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

Mid-IR lasing of Fe2+ ions via Cr2+ → Fe2+ energy transfer process with YLF:Er or laser diode pumping at 1.7 µm

Open Access Open Access

Abstract

A nonradiative Cr2+ → Fe2+ ions energy transfer process in Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) single crystal with a Cr2+ to Fe2+ ions concentration ratio about 2.5:1 was investigated under short-pulse (200 ns) ∼1.73 µm excitation by YLF:Er laser. Fe2+ ions lasing in a temperature range from 78 K up to 200 K was achieved and the central oscillation wavelength was observed to increase from 4.1 µm to 4.4 µm, depending on the temperature of the active medium. Moreover, stable Fe2+ ions lasing under a “long” pulse excitation (1 ms, 10 Hz) by a ∼1.71 µm laser diode were demonstrated. The laser generated in a temperature range from 78 K up to 110 K at a central wavelength around ∼4.17 µm. The maximum mean laser output power of 4.1 mW with a slope efficiency of 2.3% with respect to the absorbed power was achieved. These results present a new possibility to develop a simple and compact diode-pumped coherent mid-infrared (4.1–4.3 µm) laser sources.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

For several last years there is a strong demand on a development of laser systems based on Fe2+ ions operating in a spectral range of 4–5 µm pumped by a convenient laser source. Such 4–5 µm laser systems are interesting for a variety of possible applications in medicine diagnostics and treatment, spectroscopy, air pollutants measurement, free-space communications, target illumination, and industrial technologies [1]. In general, laser radiation in this wavelength range can be also obtained using difference frequency generation (DFG), hydrogen Raman lasers [2], CO lasers [3], or free-electron lasers [4]. DFG based on solid state lasers utilizing various non-oxide nonlinear crystals has been proven to be the efficient method for tunable mid-IR lasing within pulse duration range from femtosecond up to continuous-wave [5,6]. Radiation can be further amplified by optical parametric amplification (OPA), optical parametric chirped pulse amplification (OPCPA), or ZnSe:Fe2+ multi-pass amplifier [68]. Practical applications of these devices are somewhat limited by their complexity. Semiconductor lasers present other interesting and applicable alternative with the most recent output power at a level of several Watts [9,10]. Nevertheless, there is still strong interest in Fe2+:II-VI solid-state lasers broadly tunable in the mid-IR region. These lasers are usually pumped by ∼3 µm sources fitting the Fe2+ ions absorption maximum. Up to now, pumping at ∼3 µm generated mostly by free-running or Q-switched YAG:Er3+ (YSGG:Er3+), second D2 Raman Stokes of YAG:Nd3+, or electric discharge HF laser is used [1116]. Other interesting pump source is presented by an Er3+ doped ZBLAN fiber laser. A liquid-nitrogen-cooled (LN2), continuous-wave (CW) ZnSe:Fe2+ laser pumped by such fiber laser was presented recently with the maximum output power of 2.1 W and slope efficiency of 59% [17]. As for Cr2+ pumping, ZnSe:Cr2+ laser tuned to 2.94 µm was proved to be an efficient source for 9.2 W continuous-wave (CW), liquid-nitrogen-cooled ZnSe:Fe2+ laser [18]. All these pumping sources are usable but not as simple as laser diode pumping.

Other possible way to simplify the pumping scheme is a combination of Fe2+ ions with other active transitional metal ions inside the active medium and utilization of an energy transfer process. Up to now, some research was made to realize this pumping scheme of Fe2+ ions through the energy transfer from Co2+ or Cr2+ ions. Self-lasing at 3.9 µm from ZnSe:Co2+,Fe2+ and ZnS:Co2+,Fe2+ crystals was proven experimentally for temperatures below ∼24 K [14,19,20].

In the case of the energy transfer from Cr2+ ions, mid-IR luminescence of Fe2+ in ZnSe:Cr2+,Fe2+ at room temperature excited by 1.56 µm was documented in [21] at first. The Cr2+ → Fe2+ energy transfer was further confirmed also in ZnS:Cr2+,Fe2+ or under 1.77 µm excitation [11,22,23]. In-depth analysis of room temperature kinetics and energy transfer rate in the ZnSe:Cr2+,Fe2+ was published recently in [24]. Fe2+ luminescence was also detected under visible (532 nm) excitation [11,25] at cryogenic temperatures.

Cr2+ ions (donors) fluorescence spectrum should overlap Fe2+ ions (acceptors) absorption spectrum in order to obtain the energy transfer. In [26] Doroshenko et al. previously shown that absorption and fluorescence spectra of Cr2+ ions in Zn1-xMnxSe solid solutions with manganese ions presence are shifted towards longer wavelengths compared to a ZnSe crystal. At the same time, the shape of the Fe2+ ions absorption spectrum remains practically unchanged at the short-wavelength part with a steady shift of a long-wavelength part with absorption maximum towards longer wavelengths with increasing content (x) of manganese ions in the matrix. The Cr2+ → Fe2+ energy transfer process was investigated and Fe2+ ions gain-switched Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.3) laser operation under short pulse (200 ns) 1.73 µm excitation in a temperature range 78–150 K was presented by us [27]. Most recently, Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.3) laser operation under long pulse (1 ms, 10 Hz) 1.94 µm Tm:fiber pumping was also demonstrated [28].

2. Cr2+ → Fe2+ ions energy transfer process characteristics

In this paper, the energy transfer from excited Cr2+ to Fe2+ ions followed by the laser emission is presented. The lasing was obtained in the Cr2+, Fe2+ co-doped Zn1-xMnxSe (x = 0.05) single crystal within a temperature range from 78 K up to 110 K. The Zn1-xMnxSe:Cr2+,Fe2+ active medium was synthesized by solid-phase synthesis and grown by HPBM (High Pressure Bridgman Method) with active ions doping during the synthesis process. A 3.9 mm thick experimental sample was cut from a crystal boule and its faces were optically polished. There was no antireflective coating on the sample. The refractive index was measured using a prism-coupled refractometer and its value is 2.4404 at 1.7 µm. This value is for about 0.004 lower than that for the ZnSe crystal at the same wavelength. The concentrations were C(Cr) = 5×1018 cm−3 and C(Fe) = 2×1018 cm−3. Dopant ion concentrations in the synthesized sample were measured using inductively coupled plasma–optical emission spectrometry element analysis. In [27] Doroshenko et al. investigated the Cr2+ → Fe2+ energy transfer process in a highly-Mn-concentrated Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.3) crystal with about 1:2 Cr2+ to Fe2+ ions concentrations ratio. Here we present the results on low Mn concentrated Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal with about inverse (2.5:1) Cr2+ and Fe2+ ions concentration.

The absorption spectrum of Cr2+ and Fe2+ ions measured at 78 K in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal is presented in Fig. 1. For comparison the absorption spectrum of Cr2+ and Fe2+ ions in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.3) crystal at 78 K is also presented. The spectra were measured using a FTIR spectrometer (Infralum FT-08). As follows from this figure, Fe2+ ions concentration in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal is about 2.5 times lower than that for the x = 0.3 sample.

 figure: Fig. 1.

Fig. 1. Absorption spectra of Cr2+ and Fe2+ ions in the Zn1-xMnxSe:Cr2+,Fe2+ crystals with Mn content x = 0.05 (red) and x = 0.3 (blue) at 78 K.

Download Full Size | PDF

Fluorescence spectra of the Cr2+ ions together with the Fe2+ ions absorption spectra normalized to a unit area are plotted in Fig. 2. The Cr2+ ions fluorescence spectra were measured in Cr2+ doped samples (in the absence of the Fe2+ ions). As Nemec et al. discussed in [29], lower Mn content in the Zn1-xMnxSe results in a smaller shift of Cr2+ ions fluorescence maximum towards longer wavelengths. Also, higher Mn content results in both long wavelength shift and broadening of Fe2+ ions absorption lines. Both these facts lead to a significantly lower overlap of Cr2+ ions fluorescence with Fe2+ ions absorption at 78 K in the crystal with low Mn content x = 0.05 compared to the crystal with Mn content x = 0.3 (see hatched areas in Fig. 2). The overlap integral for the x = 0.05 was calculated to be about ∼ 3% of a normalized unit area and for the x = 0.3 this value was about 10% which should result in higher Cr2+ → Fe2+ energy transfer efficiency for crystal with higher Mn content.

 figure: Fig. 2.

Fig. 2. Fluorescence spectra of Cr2+ ions and absorption spectra of Fe2+ ions in the Zn1-xMnxSe:Cr2+,Fe2+ crystals with Mn content (a) x = 0.05 (red) and (b) x = 0.3 (blue) at 78 K. All spectra were area normalized. Hatched areas mark the overlap between donor (Cr2+) ions fluorescence and acceptor (Fe2+) ions absorption.

Download Full Size | PDF

Fe2+ ions fluorescence spectra are also modified depending on Mn content x with similar (to absorption spectra) shift of Fe2+ fluorescence maximum towards longer wavelengths with an increase of Mn content x as shown in Fig. 3. The fluorescence spectra were measured using a monochromator (Oriel 77250) together with a LN2-cooled mercury-cadmium-telluride (MCT) detector (Judson-Teledyne J15).

 figure: Fig. 3.

Fig. 3. Comparison of Fe2+ ions fluorescence spectra in the Zn1-xMnxSe:Cr2+,Fe2+ crystals with Mn content x = 0.05 (red) and x = 0.3 (blue) at 78 K.

Download Full Size | PDF

The fluorescence decay curve of Cr2+ ions measured at 78 K in the Zn1-xMnxSe:Cr2+ (x = 0.05) crystal was single exponential with a decay time of 5.3 µs as shown in Fig. 4 (pink hollow dots). In the presence of Fe2+ ions (in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal), the decay curve was observed to become non-exponential due to the nonradiative Cr2+ → Fe2+ ions energy transfer. It has to be noted that deviation from single exponential decay was not so pronounced as in the case of the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.3) crystal with about 2.5 times higher Fe2+ ions concentration [27].

 figure: Fig. 4.

Fig. 4. Decay curves of Cr2+ ions in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal without (pink hollow dots) and with (black hollow dots) Fe2+ ions co-doping measured at 78 K. Green line shows single exponential approximation.

Download Full Size | PDF

The efficiency of Cr2+ to Fe2+ (donor-acceptor) energy transfer (ηDA) can be directly evaluated from Fig. 4 by using equation ηDA = 1-<τCr-Fe>/τCr, where τCr is the Cr2+ ions lifetime in the sample without Fe2+ ions co-doping and <τCr-Fe> is the average Cr2+ ions lifetime in the co-doped sample [14,27]. The average lifetime can be obtained from the decay curve as <τCr-Fe> = $\mathop \smallint \nolimits_0^\infty {\textrm{I}_{\textrm{D}}}(t )\textrm{d}t/{\textrm{I}_{\textrm{o}}}$ and the energy transfer efficiency can be therefore expressed using following equation:

$${\eta _{\textrm{DA}}} = 1 - \frac{{\int\limits_0^\infty {{\textrm{I}_\textrm{N}}(t)\textrm{d}t} }}{{\int\limits_0^\infty {{\textrm{I}_\textrm{R}}(t)\textrm{d}t} }}$$
The calculated energy transfer efficiency in this case is 12% only. This value is much lower than in the case of the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.3) crystal where the Cr2+ → Fe2+ energy transfer efficiency of about 55% was presented [27].

Energy transfer function Ln(IR)-Ln(IN) plotted in $\sqrt t $ axis similar to [27] shows the first “ordered” exponential decay in the beginning (see inset in Fig. 5) which linearizes further for “disordered” Forster energy transfer stage. As follows from Fig. 5 the first exponential stage is much less pronounced compared to that in the case of the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.3) crystal with about 2.5 times higher Fe2+ ions concentration. The value of maximal decay rate Wmax obtained from the first “ordered” part was about 0.8×105 s−1 which is about an order of magnitude lower than in the case of the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.3). The average decay rate at the “disordered” energy transfer stage was calculated from the measured γ = 145 s−1/2 value to be WF = γF2 = 0.2×105 s−1. This value is again nearly an order of magnitude lower than that for the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.3) crystal [27] and again should be a result of 2.5 times lower acceptors (Fe2+ ions) concentration. It should be noted that Wmax to WF ratio in both cases is quite similar and it is close to 4. Moreover, critical time (when “ordered” stage is exchanged by “disordered one”) was also quite similar in both cases being close to 0.7 µs (see Fig. 5).

 figure: Fig. 5.

Fig. 5. Nonradiative Cr2+ → Fe2+ energy transfer function in $\sqrt t $ axis in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) and (x = 0.3) crystals at 78 K.

Download Full Size | PDF

Donor-acceptor energy transfer efficiency was estimated for the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal according to [30] as

$${\eta _{\textrm{DA}}} = \frac{{{W_\textrm{F}}}}{{{W_\textrm{F}} + \frac{1}{{{\tau _{\textrm{Cr}}}}}}}$$
For values of WF = 0.2×105 s−1 and τCr = 5.3 µs obtained above, the efficiency ηDA was calculated to be about 10%, corresponding well to the value calculated from the decay curve integrals (∼12%). The macroparameter of donor-acceptor nonradiative energy transfer value CDA in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal calculated for Fe2+ ions concentration of about 2×1018 cm−3 was ∼3.3×10−35 cm6s−1 which is not so far from that obtained in [27] for the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.3) crystal with about 2.5 times higher Fe2+ ions concentration (CDA ∼ 4.8×10−35 cm6s−1). We suggest that CDA parameter decrease should be a result of lower overlap integral of Cr2+ ions fluorescence with Fe2+ ions absorption in the case of crystal with lower Mn content x.

Serious advantage of the Zn1-xMnxSe (x = 0.05) crystal is a much lower nonradiative quenching of Fe2+ ions fluorescence resulting in sufficiently longer Fe2+ ions fluorescence lifetime especially at 78 K. In Fig. 6 measured decay curves of Fe2+ ions at 78 K are shown for two Mn contents x = 0.05 and 0.3. Both decay curves were observed to be double exponential (fits are shown in Fig. 6 by the dashed lines) with fluorescence lifetimes at initial and tail stages presented in this figure. As follows from Fig. 6 iron ions fluorescence lifetime is shortened about 4× at both stages when Mn content is increased from x = 0.05 to x = 0.3. Such longer iron fluorescence lifetime (63 µs at the tail which is quite similar to that for a ZnSe crystal ∼60 µs at ∼78 K [26,31,32]) for lower Mn content should result in higher quantum yield of Fe2+ ions fluorescence with respect to the Zn1-xMnxSe (x = 0.3) crystal. It should be noted that fluorescence decay times in Zn1-xMnxSe:Cr2+,Fe2+ crystals were measured and analyzed at 78 K and can differ from radiative lifetimes which were shown in [3133] to be shorter for temperatures below 14 K.

 figure: Fig. 6.

Fig. 6. Fluorescence decay curves of Fe2+ ions in the Zn1-xMnxSe:Fe2+ crystals with Mn content x = 0.05 (red) and x = 0.3 (blue) at 78 K.

Download Full Size | PDF

Pumping at ∼1.7 µm is quite efficient from the point of view of Cr2+ ions excitation as it well fits Cr2+ absorption maximum in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal (see Fig. 1). On the other hand, such short wavelength pumping should increase the quantum defect for Fe2+ ions excitation. For ordinarily used ∼2.9 µm pumping with Er3+ ions-based excitation sources, theoretical efficiency for generation at ∼4.1 µm as high as 70% could be achieved. It should be noted that Fe2+ ions lasing with a slope efficiency up to 59% was demonstrated in the ZnSe crystal recently [17]. For 1.7 µm pumping theoretical limit decreases more than twice down to ∼30%. Though ∼1.7 µm excitation through the Cr2+ → Fe2+ energy transfer simplifies the pump scheme and enlarge the number of commercially available compact pump sources allowing to obtain cheap and compact Fe2+ ions based tunable lasers.

3. Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) laser results

The Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) laser properties were investigated under two different pumping sources around ∼1.7 µm. The first pumping source was based on a Q-switched YLF:Er laser and the second source was a commercial laser diode (LD).

3.1 Pumping by a 1.73 µm Q-switched YLF:Er laser

In the first experiments Fe2+ ions lasing under 1.73 µm Q-switched YLF:Er laser excitation similar to [27] was tested. The laser cavity and mirrors were the same as described in [27]. The pumping beam was focused by a 150 mm CaF2 lens. Both the pumping flat mirror (PM) and concave (r = 500 mm) output coupler (OC) were placed outside the cryostat and the cavity length was ∼10 cm. The cryostat was equipped by CaF2 windows with no antireflective coating. Pumping mirror transmittance at 1.73 µm was ∼95% and reflectivities at 2.4 µm and 4.2 µm were 8.5% and >98%, respectively. The output coupler reflectivity at the same wavelengths were 92% and 94%, respectively. The YLF:Er laser generated 200 ns pulses with the repetition rate of 1 Hz. The experimental arrangement is presented in Fig. 7(a) with a detail of YLF:Er pumping in Fig. 7(b).

 figure: Fig. 7.

Fig. 7. (a) Experimental arrangement of the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) laser. Pumping sources: (b) YLF:Er pumping, (c) laser diode pumping. PM – pumping mirror, OC – output coupler, HR – high reflectivity mirror at ∼1.7 µm, PC – Pockels cell.

Download Full Size | PDF

Unlike in the case of the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.3) crystal used in [27], relatively low Cr2+ → Fe2+ energy transfer efficiency in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal resulted in simultaneous Cr2+ and Fe2+ ions lasing at 78 K (see Fig. 8(a) as an example).

 figure: Fig. 8.

Fig. 8. Temporal profiles of the Zn1-xMnxSe:Cr2+,Fe2+ (a) x = 0.05 and (b) x = 0.3 [27] laser radiation at 78 K for the output mirror reflectivity of 94% in intracavity pumping mode operation. Inset in (a) presents the Fe2+ laser beam spatial profile.

Download Full Size | PDF

Fe2+ ions intracavity pumping mode (most probably followed by the Cr2+ → Fe2+ energy transfer stage) resulting in about 6 µs long Fe2+ ions oscillation pulse – Fig. 8(a) was realized in this case with the highest output energy of 83 µJ, which corresponds to about 1% of optical-to-optical efficiency (absorbed pumping energy was ∼7 mJ).

Oscillation pulse of the Fe2+ ions in the x = 0.3 sample for the case of the Cr2+ → Fe2+ intracavity pumping is shown in Fig. 8(b). Unfortunately, pure Cr2+ → Fe2+ energy transfer mode like in the x = 0.3 sample [27] was not achieved for the x = 0.05 sample. The oscillation spectra of the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) laser were measured by the monochromator Oriel 77250 described above and Vigo PVI-4TE-6 photodetector. The laser generated in a temperature range from 78 K up to 200 K and the generated spectra examples are presented in Fig. 9. A drop in the oscillation spectra around 4.25 µm is caused by the CO2 absorption because the cavity mirrors were placed outside the evacuated cryostat. As it could be seen from this figure, the central oscillation wavelength red-shift was observed to be similar to direct 2.94 µm Fe2+ ions excitation [26].

 figure: Fig. 9.

Fig. 9. Oscillation spectra of the Zn1-xMnxSe:Cr2+,Fe2+ x = 0.05 laser under 1.73 µm Q-switched YLF:Er laser pumping in a temperature range from 78 K up to 200 K.

Download Full Size | PDF

3.2 Pumping by a 1.71 µm laser diode

Oscillation properties of the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal were also tested in more interesting case of commercial 1.71 µm laser diode (LD) pumping. This LD generated maximum CW power of 35 W. This type of optical excitation provides much lower pumping rate compared to Q-switched laser pumping. Also, a specific LD driver was characterized by relatively long build up time (∼2 ms) of the excitation pulse which resulted in a noticeable delay between the laser diode pump and Fe2+ ions oscillation pulse start. Therefore, an optical chopper (Thorlabs MC2000B) was used to get ∼1 ms (FWHM) pumping pulses at 10 Hz from CW operation of the LD. It should be noted that for available LD pumping levels, no Cr2+ ions lasing was observed for the same mirror set as in the case of the YLF:Er Q-switched laser. The experimental setup of the LD pumping system is shown in Fig. 7(c).

Oscillation pulse of Fe2+ ions was measured by the MCT detector behind a bandpass filter (Thorlabs FB4250-250). Generation was achieved in the temperature range from 78 K up to 110 K. Typical temporal profiles of the pump and Fe2+ ions laser oscillation pulses at 78 K are shown in Fig. 10. This is the first demonstration, to our best knowledge, of Fe2+ ions laser generation under direct commercial 1.71 µm laser diode excitation.

 figure: Fig. 10.

Fig. 10. Oscillogram of the 1.71 µm pump pulse (green curve) and the 4.19 µm Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) laser output pulse at 78 K (red curve).

Download Full Size | PDF

The oscillation spectra were measured again by the monochromator and MCT described above. The Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) laser oscillation spectra for temperatures of 78 K and 100 K are shown in Fig. 11. The oscillation spectrum maximum of the Fe2+ ions in the crystal was observed to be slightly shifting towards longer wavelengths with temperature increase. The output spectra were quite strongly influenced by CO2 absorption at ∼4.25 µm. The central output laser wavelength was ∼4190 nm at 78 K.

 figure: Fig. 11.

Fig. 11. Oscillation spectra of the Zn1-xMnxSe:Cr2+,Fe2+ x = 0.05 laser under 1.71 µm laser diode pumping for temperatures of 78 K and 100 K.

Download Full Size | PDF

The mean output power was measured by a power probe (Coherent PM3) behind the LP3000 and LP3750 long-pass filters (Spectrogon, transmittance of each filter ∼92% @ 4.2 µm). Laser mean output power as a function of absorbed pump power for the active medium temperature of 78 K is presented in Fig. 12. The slope efficiency was η ≈ 2.3%. The maximum mean output power was 4.1 mW (single pulse energy ∼0.41 mJ) for the absorbed pump power of ∼240 mW corresponding to the optical-to-optical efficiency of 1.6%. The output beam spatial profile was measured by a pyroelectric camera (Spiricon Pyrocam III) behind a long-pass filter LP4095 (Spectrogon). The output beam shape was observed to be distorted compared to fundamental Gaussian mode most likely due to high active crystal thermal loading due to high Cr2+ ions absorption.

 figure: Fig. 12.

Fig. 12. Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) mean laser output power as a function of the absorbed laser-diode pump power. Inset presents the beam spatial profile for the maximal output power of 4.1 mW.

Download Full Size | PDF

4. Summary

To conclude, to our best knowledge, the Fe2+ ions lasing in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) single crystal under direct commercial 1.71 µm laser diode and YLF:Er Q-switched laser excitation was demonstrated for the first time. A nonradiative Cr2+ → Fe2+ ions energy transfer parameters in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal were determined and compared to those previously shown for x = 0.3 crystal [28]. Five times lower (compared to x = 0.3 crystal) Cr2+ → Fe2+ energy transfer efficiency of about 10% in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal was demonstrated for used Cr2+ and Fe2+ ions concentrations which seems to result from an order of magnitude lower energy transfer rate at both the first “ordered” and second “disordered” stages due to 2.5 times lower Fe2+ acceptor ions concentration.

Short pulse excitation by a Q-switched YLF:Er laser (1.73 µm, 200 ns, 1 Hz) has demonstrated only intracavity Cr2+ → Fe2+ ions pumping mode (most probably followed by the Cr2+ → Fe2+ ions energy transfer stage) resulting in the Fe2+ ions oscillations. The oscillations were achieved in a temperature range from 78 K up to 200 K and the central oscillation wavelength increased from 4.1 µm to 4.4 µm depending on the active medium temperature. The highest output energy of 83 µJ and optical-to-optical efficiency of about 1% at 78 K were achieved.

Under 1.71 µm commercial laser diode pumping (1 ms, 10 Hz), Fe2+ ions excitation through the Cr2+ → Fe2+ ions energy transfer mode was realized. In this case the maximum mean laser output power of 4.1 mW and slope efficiency of 2.3% with respect to the absorbed power were obtained at 78 K. Laser was observed to operate within a temperature range from 78 K up to 110 K with central oscillation wavelength shift from 4.19 µm to 4.23 µm.

These results seem to be a good demonstration of a future possibility to develop a compact diode-pumped mid-infrared coherent laser source generating in the range of 4.1–4.3 µm. Further optimization of absolute Cr2+ and Fe2+ ions concentrations as well as Cr2+ to Fe2+ ions concentration ratio should allow to optimize energy transfer process and increase further the Fe2+ ions oscillation efficiency.

Funding

Grantová Agentura České Republiky (18-11954S); Russian Academy of Sciences (Program 5.11).

Disclosures

The authors declare no conflicts of interest.

References

1. M. Ebrahim-Zadeh and I. T. Sorokina, Mid-Infrared Coherent Sources and Applications (Springer, 2005).

2. A. Antognini, F. D. Amaro, F. Biraben, J. M. R. Cardoso, C. A. N. Conde, D. S. Covita, A. Dax, S. Dhawan, L. M. P. Fernandes, T. W. Hänsch, V. W. Hughes, O. Huot, P. Indelicato, L. Julien, P. E. Knowles, F. Kottmann, Y.-W. Liu, J. A. M. Lopes, L. Ludhova, C. M. B. Monteiro, F. Mulhauser, F. Nez, B. N. Perry, R. Pohl, P. Rabinowitz, J. M. F. dos Santos, L. A. Schaller, C. Schwob, D. Taqqu, and J. F. C. A. Veloso, “Powerful fast triggerable 6 µm laser for the muonic hydrogen 2S-Lamb shift experiment,” Opt. Commun. 253(4-6), 362–374 (2005). [CrossRef]  

3. A. A. Ionin, Y. M. Klimachev, A. A. Kotkov, A. Y. Kozlov, L. V. Seleznev, and D. V. Sinitsyn, “Carbon monoxide laser emitting nanosecond pulses with 10 MHz repetition rate,” Opt. Commun. 282(2), 294–299 (2009). [CrossRef]  

4. C. A. Brau, “Free-Electron Lasers,” Science 239(4844), 1115–1121 (1988). [CrossRef]  

5. V. Petrov, “Frequency down-conversion of solid-state laser sources to the mid-infrared spectral range using non-oxide nonlinear crystals,” Prog. Quantum Electron. 42, 1–106 (2015). [CrossRef]  

6. H. Pires, M. Baudisch, D. Sanchez, M. Hemmer, and J. Biegert, “Ultrashort pulse generation in the mid-IR,” Prog. Quantum Electron. 43, 1–30 (2015). [CrossRef]  

7. A. G. Ciriolo, M. Negro, M. Devetta, E. Cinquanta, D. Faccialà, A. Pusala, S. De Silvestri, S. Stagira, and C. Vozzi, “Optical parametric amplification techniques for the generation of high-energy few-optical-cycles IR pulses for strong field applications,” Appl. Sci. 7(3), 265 (2017). [CrossRef]  

8. E. Migal, A. Pushkin, B. Bravy, V. Gordienko, N. Minaev, A. Sirotkin, and F. Potemkin, “3.5 mJ 150 fs Fe:ZnSe hybrid mid-IR femtosecond laser at 4.4 µm for driving extreme nonlinear optics,” Opt. Lett. 44(10), 2550–2553 (2019). [CrossRef]  

9. R. Pecharromán-Gallego, “An overview on quantum cascade lasers: origins and development,” in Quantum Cascade Lasers (Intech,2017).

10. S. Slivken, D. Wu, and M. Razeghi, “Surface emitting, tunable, mid-infrared laser with high output power and stable output beam,” Sci. Rep. 9(1), 549 (2019). [CrossRef]  

11. V. V. Fedorov, S. B. Mirov, A. Gallian, D. V. Badikov, M. P. Frolov, Y. V. Korostelin, V. I. Kozlovsky, A. I. Landman, Y. P. Podmar’kov, V. A. Akimov, and A. A. Voronov, “3.77–5.05 µm tunable solid-state lasers based on Fe2+-doped ZnSe crystals operating at low and room temperatures,” IEEE J. Quantum Electron. 42(9), 907–917 (2006). [CrossRef]  

12. J. J. Adams, C. Bibeau, R. H. Page, D. M. Krol, L. H. Furu, and S. A. Payne, “4.0–4.5 µm lasing of Fe:ZnSe below 180 K, a new mid-infrared laser material,” Opt. Lett. 24(23), 1720–1722 (1999). [CrossRef]  

13. V. A. Akimov, A. A. Voronov, V. I. Kozlovsky, Y. V. Korostelin, A. I. Landman, Y. P. Podmar’kov, and M. P. Frolov, “Efficient IR Fe:ZnSe laser continuously tunable in the spectral range from 3.77 to 4.40 µm,” Quantum Electron. 34(10), 912–914 (2004). [CrossRef]  

14. S. B. Mirov, V. V. Fedorov, D. Martyshkin, I. Moskalev, M. Mirov, and S. Vasilyev, “Progress in mid-IR lasers based on Cr and Fe-doped II–VI chalcogenides,” IEEE J. Sel. Top. Quantum Electron. 21(1), 292–310 (2015). [CrossRef]  

15. M. E. Doroshenko, T. Basiev, P. Koranda, H. Jelínková, M. Němec, J. Šulc, M. Čech, V. K. Komar, A. S. Gerasimenko, V. V. Badikov, and D. V. Badikov, “Bulk Fe:ZnSe laser gain-switched by the Q-switched Er:YAG laser,” Proc. SPIE 7193, 71931K (2009). [CrossRef]  

16. S. D. Velikanov, V. P. Danilov, N. G. Zakharov, N. N. Il’ichev, S. Y. Kazantsev, V. P. Kalinushkin, I. G. Kononov, A. S. Nasibov, M. I. Studenikin, P. P. Pashinin, K. N. Firsov, P. V. Shapkin, and V. V. Shchurov, “Fe2+:ZnSe laser pumped by a nonchain electric-discharge HF laser at room temperature,” Quantum Electron. 44(2), 141–144 (2014). [CrossRef]  

17. A. V. Pushkin, E. A. Migal, H. Uehara, K. Goya, S. Tokita, M. P. Frolov, Y. V. Korostelin, V. I. Kozlovsky, Y. K. Skasyrsky, and F. V. Potemkin, “Compact, highly efficient, 2.1 W continuous-wave mid-infrared Fe:ZnSe coherent source, pumped by an Er:ZBLAN fiber laser,” Opt. Lett. 43(24), 5941–5944 (2018). [CrossRef]  

18. S. Vasilyev, I. Moskalev, M. Mirov, V. Smolski, D. Martyshkin, V. V. Fedorov, S. B. Mirov, and V. Gapontsev, “Progress in Cr and Fe doped ZnS/Se mid-IR CW and femtosecond lasers,” Proc. SPIE 10193, 101930U (2017). [CrossRef]  

19. N. Myoung, D. V. Martyshkin, V. V. Fedorov, and S. B. Mirov, “Mid-IR lasing of iron–cobalt co-doped ZnS(Se) crystals via Co–Fe energy transfer,” J. Lumin. 133, 257–261 (2013). [CrossRef]  

20. J. Peppers, N. Myoung, V. V. Fedorov, and S. B. Mirov, “Mid-IR laser oscillation via energy transfer in the Co:Fe:ZnS/Se co-doped crystals,” Proc. SPIE 8235, 823503 (2012). [CrossRef]  

21. J. Kernal, V. V. Fedorov, A. Gallian, S. B. Mirov, and V. V. Badikov, “3.9–4.8 µm gain-switched lasing of Fe:ZnSe at room temperature,” Opt. Express 13(26), 10608–10615 (2005). [CrossRef]  

22. J. Peppers, V. V. Fedorov, and S. B. Mirov, “Mid-IR photoluminescence of Fe2+ and Cr2+ ions in ZnSe crystal under excitation in charge transfer bands,” Opt. Express 23(4), 4406–4414 (2015). [CrossRef]  

23. X. Wang, Z. Chen, L. Zhang, B. Jiang, M. Xu, J. Hong, Y. Wang, P. Zhang, L. Zhang, and Y. Hang, “Preparation, spectroscopic characterization and energy transfer investigation of iron-chromium diffusion co-doped ZnSe for mid-IR laser applications,” Opt. Mater. 54, 234–237 (2016). [CrossRef]  

24. V. V. Fedorov, T. Carlson, and S. B. Mirov, “Energy transfer in iron-chromium co-doped ZnSe middle-infrared laser crystals,” Opt. Mater. Express 9(5), 2340–2347 (2019). [CrossRef]  

25. X. Wang, Z. Chen, L. Zhang, M. Xu, G. Chen, B. Jiang, C. Ke, L. Zhang, and Y. Hang, “Charge state and energy transfer investigation of iron-chromium co-doped ZnS polycrystalline prepared by step-temperature diffusion for mid-infrared laser applications,” J. Alloys Compd. 695, 3767–3771 (2017). [CrossRef]  

26. M. E. Doroshenko, H. Jelínková, V. V. Osiko, M. Jelínek, D. Vyhlídal, J. Šulc, M. Němec, N. O. Kovalenko, and A. S. Gerasimenko, “Fe:ZnMnSe laser active material at 1300–1307 K: Spectroscopic properties and laser generation at 4.2–5.0 µm,” J. Lumin. 192, 1300–1307 (2017). [CrossRef]  

27. M. E. Doroshenko, H. Jelínková, A. Říha, M. Jelínek, M. Němec, N. O. Kovalenko, and A. S. Gerasimenko, “Mid-IR (4.4 µm) Zn1−xMnxSe:Cr2+,Fe2+ (x = 0.3) laser pumped by 1.7 µm laser using Cr2+-Fe2+ energy transfer,” Opt. Lett. 44(11), 2724–2727 (2019). [CrossRef]  

28. M. E. Doroshenko, M. Jelínek, A. Říha, J. Šulc, H. Jelínková, V. Kubeček, N. O. Kovalenko, and A. S. Gerasimenko, “Long-pulse 4.4–4.6 µm laser oscillations of Fe2+ ions in a Zn1-xMnxSe (x = 0.3) crystal pumped by a 1940nm Tm:fiber laser through Cr2+ → Fe2+ energy transfer,” Opt. Lett. 44(21), 5334–5337 (2019). [CrossRef]  

29. M. Němec, M. E. Doroshenko, H. Jelínková, R. Švejkar, J. Šulc, and V. K. Komar, “Wavelength tunability of diode-pumped Cr-doped ZnMnSe laser,” Laser Physics workshop LPHYS’17, paper P.S1.2 (2017).

30. O. K. Alimov, T. T. Basiev, M. K. Ashurov, E. O. Kirpichenkova, and V. B. Murav’ev, Selective Laser Spectroscopy of Activated Crystals and Glasses, Proceedings of GPI–9, Chapter II (Nova Science Publishers, 1990).

31. N. Myoung, V. V. Fedorov, S. B. Mirov, and L. E. Wenger, “Temperature and concentration quenching of mid-IR photoluminescence in iron doped ZnSe and ZnS laser crystals,” J. Lumin. 132(3), 600–606 (2012). [CrossRef]  

32. J. W. Evans, T. R. Harris, B. Rami Reddy, K. L. Schepler, and P. A. Berry, “Optical spectroscopy and modeling of Fe2+ ions in zinc selenide,” J. Lumin. 188, 541–550 (2017). [CrossRef]  

33. J. W. Evans, R. W. Stites, and T. R. Harris, “Increasing the performance of an Fe:ZnSe laser using a hot isostatic press,” Opt. Mater. Express 7(12), 4296–4303 (2017). [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 (12)

Fig. 1.
Fig. 1. Absorption spectra of Cr2+ and Fe2+ ions in the Zn1-xMnxSe:Cr2+,Fe2+ crystals with Mn content x = 0.05 (red) and x = 0.3 (blue) at 78 K.
Fig. 2.
Fig. 2. Fluorescence spectra of Cr2+ ions and absorption spectra of Fe2+ ions in the Zn1-xMnxSe:Cr2+,Fe2+ crystals with Mn content (a) x = 0.05 (red) and (b) x = 0.3 (blue) at 78 K. All spectra were area normalized. Hatched areas mark the overlap between donor (Cr2+) ions fluorescence and acceptor (Fe2+) ions absorption.
Fig. 3.
Fig. 3. Comparison of Fe2+ ions fluorescence spectra in the Zn1-xMnxSe:Cr2+,Fe2+ crystals with Mn content x = 0.05 (red) and x = 0.3 (blue) at 78 K.
Fig. 4.
Fig. 4. Decay curves of Cr2+ ions in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) crystal without (pink hollow dots) and with (black hollow dots) Fe2+ ions co-doping measured at 78 K. Green line shows single exponential approximation.
Fig. 5.
Fig. 5. Nonradiative Cr2+ → Fe2+ energy transfer function in $\sqrt t $ axis in the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) and (x = 0.3) crystals at 78 K.
Fig. 6.
Fig. 6. Fluorescence decay curves of Fe2+ ions in the Zn1-xMnxSe:Fe2+ crystals with Mn content x = 0.05 (red) and x = 0.3 (blue) at 78 K.
Fig. 7.
Fig. 7. (a) Experimental arrangement of the Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) laser. Pumping sources: (b) YLF:Er pumping, (c) laser diode pumping. PM – pumping mirror, OC – output coupler, HR – high reflectivity mirror at ∼1.7 µm, PC – Pockels cell.
Fig. 8.
Fig. 8. Temporal profiles of the Zn1-xMnxSe:Cr2+,Fe2+ (a) x = 0.05 and (b) x = 0.3 [27] laser radiation at 78 K for the output mirror reflectivity of 94% in intracavity pumping mode operation. Inset in (a) presents the Fe2+ laser beam spatial profile.
Fig. 9.
Fig. 9. Oscillation spectra of the Zn1-xMnxSe:Cr2+,Fe2+ x = 0.05 laser under 1.73 µm Q-switched YLF:Er laser pumping in a temperature range from 78 K up to 200 K.
Fig. 10.
Fig. 10. Oscillogram of the 1.71 µm pump pulse (green curve) and the 4.19 µm Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) laser output pulse at 78 K (red curve).
Fig. 11.
Fig. 11. Oscillation spectra of the Zn1-xMnxSe:Cr2+,Fe2+ x = 0.05 laser under 1.71 µm laser diode pumping for temperatures of 78 K and 100 K.
Fig. 12.
Fig. 12. Zn1-xMnxSe:Cr2+,Fe2+ (x = 0.05) mean laser output power as a function of the absorbed laser-diode pump power. Inset presents the beam spatial profile for the maximal output power of 4.1 mW.

Equations (2)

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

η DA = 1 0 I N ( t ) d t 0 I R ( t ) d t
η DA = W F W F + 1 τ Cr
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.