeneous broadening of the inhomogeneously broadened bending Raman line of Ca3(VO4)2 as 1/(πtSc) ≈ 9 cm−1.

6. Conclusions

We have presented a comparative investigation of synchronously pumped SRS in the YVO4, GdVO4, and Ca3(VO4)2 orthovanadate crystals at both long (ν1) and short (ν2) Raman shifts. It was found that among these crystals, GdVO4 is the most suitable for SRS generation on both the ν1 stretching and ν2 bending modes of internal anionic group vibrations with the strongest SRS pulse shortening under synchronous picosecond pumping. It is as a result of GdVO4‘s widest linewidth (Δν22 = 17 cm−1) of the homogeneously broadened scheelite-type component of the bending ν2 Raman line. This property led to the strongest SRS pulse shortening down to the dephasing time 1/(πΔν22c) of the widest (scheelite-type) Raman mode at the secondary intracavity short-shifted SRS conversion. It allowed us to achieve SRS pulses with sub-picosecond duration under tens-of-picoseconds pumping due to the strongest 42-fold pulse shortening. Using the Ca3(VO4)2 crystal with essentially wider Raman lines (~50 cm−1) did not allow us to generate SRS pulses shorter than 1ps. It can be explained by inhomogeneous broadening of the Raman lines in Ca3(VO4)2 because of its structural disordering. Using the measured SRS pulse duration, the homogeneous broadening of the inhomogeneously broadened bending Raman line of Ca3(VO4)2 was estimated to be ~9cm−1. Among the orthovanadate crystals, the YVO4 crystal with the highest Raman gain and with homogeneously broadened Raman lines allowed us to realize the most efficient SRS lasing and SRS pulse shortening truly down to inverse the half-width of the bending Raman line.

Author Contributions

Conceptualization, S.N.S.; Formal analysis, M.F. and S.N.S.; Investigation, M.F., S.N.S., M.J., D.V., V.E.S., L.I.I., E.E.D., D.P.T. and I.S.V.; Methodology, M.F. and S.N.S.; Resources, V.E.S., L.I., E.E.D. and I.S.V.; Supervision, P.G.Z. and V.K.; Validation, P.G.Z. and V.K.; Visualization, M.F. and D.P.T.; Writing—original draft, M.F. and S.N.S.; Writing—review & editing, M.F., S.N.S., M.J., D.V., P.G.Z. and V.K. All authors have read and agreed to the published version of the manuscript.

Funding

The research was supported by the Czech Science Foundation−Project No 18-11954S, and by the Russian Foundation for Basic Research−Project No 19-02-00723.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zagumennyi, A.I.; Mikhailov, V.A.; Shcherbakov, I.A. Rare earth ion lasers—Nd3+. In Handbook of Laser Technology and Applications, Volume II: Laser Design and Laser Systems; Institute of Physics Publishing: Bristol, PA, USA, 2004. [Google Scholar]
  2. Ivleva, L.I.; Dunaeva, E.E.; Voronina, I.S.; Doroshenko, M.E.; Papashvili, A.G.; Šulc, J.; Kratochvíl, J.; Jelínková, H. Impact of Tm3+/Ho3+ co-doping on spectroscopic and laser properties of Ca3(VO4)2 single crystal. J. Cryst. Growth 2019, 513, 10–14. [Google Scholar] [CrossRef]
  3. Kaminskii, A.A.; Ueda, K.I.; Eichler, H.J.; Kuwano, Y.; Kouta, H.; Bagaev, S.N.; Chyba, T.H.; Barnes, J.C.; Gad, G.M.A.; Murai, T.; et al. Tetragonal vanadates YVO4 and GdVO4—New efficient χ(3) materials for Raman lasers. Opt. Commun. 2001, 194, 201–206. [Google Scholar] [CrossRef]
  4. Zverev, P.G.; Karasik, A.Y.; Basiev, T.T.; Ivleva, L.I.; Osiko, V.V. Stimulated Raman scattering of picosecond pulses in SrMoO4 and Ca3(VO4)2 crystals. Quantum Electron. 2003, 33, 331–334. [Google Scholar] [CrossRef]
  5. Basiev, T.T.; Vassilev, S.S.; Kojushkin, V.A.; Osiko, V.V.; Zagumennyi, A.I.; Kutovoi, S.A.; Shcherbakov, I.A. Diode pumped 500-picosecond Nd: GdVO4 Raman laser. Laser Phys. Lett. 2004, 1, 237–240. [Google Scholar] [CrossRef]
  6. Huang, G.; Yu, Y.; Xie, X.; Zhang, Y.; Du, C. Diode-pumped simultaneously Q-switched and mode-locked YVO4/Nd:YVO4/YVO4 crystal self-Raman first-Stokes laser. Opt. Express 2013, 21, 19723–19731. [Google Scholar] [CrossRef]
  7. Lin, J.; Pask, H.M. Cascaded self-Raman lasers based on 382 cm–1 shift in Nd:GdVO4. Opt. Express 2012, 20, 15180–15185. [Google Scholar] [CrossRef] [Green Version]
  8. Li, R.; Bauer, R.; Lubeigt, W. Continuous-wave Nd: YVO4 self-Raman lasers operating at 1109 nm, 1158 nm and 1231 nm. Opt. Express 2013, 21, 17745–17750. [Google Scholar] [CrossRef] [Green Version]
  9. Fan, S.; Zhang, X.; Wang, Q.; Liu, Z.; Li, L.; Cong, Z.; Chen, X.; Zhang, X. 1097 nm Nd:YVO4 self-Raman laser. Opt. Commun. 2011, 284, 1642–1644. [Google Scholar]
  10. Bai, F.; Jiao, Z.; Xu, X.; Wang, Q. High power Stokes generation based on a secondary Raman shift of 259 cm–1 of Nd: YVO4 self-Raman crystal. Opt. Laser Technol. 2019, 109, 55–60. [Google Scholar] [CrossRef]
  11. Lu, G.; Li, C.; Wang, W.; Wang, Z.; Xia, H.; Zhao, P. Raman investigation of lattice vibration modes and thermal conductivity of Nd-doped zircon-type laser crystals. Mater. Sci. Eng. B 2003, 98, 156–160. [Google Scholar] [CrossRef]
  12. Li, C.; Yang, W.; Chang, Y. Raman scattering study of calcium orthovanadate crystal. Jpn. J. Appl. Phys. 1985, 24, 508–509. [Google Scholar] [CrossRef]
  13. Porto, S.P.S.; Scott, J.F. Raman spectra of CaWO4, SrWO4, CaMoO4, and SrMoO4. Phys. Rev. 1967, 157, 716–719. [Google Scholar] [CrossRef]
  14. Frank, M.; Smetanin, S.N.; Jelínek, M.; Vyhlídal, D.; Ivleva, L.I.; Zverev, P.G.; Kubeček, V. Highly efficient picosecond all-solid-state Raman laser at 1179 and 1227 nm on single and combined Raman lines in a BaWO4 crystal. Opt. Lett. 2018, 43, 2527–2530. [Google Scholar] [CrossRef] [PubMed]
  15. Frank, M.; Smetanin, S.N.; Jelínek, M.; Vyhlídal, D.; Kopalkin, A.A.; Shukshin, V.E.; Ivleva, L.I.; Zverev, P.G.; Kubeček, V. Synchronously-pumped all-solid-state SrMoO4 Raman laser generating at combined vibrational Raman modes with 26-fold pulse shortening down to 1.4 ps at 1220 nm. Opt. Laser Technol. 2019, 111, 129–133. [Google Scholar] [CrossRef]
  16. Frank, M.; Smetanin, S.N.; Jelínek, M.; Vyhlídal, D.; Shukshin, V.E.; Ivleva, L.I.; Zverev, P.G.; Kubeček, V. Efficient synchronously-pumped all-solid-state Raman laser at 1178 and 1227 nm on stretching and bending anionic group vibrations in a SrWO4 crystal with pulse shortening down to 1.4 ps. Opt. Laser Technol. 2019, 119, 105660. [Google Scholar] [CrossRef]
  17. Frank, M.; Smetanin, S.N.; Jelínek, M.; Vyhlídal, D.; Shukshin, V.E.; Ivleva, L.I.; Dunaeva, E.E.; Voronina, I.S.; Zverev, P.G.; Kubeček, V. Stimulated Raman scattering in alkali-earth tungstate and molybdate crystals at both stretching and bending Raman modes under synchronous picosecond pumping with multiple pulse shortening down to 1 ps. Crystals 2019, 9, 167. [Google Scholar] [CrossRef] [Green Version]
  18. Frank, M.; Smetanin, S.N.; Jelínek, M.; Vyhlídal, D.; Shukshin, V.E.; Zverev, P.G.; Kubeček, V. 860 fs GdVO4 Raman laser at 1228 nm pumped by 36 ps, 1063 nm laser. Laser Phys. Lett. 2019, 16, 085401. [Google Scholar] [CrossRef]
  19. Voronko, Y.K.; Sobol, A.A.; Shukshin, V.E.; Zagumennyi, A.I.; Zavartsev, Y.D.; Kutovoi, S.A. Raman spectroscopic study of structural disordering in YVO4, GdVO4, and CaWO4 crystals. Phys. Solid State 2009, 51, 1886–1893. [Google Scholar] [CrossRef]
  20. Gopal, R.; Calvo, C. The structure of Ca3(VO4)2. Z. für Krist. 1973, 137, 67–85. [Google Scholar]
  21. Grzechnik, A. High-temperature transformations in calcium orthovanadate studied with Raman scattering. Chem. Mater. 1998, 10, 1034–1040. [Google Scholar] [CrossRef]
  22. Smetanin, S.N.; Jelínek, M., Jr.; Kubeček, V.; Jelínková, H. Four-wave-mixing and nonlinear cavity dumping at 1.3 μm from Nd: SrMoO4 self-Raman laser. Laser Phys. Lett. 2016, 13, 015801. [Google Scholar] [CrossRef]
  23. Grigoryan, G.G.; Sogomonyan, S.B. Synchronously pumped picosecond Raman laser utilizing an LiIO3 crystal. Sov. J. Quantum Electron. 1989, 19, 1402–1404. [Google Scholar] [CrossRef]
  24. Granados, E.; Pask, H.M.; Spence, D.J. Synchronously pumped continuous-wave mode-locked yellow Raman laser at 559 nm. Opt. Express 2009, 17, 569–574. [Google Scholar] [CrossRef] [PubMed]
  25. Granados, E.; Spence, D.J. Pulse compression in synchronously pumped mode locked Raman lasers. Opt. Express 2010, 18, 20422–20427. [Google Scholar] [CrossRef] [PubMed]
  26. Warrier, A.M.; Lin, J.; Pask, H.M.; Mildren, R.P.; Coutts, D.W.; Spence, D.J. Highly efficient picosecond diamond Raman laser at 1240 and 1485 nm. Opt. Express 2014, 22, 3325–3333. [Google Scholar] [CrossRef]
  27. Murtagh, M.; Lin, J.; Mildren, R.P.; McConnell, G.; Spence, D.J. Efficient diamond Raman laser generating 65 fs pulses. Opt. Express 2015, 23, 15504–15513. [Google Scholar] [CrossRef]
  28. Lin, J.; Spence, D.J. 25.5 fs dissipative soliton diamond Raman laser. Opt. Lett. 2016, 41, 1861–1864. [Google Scholar] [CrossRef]
  29. Isaev, S.K.; Kornienko, L.S.; Kravtsov, N.V.; Serkin, V.N. Formation of ultrashort light pulses in a laser with a bleachable filter by intracavity generation of Raman emission. Sov. J. Quantum Electron. 1981, 11, 365–370. [Google Scholar] [CrossRef]
Figure 1. The bending mode ν2 Raman line decomposed into two components, one of which (red dashed lines) is strongly dependent on temperature, in the Raman spectra of (ac) YVO4 and (df) GdVO4 at various temperatures [19].
Figure 1. The bending mode ν2 Raman line decomposed into two components, one of which (red dashed lines) is strongly dependent on temperature, in the Raman spectra of (ac) YVO4 and (df) GdVO4 at various temperatures [19].
Crystals 10 00871 g001
Figure 2. Polarized Raman spectra of Ca3(VO4)2 at room temperature.
Figure 2. Polarized Raman spectra of Ca3(VO4)2 at room temperature.
Crystals 10 00871 g002
Figure 3. The experimental setup of the synchronously pumped extracavity crystalline Raman laser.
Figure 3. The experimental setup of the synchronously pumped extracavity crystalline Raman laser.
Crystals 10 00871 g003
Figure 4. Laser output radiation spectra of the (a) YVO4, (b) GdVO4, and (c) Ca3(VO4)2 Raman lasers with OC2.
Figure 4. Laser output radiation spectra of the (a) YVO4, (b) GdVO4, and (c) Ca3(VO4)2 Raman lasers with OC2.
Crystals 10 00871 g004
Figure 5. Dependencies of the output SRS pulse energy on the input pump pulse energy in the (a,b) YVO4, (c,d) GdVO4, and (e,f) Ca3(VO4)2 Raman lasers with (a,c,e) OC1 and (b,d,f) OC2.
Figure 5. Dependencies of the output SRS pulse energy on the input pump pulse energy in the (a,b) YVO4, (c,d) GdVO4, and (e,f) Ca3(VO4)2 Raman lasers with (a,c,e) OC1 and (b,d,f) OC2.
Crystals 10 00871 g005
Figure 6. Dependencies of (a,c,e) the output pulse energy and (b,d,f) the output pulse duration on the cavity length detuning for the (a,b) YVO4, (c,d) GdVO4, and (e,f) Ca3(VO4)2 Raman laser with the output coupler OC2 under pumping with the maximal pulse energy of 330 nJ.
Figure 6. Dependencies of (a,c,e) the output pulse energy and (b,d,f) the output pulse duration on the cavity length detuning for the (a,b) YVO4, (c,d) GdVO4, and (e,f) Ca3(VO4)2 Raman laser with the output coupler OC2 under pumping with the maximal pulse energy of 330 nJ.
Crystals 10 00871 g006
Table 1. The SRS characteristics in the YVO4, GdVO4, and Ca3(VO4)2 Raman lasers with OC1 and OC2, where λ is a generated wavelength, η is a slope efficiency, and Ep is an output pulse energy.
Table 1. The SRS characteristics in the YVO4, GdVO4, and Ca3(VO4)2 Raman lasers with OC1 and OC2, where λ is a generated wavelength, η is a slope efficiency, and Ep is an output pulse energy.
Crystalsν1-Shifted Stokes Component (OC1)(ν1 + ν2)-Shifted Stokes Component (OC2)
λ [nm]η [%]Ep [nJ]λ [nm]η [%]Ep [nJ]
YVO4117427.45412285.811
GdVO4117324.74912285.010
Ca3(VO4)2116914.32112202.84

Share and Cite

MDPI and ACS Style

Frank, M.; Smetanin, S.N.; Jelínek, M.; Vyhlídal, D.; Ivleva, L.I.; Dunaeva, E.E.; Voronina, I.S.; Tereshchenko, D.P.; Shukshin, V.E.; Zverev, P.G.; et al. Stimulated Raman Scattering in Yttrium, Gadolinium, and Calcium Orthovanadate Crystals with Single and Combined Frequency Shifts under Synchronous Picosecond Pumping for Sub-Picosecond or Multi-Wavelength Generation around 1.2 µm. Crystals 2020, 10, 871. https://doi.org/10.3390/cryst10100871

AMA Style

Frank M, Smetanin SN, Jelínek M, Vyhlídal D, Ivleva LI, Dunaeva EE, Voronina IS, Tereshchenko DP, Shukshin VE, Zverev PG, et al. Stimulated Raman Scattering in Yttrium, Gadolinium, and Calcium Orthovanadate Crystals with Single and Combined Frequency Shifts under Synchronous Picosecond Pumping for Sub-Picosecond or Multi-Wavelength Generation around 1.2 µm. Crystals. 2020; 10(10):871. https://doi.org/10.3390/cryst10100871

Chicago/Turabian Style

Frank, Milan, Sergei N. Smetanin, Michal Jelínek, David Vyhlídal, Lyudmila I. Ivleva, Elizaveta E. Dunaeva, Irina S. Voronina, Dmitry P. Tereshchenko, Vladislav E. Shukshin, Petr G. Zverev, and et al. 2020. "Stimulated Raman Scattering in Yttrium, Gadolinium, and Calcium Orthovanadate Crystals with Single and Combined Frequency Shifts under Synchronous Picosecond Pumping for Sub-Picosecond or Multi-Wavelength Generation around 1.2 µm" Crystals 10, no. 10: 871. https://doi.org/10.3390/cryst10100871

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop