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

Cladding YAG crystal fibers with high-index glasses for reducing the number of guided modes

Open Access Open Access

Abstract

Yttrium aluminium garnet (YAG) crystal fibers with a core diameter of 40 μm were cladded by high index glasses using the co-drawing laser-heated pedestal growth method. Due to the extremely large cooling rates in the fabrication processes, unexpected and phenomenally large index drops of 0.018 and at least 0.02 were found from the as-grown capillary and the YAG crystal fiber cladding compared with bulk N-SF57’s, respectively. The high-index glass cladding is effective in reducing the number of guided modes, and the intensity profiles of the crystal fiber show there are only four guided modes at 532 nm.

©2013 Optical Society of America

1. Introduction

Yttrium aluminum garnet (YAG) has been widely used as a laser host because of its superior optical, thermal, mechanical properties, as well as its plurality in hosting active ions such as Ce3+, Yb3+, Nd3+, Cr4+, Er3+, Tm3+, and Ho3+ ions with fluorescence spectra ranging from 0.5 to 3 μm. The gain medium in a fiber form has the advantages over bulk gain medium including the better thermal management of the fiber structure having a larger surface-to-volume ratio and also large pump absorption. The laser-heated pedestal growth (LHPG) has been shown an effective process for single-crystal fiber growth [1]. Adding a cladding layer to the crystal fiber is helpful in improving the propagation loss and output beam quality. Growth methods of YAG-based crystal fibers with a cladding structure reported include the surface regrowth for dopant out-diffusion method [2], the extrusion method using high-index glasses as the cladding materials [3], and the co-drawing LHPG (CD-LHPG) method capable of growing both single- and double-cladding crystal fiber structures [4,5]. The recent advancement on the CD-LHPG technique has results in various applications in broadband light source, laser, and optical amplifier [69]. The quality of the crystal fiber fabricated by the LHPG method is confirmed by the record-low pump threshold and record-high slope efficiency of the Cr4+:YAG crystal fiber laser at room temperature [8]. Compared with silica fiber lasers, the crystalline core offers high emission cross section for transition metal ions because of the unique local matrix. Also the thermal conductivity and stimulated Brillouin scattering threshold of the YAG crystal are higher than those of the fused silica. However, so far, the cladded crystal fibers are highly multimode.

The Yb:YAG single-crystal un-clad fibers fabricated by the micro-pulling down (μPD) method have been used as gain medium or as the last stage power amplifier for the 1030-nm high power laser applications recently [10,11]. The pump beams were guided by the crystal fibers of 1-mm diameter but the intra-cavity laser or the signal beam was free propagating inside the fiber in both cases. Nd:GdVO4 single crystal fiber with a core diameter of 3 mm was cladded by pure GdVO4 using the double-die edge-defined film-fed growth (EFG) method [12]. High power Er:YAG lasers at eye-safe 1617- and 1645-nm for remote sensing and ranging applications have attracted intensive research activities in recent years. The laser efficiency degrades when the doping concentration of erbium ions exceeds 1 at.%, due to the energy transfer up-conversion [13]. The Er:YAG crystal fiber provides a solution for achieving sufficient pump absorption, in the case of using low erbium ion concentration Er:YAG crystal. The 0.5% Er:YAG crystal fiber grown by μPD method has a large diameter of 800 μm but no cladding [14]. The Er:YAG double-clad crystalline channel waveguide was fabricated with a core of 500 μm × 500 μm using the adhesive-free bonding method and used in cladding-pump laser at 1645 nm with an improved beam quality [15]. The LHPG method has the advantage of growing crystal fibers with smaller core diameters, compared with other fiber or waveguide fabrication methods.

Here we report the fabrication of YAG crystal fibers with core diameters of 40 μm and cladded by SF57HHT and N-SF57 high-index glasses using the CD-LHPG method for reducing the number of guided modes. A phenomenally large index drop of the glass cladding of the crystal fiber was found due to the extremely large cooling rate in the fabrication process. The measured output intensity profiles at several wavelengths show that the glass-clad crystal fiber is guiding at 532, 633 and 780 nm wavelengths where it should be un-guiding because the index of bulk N-SF57 optical glass is larger than the YAG’s. The output intensity profile at 532 nm shows the glass cladding is effective in reducing the number of guided modes and consists of mainly four modes. The finding of the index drop during CD-LHPG crystal fiber fabrication process is critical in selecting proper glass as the cladding material when further pursuing single-mode crystal fibers operating at the 1-μm Yb:YAG, 1.3- to 1.5-μm Cr4+:YAG, and 1.6-μm Er:YAG laser wavelengths.

2. Theoretical grounds

The normalized frequency, V, and numerical aperture, NA, of a glass-clad YAG crystal fiber are defined as:

V=2πλaNA=2πλanYAG2nglass2
where a and λ are the fiber core radius and the optical wavelength, and nYAGand nglass are the indices of YAG core and glass cladding, respectively. The requirements of a single-mode fiber include not only a small core radius but also a small NA so the normalized frequency V is smaller than 2.405 [16].

The available glass capillaries for the CD-LHPG process were made of fused silica or borosilicate glass [49]. The index of YAG crystal is quite high so the index difference between the YAG crystal core and the glass cladding is large. As a result, the glass-clad crystal fibers are all multi-mode. The N-SF57 (Schott) glass having a comparable high index is chosed as the cladding material. The index dispersion curves of YAG [17] and N-SF57 [18] are calculated using the Sellmeier equation and shown in Fig. 1. The index of YAG is higher than the index of bulk N-SF57 glass for wavelengths longer than 0.872 μm so the N-SF57-clad YAG crystal fiber should be guiding. The NA of N-SF57-clad YAG crystal fiber was expected to be from 0 to 0.17 for wavelength from 0.872 to 1.7 μm, as shown in Fig. 1. However, a large index drop of the as-drawn N-SF57 capillary was observed which is attributed to the extremely large cooling rate in the drawing process. The fit index curve of N-SF57 capillary to the index measurement data plotted as dots is also shown in Fig. 1. As will be depicted later, an even larger index drop of the N-SF57 cladding of the YAG crystal fiber was found after the CD-LHPG process. The N-SF57-clad YAG crystal fiber is guiding at wavelengths shorter than the original cross-over wavelength at 0.872 μm. SF57HHT glass (Schott) has a refractive index similar to N-SF57’s. The refractive index difference between N-SF57 and SF57HHT is less than 10−4 in the wavelength range from 472 to 812 nm.

 figure: Fig. 1

Fig. 1 Dispersion of YAG, N-SF57 bulk and capillary, and NA of the N-SF57-clad YAG crystal fiber. Dots: Index measurement data of the N-SF57 capillary.

Download Full Size | PDF

The index of an optical glass changes in a heat treatment process. The change of refractive index Δn is related to the ratio of annealing (cooling) rates in the processes in the logarithmic scale [19]:

Δn=mnlog(hxh0)
where h0, hxare the original and new annealing rates, and mn is the annealing coefficient for the refractive index. A fast annealing rate leads to a large index reduction of an optical glass. For N-SF57 glass, mn is −0.00263 at 589.3 nm. The dependence of the refractive index change Δn to the annealing rate of N-SF57 glass is shown in Fig. 2. The index change is compared with an original annealing rate of 7 °C/hr. The thermal treatment induced index change is pretty small, usually in the order of 10−4 to 10−5 [19].

 figure: Fig. 2

Fig. 2 Refractive index change Δn as a function of the annealing rate of N-SF57 glass. The original annealing rate is 7 °C/hr.

Download Full Size | PDF

3. Experimental results and discussion

3.1 Growth process of glass-clad YAG crystal fibers

Pure YAG crystal rods with a 0.5 mm × 0.5 mm rectangular cross-section were used as the source materials for the growth of crystal fibers. Single-crystalline 40-μm YAG fibers were grown at a growth speed of 3 mm/min through a three-step size-reduction process using the LHPG method. Two glasses including SF57HHT and N-SF57 to be used as the cladding materials were made into capillaries for the first time, to facilitate the CD-LHPG process. A 40-μm YAG single-crystal fiber was inserted into the glass capillary. Then the crystal-fiber-filled glass capillary was sealed at one end using the CO2 laser of the LHPG system, and vacuumed through the other end to remove the unwanted air in the gap between the inner wall of the glass capillary and the single crystal fiber. The glass-clad YAG crystal fiber was grown using the CD-LHPG process at a growth speed of 3 mm/min. The thermal parameters of the materials are shown in Table 1. The softening temperatures of the SF57HHT and N-SF57 glasses are 716 and 519 °C [18], and both are much lower than the 1970 °C melting temperature of YAG crystal [20]. The CO2 laser power during the co-drawing process was controlled so only the glass capillary got softened and attached to the YAG single-crystal fiber, while the crystal fiber remained unmelt, as shown in Fig. 3. During the CD-LHPG process, a very large cooling rate is expected. The reasons are two-fold. The heating zone is small since it is located approximately at the beam waist of the focused CO2 laser. Also the volume of the heated and being grown section of crystal fiber is small since the crystal fiber is so thin and consequently the heat capacitance of the heated section is small. Therefore the temperature gradient is large outside the heating zone and a large index drop of the cladding glass was induced, as described in Eq. (2). Then the index cross-over wavelength of the crystal fiber is likely to move from the original 872 nm toward a shorter wavelength.

Tables Icon

Table 1. Thermal Parameters of Materials

 figure: Fig. 3

Fig. 3 Microscope photos of (a) the glass-clad crystal fiber under growth using CD-LHPG process, (b) cross-section of N-SF57 glass capillary with inner and outer diameters of 130 and 330 μm, (c) end face of N-SF57-clad YAG crystal fiber. The diameters of core and cladding are 40 and 353 μm, respectively. The core eccentricity is 23.7 μm.

Download Full Size | PDF

3.2 SF57HHT-clad YAG crystal fiber

Due to the material availability, SF57HHT capillary was tried as a preliminary test for cladding the YAG crystal fiber. The SF57HHT glass was made into capillaries with fine machining and torch drawing. The ~1-mm outer diameters are not quite uniform, and the inner hole diameters are also large, i.e. 300 to 500 µm. The length of the SF57HHT-clad YAG crystal fiber is 5.4 cm. The measured propagation loss is 0.1 dB/cm using the cut-back method. The output intensity profiles of the crystal fibers were measured for verification of the numbers of guided modes. As shown in Fig. 4(a), the SF57HHT-clad sample was guiding even at 532 nm wavelength. The index drop of the SF57HHT cladding was at least 0.02. As shown in Figs. 4(b) and 4(c), the patterns of the as-grown sample at 633 and 1064 nm were non-circular, possibly due to the stress induced index non-uniformity of the non-concentric inner and outer diameters. An annealing process with a 24-hour soaking time at the transformation temperature Tg of SF57HHT, and a subsequently controlled slow cooling rate of 7 °C/hr from Tg to 120 °C below it was used for stress releasing and index fine-tuning. The result after annealing showed the SF57HHT-clad YAG crystal fiber sample becomes un un-guiding at 532 and 633 nm. The index drop of the SF57HHT cladding is reduced to no larger than 0.01. As shown in Fig. 4(d), the output intensity profile at 1064 nm of the annealed fiber sample becomes circular and has less speckles, which indicates that the stress was effectively released and the index of glass cladding was increased.

 figure: Fig. 4

Fig. 4 Intensity profiles of SF57HHT-clad YAG crystal fiber at (a) 532 nm, (b) 633 nm, (c) 1064 nm, as-grown, and (d) 1064 nm, annealed.

Download Full Size | PDF

3.3 N-SF57-clad YAG crystal fiber

To improve the cladding uniformity, N-SF57 capillaries were prepared from a drawing tower with inner and outer diameters of 130 and 330 μm, respectively. The length of the N-SF57-clad YAG crystal fiber was 4.8 cm. The endface photo of N-SF57 capillary is shown in Fig. 3(b). The refractive index of the as-grown N-SF57 capillary was measured using the V-block refractometer method at four wavelengths including 441, 639, 947, and 1550 nm [21]. The index dispersion curve of the glass capillary was fit using the Sellmeier equation. Figure 5 shows the wavelength-dependent index drop of the as-grown N-SF57 capillary compared with its bulk index. At 639 and 947 nm, the measured index drops were as large as 0.017, which are at least two or three orders of magnitude larger than the typical data [19]. The index drop decreased only slightly towards longer wavelength when the wavlength was larger than 1 μm, but increased significantly toward shorter wavelength when the wavelength was shorter than 0.6 μm. Large refractive index drops of 0.005 (core) and 0.01 (cladding) were ever reported after a glass fiber drawing process using two different optical glasses as the core and the cladding materials, but were assumed to be constant over the spectrum [22]. The phenomenally large and wavelength dependent index drop of the thermally treated glass is reported for the first time, to our knowledge. The simplification of wavelength-independent index drop is more appropriate for the long (infrared) wavelength range, but is less accurate in the short (visible) wavelength range.

 figure: Fig. 5

Fig. 5 Index drop curves of the N-SF57 capillary and fiber cladding. Dots: measurement data. Curve “capillary”: by fitting the measurement data. Due to the larger measurement uncertainty at 441 nm, the curve is shown as “dashed” for the wavelengths below 639 nm. The fit index drop is 0.018 at 532 nm. Curve “cladding”: estimation of the fiber cladding by scaling curve “capillary” to 0.02 at 532 nm.

Download Full Size | PDF

The endface photo of N-SF57-clad YAG crystal fiber grown by the CD-LHPG process is shown in Fig. 3(c). The core of the crystal fiber is not so clear because the indices of the core and cladding are very close to each other. The 23.7-μm eccentricity of the core is due to the large spacing between the 40-μm diameter of YAG crystal fiber and the 130-μm inner diameter of the N-SF57 capillary. Measurement of the surface reflection of a polished crytal fiber endface was attempted for deriving the index profile. However, the surface index of the glass cladding changes as much as 0.03~0.05 after the polishing process [23]. The surface reflection measurement is unable to discern the minute index difference between the crystal core and glass cladding. The output intensity profile of the crystal provides a more convincing information of the guiding condition. Figures 6(a) to 6(c) are the measured intensity profiles at 780, 633, and 532 nm, respectively. It is clear that the number of guides modes decreases at shorter wavelengths, as the difference between the indices of YAG crystal core and N-SF57 glass cladding becomes smaller. Like the SF57HHT-clad sample, the N-SF57-clad sample is also guiding at the short wavelength at 532 nm. And the intensity profile shows the number of guided modes is limited and indeed significantly reduced. Simulation was used to fit the measured profile at 532 nm. The simulated intensity distribution as shown in Fig. 6(d) is a linear combination of the intensities of the lowest four fiber modes, i.e., LP01, LP11, LP21, and LP02 modes. Existence of other higher order modes are possible but with relatively small intensities [24]. The calculated beam quality parameter, M2, based on the measured profile of Fig. 6(c) are about 2.9 [25]. If considering only four guided modes, the normalized frequncy should be less than 5.14. The corresponding NA and index difference between fiber core and cladding are smaller than 0.022 and 1.3 × 10−4, respectively. The index drop of the N-SF57 glass cladding of the crystal fiber sample with respect to its original bulk value is at least 0.02, similar to the value of SF57HHT-clad sample, but larger than the 0.018 index drop of the as-grown N-SF57 capillary as shown in Fig. 5. This implies the cooling rate in the crystal fiber cladding growth process is even larger than that of the glass capillary pulling process. The annealing process was applied to the N-SF57-clad YAG crystal fiber for tuning up the index of the glass cladding, and consequently moving the single-mode or few-mode wavelength toward the longer wavelength. However, the N-SF57 glass cladding is tentative to crack after the annealing process.

 figure: Fig. 6

Fig. 6 Intensity profiles of the N-SF57-clad YAG crystal fiber measured at (a) 780 nm, (b) 633 nm, and (c) 532 nm. (d) Simulated profile using the combination of the intensities of the lowest four modes.

Download Full Size | PDF

4. Conclusion

In conclusion, we have successfully fabricated a YAG crystal fiber with an N-SF57 glass cladding. The glass-clad crystal fiber was guiding at wavelengths where it was supposed to be un-guiding. A phenomenally large index drop of the N-SF57 glass cladding after the CD-LHPG growth process was unexpectedly found so the guiding region was extended to the visible wavelength range. The effectiveness of reducing the number of guided modes is confirmed by the measured intensity profiles. At 532 nm, the output profile shows that there are four dominant guided modes. To tune the few-mode region toward the longer wavelength, the large index drop of the glass cladding should be taken into account when choosing the suitable glass as the fiber cladding material. The high-index-glass-clad crystal fiber has the potential for designing the large-core gain fibers for high power laser applications with a good beam quality.

Acknowledgment

This work is partially supported by the National Science Council, Taiwan.

References and links

1. M. M. Fejer, J. L. Nightingale, G. A. Magel, and R. L. Byer, “Laser heated miniature pedestal growth apparatus for single crystal optical fibers,” Rev. Sci. Instrum. 55(11), 1791–1796 (1984). [CrossRef]  

2. C. A. Burrus and L. A. Coldren, “Growth of single-crystal sapphire-clad ruby fibers,” Appl. Phys. Lett. 31(6), 383–385 (1977). [CrossRef]  

3. M. J. F. Digonnet, C. J. Gaeta, D. O'Meara, and H. J. Shaw, “Clad Nd:YAG fibers for laser applications,” J. Lightwave Technol. 5(5), 642–646 (1987). [CrossRef]  

4. C. Y. Lo, K. Y. Huang, J. C. Chen, S. Y. Tu, and S. L. Huang, “Glass-clad Cr4+:YAG crystal fiber for the generation of superwideband amplified spontaneous emission,” Opt. Lett. 29(5), 439–441 (2004). [CrossRef]   [PubMed]  

5. K. Y. Huang, K. Y. Hsu, D. Y. Jheng, W. J. Zhuo, P. Y. Chen, P. S. Yeh, and S. L. Huang, “Low-loss propagation in Cr4+:YAG double-clad crystal fiber fabricated by sapphire tube assisted CDLHPG technique,” Opt. Express 16(16), 12264–12271 (2008). [CrossRef]   [PubMed]  

6. C. C. Tsai, T. H. Chen, Y. S. Lin, Y. T. Wang, W. Chang, K. Y. Hsu, Y. H. Chang, P. K. Hsu, D. Y. Jheng, K. Y. Huang, E. Sun, and S. L. Huang, “Ce3+:YAG double-clad crystal-fiber-based optical coherence tomography on fish cornea,” Opt. Lett. 35(6), 811–813 (2010). [CrossRef]   [PubMed]  

7. Y. S. Lin, T. C. Cheng, C. C. Tsai, K. Y. Hsu, D. Y. Jheng, C. Y. Lo, P. S. Yeh, and S. L. Huang, “High-luminance white-light point source using Ce,Sm:YAG double-clad crystal fiber,” IEEE Photon. Technol. Lett. 22(20), 1494–1496 (2010). [CrossRef]  

8. C. C. Lai, C. P. Ke, S. K. Liu, D. Y. Jheng, D. J. Wang, M. Y. Chen, Y. S. Li, P. S. Yeh, and S. L. Huang, “Efficient and low-threshold Cr4+:YAG double-clad crystal fiber laser,” Opt. Lett. 36(6), 784–786 (2011). [CrossRef]   [PubMed]  

9. K. Y. Hsu, D. Y. Jheng, Y. H. Liao, T. S. Ho, C. C. Lai, and S. L. Huang, “Diode-laser-pumped glass-clad Ti:sapphire crystal fiber based broadband light source,” IEEE Photon. Technol. Lett. 24, 854–856 (2012).

10. X. Délen, S. Piehler, J. Didierjean, N. Aubry, A. Voss, M. A. Ahmed, T. Graf, F. Balembois, and P. Georges, “250 W single-crystal fiber Yb:YAG laser,” Opt. Lett. 37(14), 2898–2900 (2012). [CrossRef]   [PubMed]  

11. X. Délen, L. Deyra, A. Benoit, M. Hanna, F. Balembois, B. Cocquelin, D. Sangla, F. Salin, J. Didierjean, and P. Georges, “Hybrid master oscillator power amplifier high-power narrow-linewidth nanosecond laser source at 257 nm,” Opt. Lett. 38(6), 995–997 (2013). [CrossRef]   [PubMed]  

12. M. Matsukura, O. Nakamura, S. Watanabe, A. Miyamoto, Y. Furukawa, Y. Sato, T. Taira, T. Suzudo, and H. Mifune, “Laser properties of composite Nd:GdVO4 single crystal grown by the double die EFG method,” in Conference on Lasers and Electro-Optics (CLEO 2007) Technical Digest, Baltimore (US) (Optical Society of America, May, 2007), paper CFA2.

13. J. W. Kim, D. Y. Shen, J. K. Sahu, and W. A. Clarkson, “Fiber-laser-pumped Er:YAG lasers,” IEEE J. Sel. Top. Quantum Electron. 15(2), 361–371 (2009). [CrossRef]  

14. I. Martial, S. Bigotta, M. Eichhorn, C. Kieleck, J. Didierjean, N. Aubry, R. Peretti, F. Balembois, and P. Georges, “Er:YAG fiber-shaped laser crystals (single crystal fibers) grown by micro-pulling down: Characterization and laser operation,” Opt. Mater. 32(9), 1251–1255 (2010). [CrossRef]  

15. N. Ter-Gabrielyan, V. Fromzel, X. Mu, H. Meissner, and M. Dubinskii, “High efficiency, resonantly diode pumped, double-clad, Er:YAG-core, waveguide laser,” Opt. Express 20(23), 25554–25561 (2012). [CrossRef]   [PubMed]  

16. A. Yariv and P. Yeh, Photonics: Optical Electronics in Modern Communications (Oxford University Press, 2007), Chap. 3.

17. D. E. Zelmon, D. L. Small, and R. Page, “Refractive-index measurements of undoped yttrium aluminum garnet from 0.4 to 5.0 μm,” Appl. Opt. 37(21), 4933–4935 (1998). [CrossRef]   [PubMed]  

18. Schott AG data sheet.

19. H. Bach and N. Nuroth, eds., The Properties of Optical Glass (Springer Verlag, 1995), Chap. 2.

20. W. Koechner and M. Bass, Solid-State Lasers (Springer, 2003), Chap. 2.

21. Schott Technical information TIE-29, Refractive Index and Dispersion (2007).

22. J. Nishimura and K. Morishita, “Control of spectral characteristics of dispersive optical fibers by annealing,” J. Lightwave Technol. 15(2), 294–298 (1997). [CrossRef]  

23. J. Stone and H. E. Earl, “Surface effects and reflection refractometry of optical fibres,” Opt. Quantum Electron. 8(5), 459–463 (1976). [CrossRef]  

24. J. W. Nicholson, A. D. Yablon, J. M. Fini, and M. D. Mermelstein, “Measuring the modal content of large-mode-area fibers,” IEEE J. Sel. Top. Quantum Electron. 15(1), 61–70 (2009). [CrossRef]  

25. H. Yoda, P. Polynkin, and M. Mansuripur, “Beam quality factor of higher order modes in a step-index fiber,” J. Lightwave Technol. 24(3), 1350–1355 (2006). [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 (6)

Fig. 1
Fig. 1 Dispersion of YAG, N-SF57 bulk and capillary, and NA of the N-SF57-clad YAG crystal fiber. Dots: Index measurement data of the N-SF57 capillary.
Fig. 2
Fig. 2 Refractive index change Δn as a function of the annealing rate of N-SF57 glass. The original annealing rate is 7 °C/hr.
Fig. 3
Fig. 3 Microscope photos of (a) the glass-clad crystal fiber under growth using CD-LHPG process, (b) cross-section of N-SF57 glass capillary with inner and outer diameters of 130 and 330 μm, (c) end face of N-SF57-clad YAG crystal fiber. The diameters of core and cladding are 40 and 353 μm, respectively. The core eccentricity is 23.7 μm.
Fig. 4
Fig. 4 Intensity profiles of SF57HHT-clad YAG crystal fiber at (a) 532 nm, (b) 633 nm, (c) 1064 nm, as-grown, and (d) 1064 nm, annealed.
Fig. 5
Fig. 5 Index drop curves of the N-SF57 capillary and fiber cladding. Dots: measurement data. Curve “capillary”: by fitting the measurement data. Due to the larger measurement uncertainty at 441 nm, the curve is shown as “dashed” for the wavelengths below 639 nm. The fit index drop is 0.018 at 532 nm. Curve “cladding”: estimation of the fiber cladding by scaling curve “capillary” to 0.02 at 532 nm.
Fig. 6
Fig. 6 Intensity profiles of the N-SF57-clad YAG crystal fiber measured at (a) 780 nm, (b) 633 nm, and (c) 532 nm. (d) Simulated profile using the combination of the intensities of the lowest four modes.

Tables (1)

Tables Icon

Table 1 Thermal Parameters of Materials

Equations (2)

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

V= 2π λ aNA= 2π λ a n YAG 2 n glass 2
Δn= m n log( h x h 0 )
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.