• Ei tuloksia

Complete spatial coherence characterization of quasi-random laser emission from dye doped transparent wood

N/A
N/A
Info
Lataa
Protected

Academic year: 2022

Jaa "Complete spatial coherence characterization of quasi-random laser emission from dye doped transparent wood"

Copied!
10
0
0

Kokoteksti

(1)

2018

Complete spatial coherence

characterization of quasi-random laser

emission from dye doped transparent wood

Koivurova, Matias

The Optical Society

Tieteelliset aikakauslehtiartikkelit

© Optical Society of America

© 2018 Optical Society of America. Users may use, reuse, and build upon the article, or use the article for text or data mining, so long as such uses are for non-commercial purposes and appropriate attribution is maintained. All other rights are reserved

http://dx.doi.org/10.1364/OE.26.013474

https://erepo.uef.fi/handle/123456789/6674

Downloaded from University of Eastern Finland's eRepository

(2)

Complete spatial coherence characterization of quasi-random laser emission from dye doped transparent wood

M

ATIAS

K

OIVUROVA

,

1,4,*

E

LENA

V

ASILEVA

,

2,4

Y

UANYUAN

L

I

,

3

L

ARS

B

ERGLUND

,

3 AND

S

ERGEI

P

OPOV2

1Institute of Photonics, University of Eastern Finland, P. O. Box 111, FI-80101 Joensuu, Finland

2Department of Applied Physics, SCI school, KTH-Royal Institute of Technology, Isafjordsgatan 22, SE-16440 Stockholm, Sweden

3Wallenberg Wood Science Center, Department of Fiber and Polymer Technology, KTH, Teknikringen 42, SE-10044 Stockholm, Sweden

4These authors contributed equally

*matias.koivurova@uef.fi

Abstract: We report on the experimental determination of the complete two coordinate spatial coherence function of light emitted by a quasi-random laser, implemented on recently introduced dye-doped transparent wood. The spatial coherence was measured by means of a double grating interferometer, which has some advantages over the standard Young’s interferometer. Analysis of the spatial coherence reveals that emission from such a material can be considered as a superposition of several spatial modes produced by individual emitters within semi-ordered scattering medium. The overall degree of coherence,γ, for this quasi-random laser was found to be 0.16±0.01, having possible applications in speckle free laser imaging and illumination.

© 2018 Optical Society of America under the terms of theOSA Open Access Publishing Agreement OCIS codes:(030.1640) Coherence; (140.3538) Lasers, pulsed; (140.2050) Dye lasers.

References and links

1. D. S. Wiersma, “The physics and applications of random lasers,” Nat. Phys.4, 359–367 (2008).

2. D. S. Wiersma, “Disordered photonics,” Nat. Photonics7, 188–196 (2013).

3. R. C. Polson, M. E. Raikh, and Z. V. Vardeny, “Universal properties of random lasers,” IEEE J. Sel. Top. Quantum Electron.9, 120–123 (2003).

4. H. Cao, Y. Ling, J. Y. Xu, C. Q. Cao, and P. Kumar, “Photon statistics of random lasers with resonant feedback,”

Phys. Rev. Lett.86, 4524–4527 (2001).

5. B. Redding, M. A. Choma, and H. Cao, “Speckle-free laser imaging using random laser illumination,” Nat. Photonics 6, 355–359 (2012).

6. H. Cao,Y. G. Zhao, S. T. Ho, E. W. Seelig, Q. H. Wang, and R. P. H. Chang, “Random laser action in semiconductor powder,” Phys. Rev. Lett.82, 2278 (1999).

7. B. Redding, M. A. Choma, and H. Cao, “Spatial coherence of random laser emission,” Opt. Lett.36, 3404–3406, (2011).

8. C. Gouedard, D. Husson, C. Sauteret, F. Auzel, and A. Migus, “Generation of spatially incoherent short pulses in laser-pumped neodymium stoichiometric crystals and powders,” J. Opt. Soc. Am. B10, 2358–2363 (1993).

9. M. Koivurova, H. Partanen, J. Turunen, and A. T. Friberg, “Grating interferometer for light-efficient spatial coherence measurement of arbitrary sources,” Appl. Opt.56,5216–5227 (2017).

10. E. Vasileva, Y. Li, I. Sychugov, M. Mensied, L. Berglund, and S. Popov, “Lasing from organic dye molecules embedded in transparent wood,” Adv. Opt. Mater.5, 1700057 (2017).

11. Y. Li, Q. Fu, S. Yu, M. Yan, and L. Berglund, “Optically transparent wood from a nanoporous cellulosic template:

combining functional and structural performance,” Biomacromolecules17, 1358–1364 (2016).

12. I. Burgert, E. Cabane, C. Zollfrank, and L. Berglund, “Bio-inspired functional wood-based materials – hybrids and replicates,” Int. Mater. Rev.60, 431–450 (2015).

13. M. Zhu, J. Song, T. Li, A. Gong, Y.Wang, J. Dai, Y. Yao, W. Luo, D. Henderson, and L. Hu, “Highly anisotropic, highly transparent wood composites,” Adv. Mater.28, 7563 (2016).

14. T. Li, M. Zhu, Z. Yang, J. Song, J. Dai, Y. Yao, W. Luo, G. Pastel, B. Yang, and L. Hu, “Wood composite as an energy efficient building material: guided sunlight transmittance and effective thermal insulation,” Adv. Energy Mater.6, 1601122 (2016).

#323464 https://doi.org/10.1364/OE.26.013474

Journal © 2018 Received 20 Feb 2018; revised 19 Apr 2018; accepted 19 Apr 2018; published 10 May 2018

(3)

15. M. Borrega, P. Ahvenainen, R. Serimaa, and L. Gibson, “Composition and structure of balsa (Ochroma pyramidale) wood,” Wood Sci. Technol.49, 403–420 (2015).

16. S. Fink, “Transparent wood – a new approach in the functional study of wood structure,” Holzforschung5, 403–408 (1992).

17. Li, Y., Q. Fu, S. Yu, M. Yan, and L. Berglund, “optically transparent wood from a nanoporous cellulosic template:

combining functional and structural performance,” Biomacromolecules4, 1358–1364 (2016).

18. M. Zhu, T. Li, C. Davis, Y. Yao, J. Dai, Y. Wang, F. AlQatari, J. W. Gilman and L. Hu, “Transparent and haze wood composites for highly efficient broadband light management in solar cells,” Nano Energy26, 332–340 (2016).

19. A. Maslyukov, S. Sokolov, M. Kaivola, K. Nyholm and S. Popov, “Solid-state dye laser with modified poly (methyl methacrylate)-doped active elements,” Appl. Opt.34, 1516–1518 (1995).

20. H. Lajunen, J. Tervo, and P. Vahimaa, “Overall coherence and coherent-mode expansion of spectrally partially coherent plane-wave pulses,” J. Opt. Soc. Am. A21, 2117-2123 (2004).

21. H. Partanen, J. Turunen, and J. Tervo, “Coherence measurement with digital micromirror device,” Opt. Lett.39, 1034–1037 (2014).

22. L. Mandel and E. Wolf,Optical Coherence and Quantum Optics(Cambrigde University, 1995).

23. O. C. Peterson and B. B. Snavely, “Stimulated emission from flashlamp-excited organic dyes in polymethil methacrylate,” Appl. Phys.Lett.12, 238–240 (1968).

24. F. Duarte, L. Hillman, W. Lloyd, P. Liao, and P. Kelley,Dye laser principles: with application, (Academic Press, 1990).

25. F. Duarte,Tunable Laser Applications, (CRC Press, 2008).

26. F. P. Schäfer,Dye Lasers, (Springer-Verlag, 1990).

27. H. Partanen, J. Tervo, and J. Turunen, “Spatial coherence of broad-area laser diodes,” Appl. Opt.52,3221–3228 (2013).

28. N. Stelmakh and M. Flowers, “Measurement of spatial modes of broad-area diode Lasers with 1-GHz resolution grating spectrometer,” IEEE Photon. Technol. Lett.18, 1618–1620 (2006).

29. T. Okamoto and R. Yoshitome, “Random lasing in dye-doped polymer random media with a bubble structure,” J. Opt.

Soc. Am. B34, 1497–1502, (2017).

1. Introduction

There has been considerable interest towards experimental studies on the correlation properties of light produced by exotic sources, and possible methods to control their coherence. Coherence is one of the fundamental properties of light, and it was thought for a long time that laser light always exhibits high spatial and temporal coherence compared to thermal light. Novel sources, such as random lasers [1–4], have shown that laser radiation can have varying degrees of coherence.

Particularly, lasers with a low degree of spatial coherence have been of interest for their potential applications in imaging, due to high brightness and the absence of speckle noise [5].

The coherence properties of light produced within disordered media are determined by the internal scattering processes. Experimental determinations of the spatial coherence of random laser emission have been demonstrated by measuring the emission from active material powders [6], and liquid active media solutions with suspended scatterers [7]. Typically, the spatial coherence has been estimated by making use of either the single coordinate pair Young’s double-slit experiment [7], or speckle analysis of the output beam [8]. However, complete characterization requires the measurement of the two coordinate spatial coherence function, which can be found in Young’s experiment only by scanning the slits over all possible spatial coordinate combinations on the wavefront.

The novelty of our research is that we experimentally determine the two-coordinate coherence function for a quasi-random laser operating in pulsed mode with the use of a novel configuration of double-grating interferometer with a stationary camera and z-scanning gratings [9].

The quasi-random laser we utilize, is based on transparent wood (TW) as a semi-ordered host matrix, which was doped with Rhodamine 6G (Rh6G) dye as the gain medium. This type of TW laser was introduced recently by Vasileva et al. [10]. The TW is a composite material with excellent mechanical properties [11], structural anisotropy [11,12], as well as high transmittance (about 90 %) [12], with a simultaneously high haze value (up to 95 %) [13, 14], which indicates strong scattering within the substance. The scattering is caused by the fact that natural wood has a complex hierarchical structure scaling from nanometers to tens of micrometers (vessels,

(4)

fibers, rays, etc.) [15] and constituting of several components (cellulose, hemicellulose, lignin, etc.) of different refractive indices varying from 1.53-1.61 [16]. Chemical modification of wood (delignification) followed by infiltration with a polymer of matching refractive index is required to make natural wood transparent. After delignification, the refractive index difference of the inner components is reduced and the average refractive index of the template is around 1.53 [17]. To convert the TW host material into an optically active medium, a laser dye (Rh6G) is introduced in the material together with the infiltrating polymer. More details about the material used in this work including preparation and characterization can be found elsewhere [10].

The lasing action from the TW-Rh6G has been attributed to the collective effect of cellulose fibers working as an assembly of small Fabry-Perot type resonators [10] that are partially ordered due natural growth of internal wood components [11, 18]. On the other hand, since optical feedback is realized via scattering on the wood fiber boundaries and is not based on a classical external resonator, the emission has features of random lasing. Thus, we find it appropriate to define this type of source as a quasi-random laser.

2. Experimental setup

We consider the emission from three different samples, in order to make sure that the measured coherence properties are not caused by the measurement setup itself. For this, we need one non-lasing sample (low coherence), one lasing sample (higher coherence) and the TW-Rh6G sample (unknown). We prepared reference samples based on Rh6G in poly-methyl-metacrylate (PMMA) [19] and compare these results with the TW-Rh6G. The first reference sample was a simple PMMA slab with Rh6G and the second one was otherwise similar, but it had a gold mirror at one end. The two comparison samples are functionally the same Fabry-Perot type cavities, but the sample with a gold mirror has a higher quality factor, and thus lasing is easier to achieve with it.All the samples were of a rectangular shape, as is shown in Fig. 1, with facets polished to diminish output scattering. We always used the same side-pumping scheme for the samples, which is depicted in Fig. 1 (d). The pumping intensity in transversal direction decreased with the penetration depth due to the absorption of light by Rh6G molecules (cross-section inset in Fig. 1 (d)). Further on, we will refer to this feature to understand some of the experimental findings.

TW+Rh6G

Rh6G Rh6G

Au M

(b)

(a) (c)

Pumping

Output Pumped volume

cross-section

(d)

Fig. 1. Schematic representation of the experimental samples, (a) reference sample; (b) reference sample with a gold mirror (Au M) attached to a facet; (c) TW-dye sample; (d) schematics of the pumped area in the sample. The intensity distribution of the pump beam is nonuniform due to absorption and depth of focus.

We used the setup depicted in Fig. 2 for our experiments. A side-pumping scheme (Fig. 2 (a)) with the second harmonic of Nd:YAG (532 nm) at 0.70 mJ pulse energy was realized, where the active material was pumped using a 2 mm long and∼300µm wide line, and the pumping conditions were kept constant over all measurements. The lens L1 expanded the pump beam, and the square aperture A was used to sample the central part, where the spatial intensity distribution was nearly uniform. The cylindrical lens CL focused the pump light to a line, and the lens L2 collimated the output emission. To remove scattered pump light, a long-pass filter F was

(5)

implemented.

Luminescence emitted by the sample was then directed to a modified double grating interfer- ometer [9] (Fig. 2 (b)). It consisted of binary phase gratings G1and G2, electronic shutters ES1

and ES2, a microscope objective MO, and a camera C. The camera was a typical square-law CCD-array detector, with a broad spectral response. The light first passed through the binary grating G1, where it was split to several diffraction orders and the orders±1 were maximized by the grating design. All other orders were either blocked (Wbeing the zeroth order block) or propagated outside the measurement system. The two beams were then recombined by the second grating G2at the image plane, where they interfered and the interference pattern was imaged with a microscope objective onto the camera. By scanning the gratings in thez-direction, it was possible to shear the replicas of the emission with respect to each other, while keeping the overall propagation distance (from the source to the camera) approximately constant. Spatial coherence information was obtained from the visibility of the interference pattern.

(a) (b)

Nd:YAG 532 nm L1 A CL

Sample L2 F

x

y z

G1 G2

D1 D2

ES1

ES2

MO C

W

Fig. 2. (a) Pumping scheme, lens L1 expands the beam towards a square aperture A, after which the cylindrical lens CL focuses a line on the studied sample. The emission is collected by lens L2 and filter F removes any remaining pump light. (b) Interferometer, grating G1 disperses the beam and G2recombines±1 orders at the image plane, W being the zeroth order block. Electronic shutters ES1and ES2are used to find the intensity from individual arms, and the microscope objective MO images the interference pattern onto the detector C.

Since the sources we consider are quasi-monochromatic, we can ignore the spectral response of our system and the visibility of the interference fringes,V, in the grating interferometer is then related to the space–time domain degree of coherence as [9]

V(x,∆x)=2p

I0(x−∆x)I0(x+∆x)

I0(x−∆x)+I0(x+∆x) |γ0(x−∆x,x+∆x)|, (1) whereI0(x−∆x)andI0(x+∆x)are the intensities from different arms of the interferometer, andγ0(x−∆x,x+∆x)is the time–domain degree of spatial coherence in average and difference coordinates, xand ∆x, respectively. The spatial degree of coherence is a time independent quantity because the path lengths of the two beams are equal and the degree of spatial coherence is inherently measured as an ensemble over many pulses. In general, the interference pattern may fluctuate even when it is observed at a singlez-coordinate, which then smooths out if the measurement is done over a large number of pulses.

To conveniently quantify the results with a single numerical value, we define the overall degree of spatial coherence as [20]

γ2=

∬ I0(x−∆x)I0(x+∆x)|γ0(x−∆x,x+∆x)|2dxd∆x

∬ I0(x−∆x)I0(x+∆x)dxd∆x , (2) where the integration is taken over the whole beam andγmay take on values ranging from zero to unity, where zero signifies complete incoherence whereas unity is complete coherence. This is

(6)

a root-mean-square value of the absolute value of spatial coherence, giving high intensity areas more significance.

The displacement∆xbetween the two interfering fields is related to the scanning of the two gratings in thez-direction as

∆x(ω)=∆ztan

arcsin2πc

ωd ≈ 2πc∆z

ωd , (3)

wheredis the period of the observed interference fringes,cis the speed of light in vacuum andω is the frequency of light. Here we also neglect theωdependence of∆x, as the considered sources were found to be quasi-monochromatic based on the emission spectra that are presented later on.

To avoid excessive diffractive spreading, the length of the device was minimized, in accordance with [9]. By employing trigonometry and the grating equation, the parameters of the interferometer were chosen asD1 =50 mm andD2 ≈158 mm. For the other parameters, we setW =4 mm, so that the period of the first and second gratings becomed1 ≈7.94 µm andd2 ≈6.02 µm, respectively. The interference was imaged with a microscope objective MO – featuring 5× magnification – onto a Thorlabs DCC 1545M monochromatic camera. The fringe period was set tod = 12.5 µm, corresponding to ∼12 pixel widths on the camera and thus avoiding undersampling.

An important feature of the double grating interferometer is the absence of time delay between the two interfering beams, in which case temporal coherence of the emission does not affect spatial coherence measurements [9]. The two gratings should be parallel in the ideal case, but because the propagation direction of the diffracted orders is dependent on the angle of incidence, the system is also self correcting to a certain degree. The main advantages of using the grating interferometer instead of a Young’s interferometer with moving slits [21], are the speed and efficiency of the device, to avoid dye photobleaching during measurement. Significant photobleaching occurs after more than 50 000 shots, which is∼1.4 hours at 10 Hz repetition rate. Young’s interferometer has to scan over all possible combinations of spatial positions with its pinholes, meaning that the measurement is time consuming and has a low light efficiency. On the other hand, the employed grating interferometer features a light efficiency of∼32 %, and it measures all sets of points along thexdirection for a given∆xvalue. This allows for a measurement time of∼45 minutes in the present case.

It is worth mentioning that propagation through free space or optical components can modify the spatial coherence of light [22]. For example, light passed through a small aperture features increased spatial coherence, even if the input light is completely incoherent. In our experiments, we avoided these effects by sampling the emitted light with a large aperture and keeping the propagation length constant (the camera and sources had constant distance to each other), so that we could measure only the inherent properties of the light.

3. Results and discussion

The results for the emission from the first reference sample are depicted in Fig. 3. Its emission spectrum is narrowed to 12-14 nm (Fig. 3 (a)) in comparison to the typical 50 nm-width of the photoluminescent spectrum of Rh6G [23–26]. The narrowing of the emission line is caused by the influence of the low quality Fabry-Perot resonator formed by the sample facets, with a reflection of about 4 % [10]. As the optical amplification in the system is not enough to compensate the losses, amplified spontaneous emission (ASE) is observed. In fact, the cavity had such low quality that the lasing threshold would not be reached without considerably damaging the sample, and therefore, its value was not estimated. It is however, considerably higher than our pump energy of 0.70 mJ.

The measurement of the spatial coherence was performed over an ensemble of 10 pulses at each z-coordinate, and the single-shot interference pattern was observed to be stable. The measured

(7)

spatial coherence function of ASE was nearly incoherent, with the coherent area located mainly at zero separation (Fig. 3 (b)). The coherent area found here was produced by the propagation through the optical system, which is in agreement with the Van Cittert–Zernike theorem [22].

The overall degree of spatial coherence for this sample was found to beγ=0.11±0.01.

540 550 560 570 580 590 600

(a)

Rh6G

Wavelength [nm]

(b)

x[µm]

x[µm]

400 200 0 -200 -400

-400 -200 0 200 400

1.0

0.5

0.0

Fig. 3. PMMA + Rh6G sample below lasing threshold, (a) typical spectrum, (b) the measured spatial coherence,γ=0.11±0.01.

To increase the Q-factor of the Fabry-Perot cavity, a gold mirror was placed at the other facet of the Rh6G-PMMA sample as was depicted in Fig. 1 (b). With such design, optical losses in the cavity were reduced and the Q-factor of the resonator was increased in comparison with the first case. This reduced the threshold energy to around 0.44 mJ, which we could easily exceed.

Lasing peaks appeared at the background of an ASE spectrum, shown in Fig. 4 (a), although the pumping conditions were the same as for the previous sample.

The measurement of the spatial coherence was performed in a similar manner as before, with averaging over 10 pulses and no large fluctuations in the single shot interference patterns were observed. The measured spatial coherence function (Fig. 4 (b)) is clearly more coherent than the one measured for the low quality cavity, as can be seen from the prominent side-lobes, and with the overall degree of coherence beingγ=0.34±0.02. This coherence function has similarities to the broad area laser diode (BALD) [27, 28], which is a perfect example of multimode lasing.

540 550 560 570 580 590 600

(a)

Rh6G + M

Wavelength [nm]

(b)

x[µm]

x[µm]

400 200 0 -200 -400

-400 -200 0 200 400

1.0

0.5

0.0

Fig. 4. PMMA + Rh6G sample above lasing threshold, (a) typical spectrum, (b) the measured spatial coherence,γ=0.34±0.02.

As for the TW material, the internal structure is much more complicated. It is essentially a stack of cellulose fibers of various lengths (orders of mm) and diameters (in the range of 10−50µm).

In this medium, the cellulose fibers can act as small uncoupled Fabry-Perot resonators, which contribute to the output radiation. At the same time, the fibers are only semi-ordered, and strong scattering can occur at their boundaries, adding a certain degree of randomness to the emission.

(8)

Therefore, for careful interpretation of the experimental results, we first imaged the emission surface of the TW sample (Fig. 5(a)) by substituting the collimating lens (L2 in Fig. 2 (a)) with a microscope objective.

The majority of luminescent dye molecules are located on the walls of the cellulose fibers as is seen in Fig. 5 (a), creating a honeycomb-like structure, also reported in [10]. To examine if there is any significant correlations between the individual emitters (cellulose fibers), two copies of the near zone images were overlapped at the image plane by using the grating interferometer.

The produced interference patterns at zero (Fig. 5 (b)) and∼10 µm (Fig. 5 (c)) separations were recorded. The interference fringes rapidly disappear with increasing separation distance, and therefore show that the emitters are mutually uncorrelated. This type of behavior is also reminiscent of BALD, where several mutually uncorrelated cavity modes along the wide axis [28]

cause a complicated spatial coherence function [9, 27].

(a) (b) (c)

x

y

100µm 50µm

High

arb.units

Low

Fig. 5. Images of the emission surface, (a) the intensity distribution where some of the individual cells are visible, (b) portion of the image interfered with itself at zero separation, and (c) at a separation of∼10µm.

To perform the spatial coherence measurements for the TW output emission, the collimating lens L2 was placed back into the setup instead of the microscope objective. Again, the measurements were averages over 10 pulses, and almost no fluctuations were found from single-shot interference patterns. The measured values for the emission from the TW sample are shown in Fig. 6. We estimate the threshold energy to be about 0.54 mJ, which is slightly higher than for the Rh6G sample with gold mirror, but is still easily exceeded in our experiment. The main difference between coherent and incoherent feedback in random lasers is the presence or absence of narrow spectral lines. In our case, the narrow peaks are absent from the output emission (Fig. 6 (a)) and the spectral line is broadened (FWHM≈6 nm) due to collective lasing action by a set of uncorrelated wood fibers. Thus, the spectrum indicates quasi-random laser action with incoherent feedback.

The measured spatial coherence function is displayed in Fig. 6 (b). It is clear from these results that the spatial coherence function has some multimode structure, and that the overall degree of coherence is between the values measured for the two reference samples, which was found to be γ=0.16±0.01. This intermediate case is clearly seen in Fig. 6 (c), where the comparison of coherence functions of all three cases is presented.

(9)

540 550 560 570 580 590 600

(a)

TW + Rh6G

Wavelength [nm]

400 200 0 -200 -400

-400 -200 0 200 400

(b)

x[µm]

x[µm]

1.0

0.5

0.0

|γ0(∆x)|

x[µm]

TW + Rh6G Rh6G Rh6G + Au

(c)

0.0 0.2 0.4 0.6 0.8 1.0

-400 -300 -200 -100 0 100 200 300 400

Fig. 6. TW above lasing threshold, (a) typical spectrum, (b) the measured spatial coherence, γ=0.16±0.01, and (c) a comparison of the coherence from all three samples corresponding to the positionx=−200µm.

Comparison with the reference samples shows that the measurement setup does not significantly modify the coherence of the radiation, since in the opposite case the spatial coherence would yield systematically similar shape and overall degree of coherence for all of the considered samples. Thus, it is safe to conclude that the measured spatial coherence functions characterized intrinsic properties of the sources themselves.

One immediately sees that for the second reference sample (Fig. 4 (b)) and TW (Fig. 6 (b)), the coherence function has lower values forx>0, which is caused by the pump beam absorption in the samples. The pump beam intensity exponentially decreases while entering the active medium, as was shown in Fig. 1 and the varying distribution of the pump power has been shown to affect the properties of the output light [29]. The output consists of laser emission from areas with high pump intensities (the side of pump incidence) and ASE and photoluminescence from less pumped areas (away from the pump). In other words, the output beam features high (laser emission) and low (ASE and photoluminescence) coherence areas, with the degree of coherence decreasing across the beam along the direction of the pump.

4. Conclusions

To summarize, we have demonstrated experimental determination of the two coordinate spatial coherence function for a quasi-random laser based on a TW-Rh6G material which has neither perfectly ordered nor random structure. We have used double-grating interferometer of modified configuration with a stationary camera and z-scanning gratings in the short pulse regime.

The presented technique applied for this type of source can open new opportunities in the characterization of random lasers. The nature of the emission produced by gain media with randomly distributed scatters has been under discussion, and measuring the two-coordinate spatial coherence function can be used as a fundamental parameter to distinguish between ASE and lasing, as we have shown here.

The emitted light from the TW sample was found to be a mixture of laser emission produced

(10)

by a number of uncorrelated laser cavities on the strongly pumped side of the sample, and low coherence light in the areas of less intense pumping. These types of quasi-random lasers can be constructed from any semi-ordered material that features optical gain, with coherence properties determined by the ordering of the internal structure. It is conceivable that such sources can offer novel solutions in a wide range of imaging and lighting applications.

Funding

University of Eastern Finland (4900024); Swedish Research Council (621-2012-4421); European Research Council Advanced Grant (742733); European Research Council Advanced Grant No 742733, Wood NanoTech.

Viittaukset

LIITTYVÄT TIEDOSTOT

Öljyn kokonaiskäyttö kasvaa kaikissa skenaarioissa hieman vuoteen 2010 mennessä mutta laskee sen jälkeen hitaasti siten, että vuonna 2025 kulutus on jo nykytason ala-

Jos valaisimet sijoitetaan hihnan yläpuolelle, ne eivät yleensä valaise kuljettimen alustaa riittävästi, jolloin esimerkiksi karisteen poisto hankaloituu.. Hihnan

If the distance from SPP creation to the nanostripe N is small compared to the plasmon survival lengths, the principal features of the spatial coherence do not effec- tively depend

Työn merkityksellisyyden rakentamista ohjaa moraalinen kehys; se auttaa ihmistä valitsemaan asioita, joihin hän sitoutuu. Yksilön moraaliseen kehyk- seen voi kytkeytyä

The new European Border and Coast Guard com- prises the European Border and Coast Guard Agency, namely Frontex, and all the national border control authorities in the member

The US and the European Union feature in multiple roles. Both are identified as responsible for “creating a chronic seat of instability in Eu- rope and in the immediate vicinity

Indeed, while strongly criticized by human rights organizations, the refugee deal with Turkey is seen by member states as one of the EU’s main foreign poli- cy achievements of

A Bruker IFS-120 high-resolution Fourier transform interferometer has also been employed to spectrally analyse the fluorescence spectra in the emission study and the laser output in