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

New method to determine the refractive index and the absorption coefficient of organic nonlinear crystals in the ultra-wideband THz region

Open Access Open Access

Abstract

A method for simultaneously measuring the refractive index and absorption coefficient of nonlinear optical crystals in the ultra-wideband terahertz (THz) region is described. This method is based on the analysis of a collinear difference frequency generation (DFG) process using a tunable, dual-wavelength, optical parametric oscillator. The refractive index and the absorption coefficient in the organic nonlinear crystal DAST were experimentally determined in the frequency range 2.5–26.2 THz by measuring the THz-wave output using DFG. The resultant refractive index in the x-direction was ~2.3, while the absorption spectrum was in good agreement with FT-IR measurements. The output of the DAST-DFG THz-wave source was optimized to the phase-matching condition using the measured refractive index spectrum in THz region, which resulted in an improvement in the output power of up to a factor of nine.

©2010 Optical Society of America

1. Introduction

THz-wave radiation has been used in several fields of science, including physics, chemistry, biology, medicine, and environmental science. In THz technology, a monochromatic source is an extremely important device and has enabled many applications. Ito et al. have developed tunable monochromatic THz-wave sources based on nonlinear optical processes. These systems use optical parametric oscillations (OPO) [1,2] and difference-frequency generation (DFG) [3,4]. They have achieved THz-wave sources that are widely tunable and frequency agile by using organic nonlinear crystals for the DFG process.

In these nonlinear processes, high output power THz radiation is expected when the phase-matching condition is met and the THz-wave gain is higher than the loss. Therefore, the refractive index and absorption coefficient in the THz region, which are related to the phase-matching condition and losses in a nonlinear material, are very important. These optical properties provide us the information not only to optimize the pump wavelength and the suitable crystal thickness for THz generation [5], but also to assign the inter- and intramolecular vibration modes in organics [6].

Optical crystals of 4-dimethylamino-N-methyl-4-stilbazolium-tosylate (DAST) [7] are remarkable in their ability to generate THz waves due to the high optical nonlinearity [8] and low refractive index dispersion in the THz and near-infrared regions [911]. Moreover, wideband THz-wave generation has been reported [12,13]. The refractive index of DAST has been studied in the near-infrared region [9], and the refractive index and absorption in the region 0–3 THz using a terahertz time-domain spectroscopy [14]. Directly measuring the reflective index in the wideband from the THz-region to the mid-infrared region is difficult because of the large absorption coefficient in DAST, and the absence of a tunable source or detector in this frequency region. In this paper, we report the development of a simultaneous method to obtain both the refractive index and the absorption coefficient in a nonlinear optical crystal in the ultra-wideband THz region using a collinear DFG process. Our method is based on the fact that the output power from the collinear DFG processes depends strongly on the refractive index, absorption, and the wavelengths of the input and output electromagnetic waves.

2. Theory

Consider the situation where two pump beams with wavelengths λ1 and λ2, where λ1 < λ2, are collinearly incident on a nonlinear crystal of thickness L. The THz-wave radiation with wavelength λTHz is generated in the crystal by a DFG process. The functional form of the output power from the process is given by [11,12]

S=Aexp(αTHzL)1+exp(ΔαL)2exp(12ΔαL)cos(ΔkL)(ΔkL)2+(12ΔαL)2,
Δk=k1k2kTHz=2π(nλ1λ1nλ2λ2nTHzλTHz),
Δα=|α1α2αTHz|,
where A is the proportional constant. The refractive indices for each wavelength are n λ1, n λ2, and n THz, and the absorption coefficients are αλ1, αλ2, and αTHz. The wavelengths λ1 and λ2 can be measured (e.g., by using an optical spectrum analyzer) and then λTHz can be determined from energy conservation for each angular frequency, ω 1 = ω 2 + ω THz. Therefore, if n λ1, n λ2, αλ1, and αλ2 are known, one can estimate n THz and αTHz from the spectral shape of S1, λ2).

Here, we consider the case of THz-wave generation using a DAST crystal as reported in Ref. 3. In the optical region, the refractive index, n, of DAST has been established by F. Pan et al. [9] Therefore, n λ1 and n λ2 are known in this case. Because the absorption coefficient in THz frequency region is very large (α > 100 cm−1, as shown in FT-IR measurement in experimental section), we assume that the absorption in the optical region can be ignored in Eq. (3), yielding ΔααTHz. So, in the DAST case, one can obtain n THz and αTHz from the two-dimensional measurement of S1, λ2). Since we are considering a nonlinear process in which three wavelengths are coupling with the component, d 111, in the nonlinear coefficient tensor of the DAST, n λ1 and n λ2 in the x-direction were used for the calculation and n THz and αTHz in the same direction were obtained.

Note that, with a traditional method (e.g., THz-TDS) one must prepare thin samples to measure the refractive index and the absorption in opaque materials; however, with this method, controlling the sample thickness is not necessary because these are given within the functional form factor of the peak shift and line width of S1, λ2). Furthermore, one need not consider the Fresnel loss at the sample surfaces to obtain the absorption coefficient. This method can be applied not only to DFG processes, but also to other nonlinear processes including sum frequency generation and higher order processes.

3. Experimental

The collinear DFG and detection experimental setup is shown in Fig. 1 . It was similar to that of the broad tunable THz-wave source reported earlier by Ito et al. [4] λ1 and λ2 are generated in a dual-wavelength source (KTP-OPO), which was pumped by pulsed 532-nm radiation from a frequency-doubled Nd:YAG laser (8 ns, 100 pps, 95 mJ/pulse). KTP-OPO is based on a single resonance-type OPO in which the signal wave (762–921 nm) is confined by mirrors M1 and M2. M1 is coated for high reflectivity of the signal waves and high transmittance of the 532 nm and idler waves (1260–1760 nm), M2 is a broadband mirror, and M3 is coated for high reflection of idler waves but high transmittance at 532 nm. Since the two KTP crystals (a) and (b) are mounted on galvano scanners in one cavity, one can independently select a pair of wavelengths λa and λb from 1260 to 1760 nm by setting the incident angles for the pump beam of each crystal. The shorter of the wavelengths λa and λb is designated λ1 and the longer λ2.

 figure: Fig. 1

Fig. 1 Experimental setup used to measure DFG in DAST for arbitrary λa and λb in the range 1260–1760 nm.

Download Full Size | PDF

The two waves from the KTP-OPO propagate onto the c-axis of a 1-mm-thick DAST crystal. In the DAST crystal, the THz wave is generated by a type-zero DFG process in which the polarization direction of every wave is along to the a-axis of the DAST crystal. The THz waves pass though the low-pass filter to clip λ1 and λ2 and are then collimated and focused onto a 4 K Si bolometer by a pair of off-axis parabolic mirrors. Note that the maximum measurable frequency was limited to 30 THz due to the sensitivity of the bolometer; other types of detectors (e.g., pyro-detector) could potentially improve the range.

4. Results and discussion

The DFG output power was measured while varying λa and λb independently from 1260 to 1760 nm. Figure 2 (a) shows a two-dimensional mapping of the output power. The output power was normalized to the maximum at each frequency difference, and the frequency dependence comes from the dispersion of the refractive index in the THz region. The width of the absorption band also depends on the THz frequency. The measured area in Fig. 2 was limited by both the range in rotation angle of the KTP crystals in the OPO cavity and the frequency sensitivity of the bolometer used as a detector. The output spectra were measured with good signal-to-noise ratios in the frequency regions of 2.5–20 THz and 22.5–26.2 THz. The horizontal fine structure in Fig. 2(a) results from the power fluctuation of the KTP-OPO, and is a consequence of optical cavity modes of the crystal. Figure 2(b) shows the spectral mapping with this fine structure removed by applying a low intensity threshold to the data.

 figure: Fig. 2

Fig. 2 Output mapping of DFG power. (a) Power was normalized at each frequency (ωa – ωb)/2π . (b) The low-power region caused by the cavity mode of the KTP-OPO was removed

Download Full Size | PDF

To obtain the refractive index and absorption coefficient as a function of the THz frequency, the cross sections at each frequency difference on the mapping were analyzed by applying a least square fit to Eq. (1). For the fitting, the refractive indices n λ1 and n λ2 in the x-direction were given by the Sellmeier formula using parameters from Ref. 8 and L = 1 mm was substituted into Eq. (1) and A, n THz and α THz were adjustable. Figure 3 shows an example of this fitting; the spectral shift and line width were replicated by the fitting function. The refractive index and absorption coefficient at 18.9 THz were n THz = 2.295 ± 0.001 and α THz = 125 ± 10 cm−1, respectively. Since the accuracy of refractive index has influence to the spectral shift, it was depended on the accuracy of n λ1 and n λ2, which were given to four significant figures in Ref. 8, the wavelength determination accuracy of the KTP-OPO and the uncertainty of the fitting. The accuracy of absorption coefficient is linked to the spectral line width of the output function and is limited by the accuracy of L and the broadband flatness of the KTP-OPO output power in addition to the factors for refractive index. In our case, the uncertainty of the fitting was the most dominant in these limiting factors. Thus, the error was estimated from the statistical uncertainty of the least square fit in which we took the confidence interval to 95%. The same fitting process was performed at all frequencies.

 figure: Fig. 3

Fig. 3 An example of the fit at (ωa – ωb) / 2π = 18.9 THz. The red points are THz-wave output experimentally obtained. The green line is a regression curve to Eq. (1).

Download Full Size | PDF

Figure 4(a) shows the results for the refractive index as best-fit coefficients for all frequencies. The positive and negative difference frequency points both oscillated around n = 2.3, and the functional form follows that of the absorption coefficient. The refractive index obtained using the new method is in satisfactory agreement with the data from Ref. 14, which was measured using a THz time-domain spectroscopic method. Figure 4(b) shows results for the absorption coefficient, together with the absorption coefficient spectrum along the x-axis of the DAST crystal measured using an FT-IR spectrometer. Because of the large absorption of a DAST in the THz region, for the FT-IR measurement we thinned the DAST crystal to 0.25 mm using a special turning lathe [15]. The Fresnel losses at the surfaces of the DAST crystal were considered due to the refractive index of 2.3. The absorption spectra were in good agreement, both in terms of functional form and magnitude. The absorption coefficient at around 15 THz was too large to measure using FT-IR, however, could be measured using the present method. We also took into account that the spectral resolution of the refractive index and the absorption coefficient were reduced by the 180-GHz spectral line width of the KTP-OPO output. The discrepancy of absorption coefficient between the present method and the FT-IR measurement was rather large in between from 10 to 20 THz. We consider that this is due to the assumptions that we neglected the absorption of λ1 and λ2 in the DAST for the present method and that we also neglected the dispersion of the refractive index to obtain the Fresnel losses for the FT-IR measurement. In the range, 20–23 THz the absorption coefficient is large, and the signal-to-noise ratio is too low to make reliable measurements. For this reason, there is a gap in the data in Fig. 4. The shaded region in Fig. 4(a) shows the frequency range where we cannot determine the refractive index using the difference frequency system. However, the refractive index is within these limits for almost all frequencies. Following the publication of this work, the numerical data will become available on the THz-database [16].

 figure: Fig. 4

Fig. 4 Frequency dependence of the (a) refractive index and (b) absorption along the x-axis in a DAST crystal. Red (Blue) points were obtained from the region in which ωa – ωb was positive (negative). The gray curve in (a) was taken from Ref. 14. In our method, the refractive index can be obtained if the value is within the un-shaded area. The gray curve in (b) was measured using an FT-IR spectrometer.

Download Full Size | PDF

The refractive index spectra of nonlinear crystals in the THz region make it possible to optimize the phase matching condition at all frequencies5. This can be achieved by controlling the phase mis-match, Δk, which appears in Eq. (2) and which depends on n THz. Figure 5 shows a comparison of the normalized THz output power spectrum with and without the optimized phase matching. The latter in which one wavelength was fixed to 1.3 μm as provided in Ref. 3 was useful for achieving the wide tunability and easy controlling of KTP-OPO, but had disadvantage in the high power output. An increase in the output power of up to a factor of nine was observed.

 figure: Fig. 5

Fig. 5 THz output of DAST-DFG as a function of frequency. The red curve shows optimized output for phase matching condition by controlling λ1 and λ2. The gray curve shows the output with λ1 = 1.3 μm.

Download Full Size | PDF

5. Conclusion

We have developed a method to measure the refractive index and absorption coefficient of nonlinear crystals in the THz frequency range. Using this technique, we measured a DAST crystal, and characterized n and α in the range 2.5–26 THz. The frequency-dependent refractive index data was used in improve the phase matching condition for DFG, and an improvement in the output power of up to a factor of nine was observed. This method is, in principle, applicable to all nonlinear crystals that can be used for DFG of THz-wave radiation.

Acknowledgment

The authors thank A. Nawahara for his valued assistance in the experimental procedures. We also thank C. Takyu, T. Shoji, and Y. Konno, who respectively performed the dielectric and metal coating, the polishing of the various optical components, and the turning of the DAST crystals. This work was partly supported by a Grant-in-Aid for Scientific Research (A) from the Japan Society for the Promotion of Science (JSPS) (No. 19206009) and a Research for Promoting Technological Seeds for 2008 from the Japan Science and Technology Agency (JST) (No. 03-036).

References and Links

1. K. Kawase, J. Shikata, and H. Ito, “Terahertz wave parametric source,” J. Phys. D Appl. Phys. 35(3), 201 (2002). [CrossRef]  

2. H. Minamide, T. Ikari, and H. Ito, “Frequency-agile terahertz-wave parametric oscillator in a ring-cavity configuration,” Rev. Sci. Instrum. 80(12), 123104 (2009). [CrossRef]  

3. H. Ito, K. Suizu, T. Yamashita, A. Nawahara, and T. Sato, “Random frequency accessible broad tunable terahertz-wave source using phase-matched 4-dimethylamino-n-methyl-4-stilbazolium tosylate crystal,” Jpn. J. Appl. Phys. 46(11), 7321–7324 (2007). [CrossRef]  

4. K. Miyamoto, H. Minamide, M. Fujiwara, H. Hashimoto, and H. Ito, “Widely tunable terahertz-wave generation using an N-benzyl-2-methyl-4-nitroaniline crystal,” Opt. Lett. 33(3), 252–254 (2008). [CrossRef]   [PubMed]  

5. K. Miyamoto, S. Ohno, M. Fujiwara, H. Minamide, H. Hashimoto, and H. Ito, “Optimized terahertz-wave generation using BNA-DFG,” Opt. Express 17(17), 14832–14838 (2009). [CrossRef]   [PubMed]  

6. M. Walther, B. Fischer, M. Schall, H. Helm, and P. Uhd Jepsen, “Far-infrared vibrational spectra of all-trans, 9-cis and 13-cis retinal measured by THz time-domain spectroscopy,” Chem. Phys. Lett. 332(3-4), 389–395 (2000). [CrossRef]  

7. H. Nakanishi, H. Matsuda, S. Okada, and M. Kato, “Organic polymeric ion-complexes for nonlinear optics”, MRS International Meeting on Advanced Materials Proceedings1, 97 (1989).

8. U. Meier, M. Bösch, C. Bosshard, F. Pan, and P. Günter, “Parametric interactions in the organic salt 4-N-N-dimethylamino-4’-N’- methyl-stilbazolium tosylate at telecommunication wavelengths,” J. Appl. Phys. 83(7), 3486 (1998). [CrossRef]  

9. F. Pan, G. Knöpfle, Ch. Bosshard, S. Follonier, R. Spreiter, M. S. Wong, and P. Günter, “Electro‐optic properties of the organic salt 4‐N,N‐dimethylamino‐4’‐N’‐methyl‐stilbazolium tosylate,” Appl. Phys. Lett. 69(1), 13 (1996). [CrossRef]  

10. A. Schneider, M. Neis, M. Stillhart, B. Ruiz, R. U. A. Khan, and P. Günter, “Generation of terahertz pulses through optical rectification in organic DAST crystals: theory and experiment,” J. Opt. Soc. Am. B 23(9), 1822 (2006). [CrossRef]  

11. K. Kawase, M. Mizuno, S. Sohma, H. Takahashi, T. Taniuchi, Y. Urata, S. Wada, H. Tashiro, and H. Ito, “Difference-frequency terahertz-wave generation from 4-dimethylamino-N-methyl-4-stilbazolium-tosylate by use of an electronically tuned Ti:sapphire laser,” Opt. Lett. 24(15), 1065–1067 (1999). [CrossRef]  

12. T. Taniuchi, S. Okada, and H. Nakanishi, “Widely tunable terahertz-wave generation in an organic crystal and its spectroscopic application,” J. Appl. Phys. 95(11), 5984 (2004). [CrossRef]  

13. M. Ashida, R. Akai, H. Shimosato, I. Katayama, T. Itoh, K. Miyamoto, and H. Ito, “Ultrabroadband THz Field Detection beyond 170THz with a Photoconductive Antenna,” CLEO/QELS( 2008), CTuX6.

14. M. Walther, K. Jensby, S. R. Keiding, H. Takahashi, and H. Ito, “Far-infrared properties of DAST,” Opt. Lett. 25(12), 911–913 (2000). [CrossRef]  

15. Y. Namba, M. Tsukahara, A. Fushiki, K. Suizu, and H. Ito, “Single-point diamond turning of DAST crystals“, Proceedings of SPIE 5180 Optical Manufacturing and Testing V, 55 (2003)

16. Terahertz Database, http://www.thzdb.org

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 (5)

Fig. 1
Fig. 1 Experimental setup used to measure DFG in DAST for arbitrary λa and λb in the range 1260–1760 nm.
Fig. 2
Fig. 2 Output mapping of DFG power. (a) Power was normalized at each frequency (ωa – ωb)/2π . (b) The low-power region caused by the cavity mode of the KTP-OPO was removed
Fig. 3
Fig. 3 An example of the fit at (ωa – ωb) / 2π = 18.9 THz. The red points are THz-wave output experimentally obtained. The green line is a regression curve to Eq. (1).
Fig. 4
Fig. 4 Frequency dependence of the (a) refractive index and (b) absorption along the x-axis in a DAST crystal. Red (Blue) points were obtained from the region in which ωa – ωb was positive (negative). The gray curve in (a) was taken from Ref. 14. In our method, the refractive index can be obtained if the value is within the un-shaded area. The gray curve in (b) was measured using an FT-IR spectrometer.
Fig. 5
Fig. 5 THz output of DAST-DFG as a function of frequency. The red curve shows optimized output for phase matching condition by controlling λ1 and λ2. The gray curve shows the output with λ1 = 1.3 μm.

Equations (3)

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

S = A exp ( α THz L ) 1 + exp ( Δ α L ) 2 exp ( 1 2 Δ α L ) cos ( Δ k L ) ( Δ k L ) 2 + ( 1 2 Δ α L ) 2 ,
Δ k = k 1 k 2 k T H z = 2 π ( n λ 1 λ 1 n λ 2 λ 2 n T H z λ T H z ) ,
Δ α = | α 1 α 2 α T H z | ,
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.