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

Nd:YAG optical electronic nonlinearity and energy transfer upconversion studied by the Z-scan technique

Open Access Open Access

Abstract

A procedure to obtain the Energy Transfer Upconversion parameter, γ, for Nd3+:YAG by using the Z-scan technique is presented in this paper. It was found γ = (2.0 ± 0.3) × 10−16 cm3/s, from the open aperture transmittance. Also obtained was the nonlinear refractive index, the polarizability difference and absorption cross-section difference between excited and ground state at 808.7 nm.

© 2015 Optical Society of America

1. Introduction

Energy transfer upconversion (ETU) process is a well known mechanism responsible by heating in laser crystals under high intensity pump conditions [1, 2]. This deleterious effect enhances thermal lensing and decreases the population inversion, thus limiting the performance of solid state lasers. Therefore, the knowledge of ETU is very important in the design of solid-state lasers. For instance, Chen demonstrated that ETU has a critical effect in the choice of pump-to-mode size ratio of a laser cavity [3]. Some methods has been used in order to characterize the ETU process such as fluorescence decay behavior [1, 2], different thermal lens techniques [4, 5] and analysis of the laser behavior. However, the emission dynamics of Nd:YAG is quite complex and departures from single exponential decay even at quite low concentrations and low pump intensity (negligible ETU effect) [5]. Consequently, the ETU effect can only be detected at very high excitation levels, higher than those usually found in operating lasers. Therefore, taking advantage on the dependence of the transmitted light through the sample with the ETU process, the Z-scan technique can be applied as an alternative approach in the study of the energy transfer mechanisms. Z-scan is the most popular technique for the characterization of nonlinear refraction and absorption behavior in several classes of materials [6]. The time resolved Z-scan technique was introduced to study materials presenting slow nonlinearities (response times > 10−4 s) using cw lasers [7]. This technique is able to monitor the transient behavior of the excited state population [8]. It is an attractive method to quantitatively study the ETU effect due to its simplicity, sensitivity and accuracy [9, 10]. Although the ETU effect is known to be important in Nd:YAG, the literature data is scarce with conflicting results [1–3]. The aim of this article is to study the ETU process in Nd:YAG crystals and in its excited state population dynamics. Moreover, the nonlinear properties related to refractive index change and absorption changes were also determined.

2. Theoretical background

In ion doped laser materials (transparent crystals and glasses), it is well known that a refractive index change appears with the variation of the excited state population or, consequently, the gain. This effect is due to the fact that the ion excited state polarizability, αpex, is different from that the ground state one, αpg. For Nd3+ ion, the ground, 4I9/2, and excited, 4F3/2, states of Nd3+ ions will be hereafter referred as g and ex, respectively. The contribution of the ion to the total susceptibility is given byχion=αpgNg+αpexNex [11], with Ng (Nex) the ground (excited) ion population density. Here, the usual case where the system has a main excited metastable state that traps all excited population is considered. Therefore, in the stationary regime the total ion concentration is given by Nt ~Nex + Ng and consequently χion=αpgNt+ΔαpNex where Δαp = αpex - αpg . Under cw excitation, and in the stationary regime, the excited state population is given by Nex ~NtI/Is for I/Is<<1, where IS=hν/τ0ηpσg is the pump saturation intensity, is the photon energy, το is the excited state lifetime, σg is the ground state absorption cross section and ηp is the pump efficiency (i.e. the fraction of absorbed pump photons that result in excitation of the emitting level). Lower than unity ηp may be caused by the presence of impurities or “dead sites”. However, there is no clear evidence of these effects, so it is usually assumed ηp ~1 [12]. Therefore, in first order the electronic athermal refractive index change is proportional to the intensity asΔn=n2I, so that [11]:

n2=NtIs(2πn0fL2Δαpiλ4πΔσ)
where n0 is the real part of the linear refractive index, fL = (n02 + 2)/3 is the Lorenz local field correction factor and the imaginary part of n2 is related to Δσ = σex - σg, where σex is the excited state absorption cross section. So the system behaves like a saturable absorber if Δσ < 0 or a reverse saturable absorber otherwise. Equation (1) is an important expression in this paper because it relates n2 to Δαp, which is a microscopic parameter so this connection depends on the local field factor. The time-resolved Z-scan technique can be used to study these refractive index changes in ion-doped solids. The reliability of this technique was demonstrated in different papers [8, 13–15]. It consists in monitoring the changes in the far-field intensity pattern while the nonlinear sample is translated through the focal plane of a focused Gaussian beam. By monitoring the transmittance change as the sample is scanned through the laser beam focal plane, one is able to determine the complex refractive index change Δn = n2I. The response time of the nonlinear effect of ion doped solids is governed by the population temporal evolution, which is related to excited state lifetime of the metastable level, ~250 μs for Nd:YAG. For materials with slow nonlinear response time, the time-resolved Z-scan method [16] has been used successfully with a cw laser modulated by a mechanical chopper. As usually done in the Z-scan method, the beam was split after the sample, in order to detect simultaneously both open and closed aperture signals (with aperture factor S1 = 100% and S2 = 50%, respectively, as defined in Ref [6]. Details on the experimental apparatus and method can be found elsewhere [14].

On the other hand, it is already known that the ETU in Nd3+ systems involves the interaction of two excited ions in the 4F3/2 metastable laser level. Here one ion returns to the 4I11/2 (and/or 4I13/2) level while the other is promoted to the higher-lying excited level followed by a multiphonon decay back to the 4F3/2 level [17, 18]. The ETU process in doped solids is satisfactorily described through standard population rate equation using the ETU rate given by Wup = γNex, where γ is the upconversion parameter [1]:

dNexdt=Sτ0Nt(1+S)τ0NexγNex2
where τ0 is the lifetime and S = I/IS the saturation parameter. The stationary solution of Eq. (2) is:
Nex=Nt(1+S)2+4βS(1+S)2β
where β = γτ0Nt is a dimensionless parameter related to the strength of the upconversion process. In the limit of β → 0 (no ETU) Nex = NtS/(1 + S), as expected. On the other hand, the expansion of Eq. (3), for small S, results in Nex ~Nt[S – (1 + β)S2]. Therefore, in first order in S, Nex is not affected by ETU. The Eq. (3) also shows that Nex is strongly depleted by the ETU process since the required pump intensity to reach the saturation condition (Nex = Ng = ½ for S = 1), i.e. the saturation intensity, is now I = Is(1 + β/2). This increment (factor β/2) appears because part of the pumped excited state population is transferred to ground state nonradiatively due to ETU. This effect should be more dramatic in stoichiometric systems, were the emitter ion is part of the matrix so that β >>1 [17].

3. Experimental

In this work the used sample was a 2.1 mm thick Nd3+:YAG crystal, pumped in resonance with 4I9/24F5/2 transition (808.7 nm), from a cw Ti:Sapphire laser. The linear absorption coefficient was found as αabs = (7.15 ± 0.02) cm−1. The Z-scan experiments were performed using a 10 cm focal length lens to focalize the laser beam, with a w = (35 ± 1) μm beam waist. The chopper frequency was adjusted to its maximum value (820 Hz) in order to minimize the aperture time (~10 µs). The time-resolved Z-scan measurements were performed at ti = 20 μs and tf = 600 μs (initial and final times for data acquisition). More details of the experimental set-up can be found if Refs [7, 14, 16]. For the transient measurements the sample was positioned at z = 5 mm, which corresponds to the peak of the purely refractive curve (closed/open ratio).

In ion doped materials, it is well known that both thermal and the electronic effects should contribute to the Z-scan the signal. The magnitude of the TL effect (in the cw regime) is proportional to phase parameter θ [14]:

θ=φPabsλKdsdT
where φ is the fraction of the absorbed pump power (Pabs) that is converted into heat, K is the thermal conductivity, λ the wavelength and ds/dT the optical path. The electronic contribution is proportional to the parameter ΔΦο, given by:
ΔΦo=2πλLeff n2,2Pπw2
Where n2, is the real part of n2, Leff=[1exp(αabsL)]/αabs is the effective length of the sample, L is the actual length of the sample. Therefore, in order analyze to influence of the thermal effect on the Z-scan results, it is interesting the magnitude of the dimensionless parameter:
θΔΦo=φ4KdsdTw2(n2/αabs)
This parameter θ /ΔΦo depends on material thermo-optical and electronic parameters and w. Our experiment we expect to have ΔΦo >> θ due to the relatively high thermal conductivity of YAG compared to glasses and fluoride crystals [14]. In this situation, the total Z-scan signal can be calculated considering the superposition of thermal and electronic effects [19].

The characteristic Thermal Lens response time is related to the radial heat diffusion, and is given by tc = wo2/4D, where D is the thermal diffusivity. In our experiment tc ~70 μs, so the response time of the thermal lens is typically smaller than the electronic population lens (τo = 240 μs).

4. Results and discussion

Firstly, the results obtained in the low intensity regime, I<<Is, where saturation effects are negligible, will be analyzed. This corresponds to the regime of validity of the standard Z-scan theory and of Eq. (1) since it assumes that Nex ~NtI/Is. Fig. 1 shows typical Z-scan curves obtained at 808.7 nm.

 figure: Fig. 1

Fig. 1 – Time resolved Z-scan measurement for (a) closed aperture, (b) open aperture and (c) the ratio closed/open at λ = 808.7 nm, chopper frequency f = 820 Hz, P = 120 mW and ω0 = 35μm.

Download Full Size | PDF

Since for Nd3+:YAG the magnitude of the nonlinear absorption is important, we normalized the closed-aperture signal (Fig. 1(a)), dividing it by the open aperture data (Fig. 1(b)), as is typically done in the Z-scan technique [6]. Therefore, Fig. 1(c) presents the purely refractive Z-scan curve. The open aperture curve presented an increase of the transmitted intensity at the focus position which indicates that absorption saturation is the predominant effect, Fig. 1(b).

Transient measurements were also performed in the low intensity regime in order to further investigate the origin of the measured refractive index changes. In these experiments, the sample was placed at the peak of the Z-scan curve and the transmittance transient signals in both detectors were recorded, Fig. 2.

 figure: Fig. 2

Fig. 2 – Transient transmittance obtained from both closed, 50% transmittance, and opened, 100% transmittance, Z-scan apertures. The solid lines are exponential fits.

Download Full Size | PDF

Both curves where fit by single exponentials, in good agreement with the fluorescence lifetime of this sample, τo = (240 ± 10) μs. In three level systems like ruby, this behavior would indicate a predominance of the electronic contribution compared to the thermal contribution. However, in Nd3+ doped materials, the heat released has a fast and a slow component. The fast one is due to the nonradiative transition 4F5/24F3/2, so that its dynamics is given by the typical heat diffusion time (tc = w2/4D), where D is the thermal diffusivity. The slow component is due to the transitions after the radiative decay, the cascade decay from the terminal levels (4I15/2, 4I13/2 4I11/2) to the ground state (4I9/2). In our experiment tc << τo, so the slow heat component has a dynamics similar to the one of the excited state population. In Nd:YAG, ~67% of the heat is released in the slow process and 33% in the fast one. Using the value (φK−1ds/dT) = 2.7 × 10−6 cm/W from Ref [20]. in Eq. (6) we estimated θ/ΔΦο ∼10%. Moreover, Fig. 1(c) presents a peak-to- valley distance ΔZp–v ~1.7zc (with zc = πw2) corroborating our assumption that the electronic contribution is the dominant one. In the cw regime, the thermal lens effect is known to present ΔZp–v, ~3.4zc, as a consequence of heat diffusion [19, 21]. Considering the superposition of thermal and electronic effects [19] we estimate that ~14% of the Z-scan signal is due to the thermal effect. Therefore, the electronic component of n2,=(4.4±0.2)×109 cm2/W. The imaginary part, n2,, is not affected by the thermal effect, and is obtained from the theoretical fitting of the Fig. 1(b) as n2,,=(2.4± 0.2)  109 cm2/W.

In principle, the values of Nt and Is should be known in order to obtain Δαp and Δσ from n2 through Eq. (1). However, the reported σg values in the literature varies in the range of (4 – 10) × 10−20 cm2 (at 808nm) [2, 22–24], probably due to the uncertainty of concentration determination, since αabs = Ntσg. This difficulty can be eliminated by noticing that Nt/Is = αabsτ0/hν, where both αabs and τ0 can be obtained with relative good precision. Here it was assumed unity pump efficiency (ηp ~1), for our Nd:YAG sample. This assumption is based on previous photothermal measurements performed on this sample that indicates no presence of defects that could affect the pump efficiency [12, 24]. The Z-scan allow the determination of both Δαp and Δσ by Eq. (1). Thus, for λ = 808.7nm, Δαp = (5.5 ± 1) × 10−26 cm3 and Δσ = (−5.1 ± 0.5) × 10−20 cm2 were found. This Δαp value is comparable with 4.9 × 10−26 cm3 obtained from four-wave-mixing experiments [25], 4.0 × 10−26 cm3 from interferometric measurements [26] and Δαp = (3.0 ± 0.2) × 10−26 cm3 by both interferometry and transient grating (λ = 633nm) [27]. In principle, we could expect some dispersion in Δαp(ω) due to the superposition of resonant and nonresonant contributions. In our experiment, the laser is tuned at line-center of an absorption transition (808.7nm), so the resonant contribution vanishes as demonstrated in pump-probe n2 spectroscopy measurements [15]. For instance, in the case of ruby (Al2O3:Cr3+), is was observed by Z-scan measurements that in the 543 – 457 nm range, Δαp(ω) varies only between 1.7 – 1.9 (10−25 cm3) and this variation is within the experimental uncertainty [28]. Similarly, the nonresonant contribution to Δαp in Nd:YAG is expected to be the most important one, consequently Δαp(ω) should be nearly constant in the 808 – 515 nm range [23].

The Δσ data is related to Nd:YAG excited state absorption (ESA) spectra studied in detail by Guyot et al. [2] and Kuck et al. [22]. Both papers indicate that at 808 nm, σex is at least one order of magnitude smaller than σg. From the spectra of Guyot, we estimated σex ~0.5 × 10−20 cm2 and Δσ = −5.5 × 10−20 cm2, in very good agreement with the value we obtained from n2.

Let us now consider the results for the pump saturation regime, where I is comparable to Is so both saturation and upconversion processes should be taken into account. Firstly, applying the Beer law to the beam crossing a sample, of thickness L, the transmittance can be written as T~T0(1- ΔσLNex), for NexΔσ <<1. In this expression T0 = exp(abs L) is the linear transmittance, found for Nex = 0. Therefore, the amplitude of the open aperture Z-scan signal is given by:

ΔT=ΔσLeffNex.

Figure 3 shows ΔT, obtained for different incident intensities, as a function of average radial intensity, <I(r)> = Io/2, where I0 is the on-axis intensity of the Gaussian beam profile I(r) = Io.exp(−2r2/w2).

 figure: Fig. 3

Fig. 3 – Open aperture transmittance variation as a function of average intensity, <I> = I0/2. For comparison it is also depicted the β = 0 case.

Download Full Size | PDF

In principle, ΔT should be calculated integrating over this profile using Nex given by Eq. (3). Fortunately, numerical results of this integral are in very good agreement with calculations using just the average radial intensity in Eq. (7). For the intensity range used in this paper, the difference between these two procedures is better than 8%. So, the full line in Fig. 3 was calculated using the average intensity. In the same Figure one can see the case β = 0 for comparison. The difference between them is due to ETU mechanism that decreases the excited state population, weakening the population lens, therefore reducing ΔT. The experimental points were fitted with Eq. (7) using Δσ obtained in the low intensity Z-scan measurement and leaving β and Is as fit parameters. Therefore, β = (6.4 ± 0.8) and Is = (18.1 ± 0.9) kW/cm2 were obtained. Reminding that Is = /ηpσgτo, with ηp = 1, τo = 240 μs results in σg ~(5.6 ± 0.4) × 10−20 cm2, in agreement with σg ~5.8 × 10−20 cm2 of Ref [23]. Moreover, it also agrees with Δσ = −5.1 × 10−20 (obtained from n2) since no significant ESA absorption is expected at 808 nm [22, 23]. Since αabs = Ntσg it was possible to obtain the Nd concentration as Nt = (1.28 ± 0.09) × 1020 ions/cm3 or 0.92 at. %. Finally, the ETU effect could be evaluated from the dimensionless parameter β = Ntγτo as γ = (2.0 ± 0.3) × 10−16 cm3/s. This value is in good agreement with γ = (2.8 ± 1) × 10−16 cm3/s and γ = (1.8 ± 0.2) × 10−16 cm3/s, found by Guyot et al [2] and Chen et al [3], while it is slightly higher than γ = 0.5 × 10−16 cm3/s found by Guy et al [1]. It should be noticed that these γ values were evaluated in samples with ~1 at. %-Nd, since γ increases linearly with Nd concentration [9, 10, 14]. However, the literature results cited above were obtained via the fluorescence decay behavior, which is sensitive to ETU process only in the initial decay curve. This can increase the uncertainty in γ determination since it requires high pump intensities and/or high doping. On the other hand, time resolved Z-scan transmittance is defined as the ratio between the transmitted intensity before any nonlinear effect and the intensity at the stationary state. Therefore, it is unnecessary to know the system dynamics (with its initial conditions, e. g., Nex(0)) after pumping. It can also be noted that the value obtained for γ presents a good accuracy when compared with the literature.

The Z-scan theory was obtained with the assumption that the refractive index, and consequently, the phase profile due to the refractive index variation, have a Gaussian profile. However, in the saturation regime, Nex(r) is no longer proportional to a Gaussian profile. It is well known that the standard Z-scan theory cannot be applied in the saturation regime, where the refractive index is proportional to S/(1 + S) [7]. In the present paper, both saturation and upconversion effects should contribute to a non-Gaussian refractive index profile. To approach this problem, we used the Fresnel-Kirchhoff diffraction integral considering the Gaussian beam propagation through a radial phase profile, Δϕ(r) proportional to Nex(r) given by

Δφ=2πλLeffn2'IS(1+S)2+4βS(1+S)2β
were n2 is the real nonlinear refractive index. Unfortunately this integral has no analytic solution, so that it was evaluated numerically only. We computed Z-scan curves for different S and β values.

Figure 4 shows the power dependence of ΔTpv (the difference between peak and valley transmittance) for β between 4 and 8. A good agreement between experimental and calculated values was obtained using β = 6, n2 = 4.4 x 10−9 cm2/W and Is = 18 kW/cm2, corroborating the values previously obtained.

 figure: Fig. 4

Fig. 4 – Peak-and-valley transmittance variation, of the closed Z-scan curve, as a function of the saturation parameter at the peak position, for λ = 808.7 nm.

Download Full Size | PDF

5. Conclusions

In this paper time resolved Z-scan measurements on a Nd:YAG crystal, on saturated and unsaturated regimes, were performed. The nonlinear refractive index, n2, was determined from the linear dependence of the Z-scan signals, from both closed an open aperture, with intensity, leading to Δαp = (5.5 ± 1) × 10−26 cm3and Δσ = (−5.1 ± 0.5) × 10−20 cm2. However, a nonlinear behavior was observed which could be attributed to an excited state population Nex I/Is - (1 + β)I2/Is2 + …, which is equivalent to consider Δn = n2I + n4I2 + ... where the real and imaginary parts of Δn are both in good agreement with the values β = 6 and Is = 18 kW/cm2. In this way, the Z-scan method was important to corroborate that both, the nonlinear refraction and absorption, are proportional to Nex(I). From the open aperture transmittance variation the ETU parameter, γ, could also be obtained. The saturation intensity was also obtained in the same measurement. It was shown that on the presence of ETU an effective saturation intensity (1 + β/2) times greater than the usual one appears. These parameters are particularly useful for laser designers in order to optimize the Nd:YAG gain transitions.

The results also show that, while the optical electronic nonlinearity and the energy transfer upconversion can be determined from several techniques, the Z-scan technique is confirmed as a good option because it combines sensitivity, simple analysis and simple experimental apparatus.

Acknowledgments

The authors would like to acknowledge the Brazilian agencies FAPESP, CNPq and FAPEMIG for the financial support.

References and links

1. S. Guy, C. L. Bonner, D. P. Shepherd, D. C. Hanna, A. C. Tropper, and B. Ferrand, “High-inversion densities in Nd : YAG: Upconversion and bleaching,” IEEE J. Quantum Electron. 34(5), 900–909 (1998). [CrossRef]  

2. Y. Guyot, H. Manaa, J. Y. Rivoire, R. Moncorgé, N. Garnier, E. Descroix, M. Bon, and P. Laporte, “Excited-State-Absorption and up-Conversion Studies of Nd3+-Doped Single Crystals Y3Al5O12, YLiF4, and LaMgAl11O19,” Phys. Rev. B Condens. Matter 51(2), 784–799 (1995). [CrossRef]   [PubMed]  

3. Y. F. Chen, C. C. Liao, Y. P. Lan, and S. C. Wang, “Determination of the Auger upconversion rate in fiber-coupled diode end-pumped Nd:YAG and Nd:YVO4 crystals,” Appl. Phys. B Lasers Opt. 70(4), 487–490 (2000). [CrossRef]  

4. V. Pilla, T. Catunda, H. P. Jenssen, and A. Cassanho, “Fluorescence quantum efficiency measurements in the presence of Auger upconversion by the thermal lens method,” Opt. Lett. 28(4), 239–241 (2003). [CrossRef]   [PubMed]  

5. M. Pollnau, P. J. Hardman, M. A. Kern, W. A. Clarkson, and D. C. Hanna, “Upconversion-induced heat generation and thermal lensing in Nd:YLF and Nd:YAG,” Phys. Rev. B 58(24), 16076–16092 (1998). [CrossRef]  

6. M. Sheik-Bahae, A. A. Said, and E. W. Van Stryland, “High-sensitivity, single-beam n(2) measurements,” Opt. Lett. 14(17), 955–957 (1989). [CrossRef]   [PubMed]  

7. L. C. Oliveira, T. Catunda, and S. C. Zilio, “Saturation effects in Z-scan measurements,” Jpn. J. Appl. Phys. 35(5), 2649–2652 (1996). [CrossRef]  

8. V. Pilla, P. R. Impinnisi, and T. Catunda, “Measurement of saturation intensities in ion doped solids by transient nonlinear refraction,” Appl. Phys. Lett. 70(7), 817 (1997). [CrossRef]  

9. C. Jacinto, D. N. Messias, A. A. Andrade, and T. Catunda, “Energy transfer upconversion determination by thermal-lens and Z-scan techniques in Nd3+-doped laser materials,” J. Opt. Soc. Am. B 26(5), 1002 (2009). [CrossRef]  

10. W. J. Lima, V. M. Martins, A. F. G. Monte, D. N. Messias, N. O. Dantas, M. J. V. Bell, and T. Catunda, “Energy transfer upconversion on neodymium doped phosphate glasses investigated by Z-scan technique,” Opt. Mater. (Amst) 35(9), 1724–1727 (2013). [CrossRef]  

11. R. C. Powell, Physics of Solid-State Lasers Materials (Springer-Verlag, 1998).

12. V. Lupei and A. Lupei, “Nd:YAG at its 50th anniversary: Still to learn,” J. Lumin. (to be published).

13. D. N. Messias, T. Catunda, J. D. Myers, and M. J. Myers, “Nonlinear electronic line shape determination in Yb3+-doped phosphate glass,” Opt. Lett. 32(6), 665–667 (2007). [CrossRef]   [PubMed]  

14. C. Jacinto, D. N. Messias, A. A. Andrade, S. M. Lima, M. L. Baesso, and T. Catunda, “Thermal lens and Z-scan measurements: Thermal and optical properties of laser glasses – A review,” J. Non-Cryst. Solids 352(32–35), 3582–3597 (2006). [CrossRef]  

15. S. M. Lima and T. Catunda, “Discrimination of resonant and nonresonant contributions to the nonlinear refraction spectroscopy of ion-doped solids,” Phys. Rev. Lett. 99(24), 243902 (2007). [CrossRef]   [PubMed]  

16. L. C. Oliveira and S. C. Zilio, “Single-beam time-resolved Z-scan measurements of slow absorbers,” Appl. Phys. Lett. 65(17), 2121–2123 (1994). [CrossRef]  

17. C. Jacinto, T. Catunda, D. Jaque, and J. G. Solé, “Fluorescence quantum efficiency and Auger upconversion losses of the stoichiometric laser crystal NdAl3(BO3)4,” Phys. Rev. B 72(23), 235111 (2005). [CrossRef]  

18. A. Ikesue, Y. L. Aung, and V. Lupei, Ceramic Lasers (Cambridge University Press, 2013).

19. A. A. Andrade, E. Tenorio, T. Catunda, M. L. Baesso, A. Cassanho, and H. P. Jenssen, “Discrimination between electronic and thermal contributions to the nonlinear refractive index of SrAlF5 : Cr+3,” J. Opt. Soc. Am. B 16(3), 395–400 (1999). [CrossRef]  

20. C. Jacinto, A. A. Andrade, T. Catunda, S. M. Lima, and M. L. Baesso, “Thermal lens spectroscopy of Nd:YAG,” Appl. Phys. Lett. 86(3), 034104 (2005). [CrossRef]  

21. S. M. Lima, T. Andrade, R. Catunda, F. Lebullenger, Y. Smektala, Jestin, and M. L. Baesso, “Thermal and optical properties of chalcohalide glass,” J. Non-Cryst. Solids 284(1–3), 203–209 (2001). [CrossRef]  

22. S. Kuck, L. Fornasiero, E. Mix, and G. Huber, “Excited state absorption and stimulated emission of Nd3+ in rystals. Part I: Y3Al5O12, YAlO3, and Y2O3,” Appl. Phys. B 67(2), 151–156 (1998). [CrossRef]  

23. J. Margerie, R. Moncorgé, and P. Nagtegaele, “Spectroscopic investigation of variations in the refractive index of a Nd:YAG laser crystal: Experiments and crystal-field calculations,” Phys. Rev. B 74(23), 235108 (2006). [CrossRef]  

24. C. Jacinto, T. Catunda, D. Jaque, L. E. Bausá, and J. García-Solé, “Thermal lens and heat generation of Nd:YAG lasers operating at 1.064 and 1.34 microm,” Opt. Express 16(9), 6317–6323 (2008). [CrossRef]   [PubMed]  

25. R. C. Powell, S. A. Payne, L. L. Chase, and G. D. Wilke, “Index-of-refraction change in optically pumped solid-state laser materials,” Opt. Lett. 14(21), 1204–1206 (1989). [CrossRef]   [PubMed]  

26. O. L. Antipov, A. S. Kuzhelev, A. Y. Luk’yanov, and A. P. Zinov’ev, “Changes in the refractive index of an Nd:YAG laser crystal on excitation of the Nd3+ ions,” Quantum Electron. 28(10), 867–874 (1998). [CrossRef]  

27. R. Soulard, A. Zinoviev, J. L. Doualan, E. Ivakin, O. Antipov, and R. Moncorge, “Detailed characterization of pump-induced refractive index changes observed in Nd:YVO4, Nd:GdVO4 and Nd:KGW,” Opt. Express 18(2), 1553–1568 (2010). [CrossRef]   [PubMed]  

28. T. Godin, R. Moncorgé, J.-L. Doualan, M. Fromager, K. Ait-Ameur, R. A. Cruz, and T. Catunda, “Optically pump-induced athermal and nonresonant refractive index changes in the reference Cr-doped laser materials: Cr:GSGG and ruby,” J. Opt. Soc. Am. B 29(5), 1055 (2012). [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 (4)

Fig. 1
Fig. 1 – Time resolved Z-scan measurement for (a) closed aperture, (b) open aperture and (c) the ratio closed/open at λ = 808.7 nm, chopper frequency f = 820 Hz, P = 120 mW and ω0 = 35μm.
Fig. 2
Fig. 2 – Transient transmittance obtained from both closed, 50% transmittance, and opened, 100% transmittance, Z-scan apertures. The solid lines are exponential fits.
Fig. 3
Fig. 3 – Open aperture transmittance variation as a function of average intensity, <I> = I0/2. For comparison it is also depicted the β = 0 case.
Fig. 4
Fig. 4 – Peak-and-valley transmittance variation, of the closed Z-scan curve, as a function of the saturation parameter at the peak position, for λ = 808.7 nm.

Equations (8)

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

n 2 = N t I s ( 2π n 0 f L 2 Δ α p i λ 4π Δσ )
d N ex dt = S τ 0 N t (1+S) τ 0 N ex γ N ex 2
N ex = N t (1+S) 2 +4βS (1+S) 2β
θ= φ P abs λK ds dT
Δ Φ o = 2π λ L eff   n 2 , 2P π w 2
θ Δ Φ o = φ 4K ds dT w 2 ( n 2 / α abs )
ΔT=Δσ L eff N ex .
Δφ= 2π λ L eff n 2 ' I S (1+S) 2 +4βS (1+S) 2β
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.