Generation of Quasi-Continuous-Wave Vacuum-Ultraviolet Coherent Light by Fourth-Harmonic of A Ti:sapphire Laser With KBBF Crystal

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Generation of quasi-continuous-wave vacuum-

ultraviolet coherent light by fourth-harmonic of


a Ti:sapphire laser with KBBF crystal
Yun Zhang,1* Yusuke Sato,1 Nobuyoshi Watanabe,1 Riskey Ananda1,
Yoshiko Okada-Shudo,1 Masayoshi Watanabe,1 Masaharu Hyodo,2 Xiaoyang Wang,3
Chuangtian Chen,3 Teruto Kanai,4 and Shuntaro Watanabe4
1
Department of Electronic Engineering, The University of Electro-Communications, 1-5-1 Chofugaoka, Chofu-shi,
Tokyo 182-8585, Japan
2
Kobe Advanced Research Center, National Institute of Information and Communications Technology, 588-2 Iwaoka,
Nishi-ku, Kobe, Hyogo 651-2492, Japan
3
Technical Institute of Physics and Chemistry, China Academy of Science, P.O. Box 2711, Beijing 100080, China
4
Institute for Solid State Physics, University of Tokyo, 5-1-5 Kashiwanoha, Kashiwa-shi, Chiba 277-8581, Japan
*
Corresponding author: zhang@ee.uec.ac.jp

Abstract: We report the generation of quasi-continuous-wave vacuum-


ultraviolet (VUV) coherent light based on a Ti:sapphire laser with two
successive frequency doubling stages. In the first stage, UV light at 399 nm
with power of 1.1 W was obtained by exploiting an enhanced cavity. With a
KBBF crystal as nonlinear material, quasi-continuous-wave VUV coherent
light with power of about 25 mW at 199.5 nm and 4.7 mW at 193.5 nm
were achieved through a single-pass SHG configuration, respectively, in the
second stage.
2009 Optical Society of America
OCIS codes: (190.2620) Harmonic generation and mixing; (140.3515) Lasers, frequency
doubled; (140.7240) UV, EUV, and X-ray lasers

References and links


1. R. W. Boyd, Nonlinear Optics (Third version), Academic Pr. (2008).
2. M. Watanabe, K. Hayasaka, H. Imajo, and S. Urabe, “Continuous-wave sum frequency generation near 194
nm in BBO crystals with an enhancement cavity,” Opt. Lett. 17, 46-48 (1992).
3. C. T. Chen, Y. B. Wang, B. C. Wu, K. C. Wu, W. L. Zeng, and L. H. Yu, “Design and synthesis of an
ultraviolet-transparent nonlinear-optical crystal KBBF,” Nature 373, 322-325 (1995).
4. C. T. Chen, G. L. Wang, X. Y. Wang, Y. Zhu, Z. Y. Xu, T. Kanai, and S. Watanabe, “Improved sellmeier
equations and phase matching characteristics in deep ultraviolet region of KBBF crystal,” IEEE J.
Quantum Electron. 44, 617-621 (2008).
5. T. Kanai, T. Kanda, T. Togashi, T. Sekikawa, S. Watanabe, C. Chen, C. Zhang, Z. Y. Xu, and J. Wang,
“Generation of vacuum ultraviolet light and measurement of phase matching angles in KBBF crystal,”
presented at the Conference on Lasers and Electro-Optics/Quantum Electronics and Laser Science
Conference, Technical Digest (Optical Society of America, 2003), paper CTuM1.
6. T. Kanai, T. Kanda, T. Sekikawa, S. Watanabe, T. Togashi, C. T. Chen, C. Q. Zhang, Z. Y. Xu, and J. Y.
Wang, “Generation of vacuum-ultraviolet light below 160 nm in a KBBF crystal by the fifth harmonic of a
single-mode Ti:sapphire laser,” J. Opt. Soc. Am. B 21, 370-375 (2004).
7. C. T. Chen, T. Kanai, X.Y. Wang, Y. Zhu, and S. Watanabe, “High-average-power light source below 200
nm from a KBe2BO3F2 prism-coupled device,” Opt. Lett. 33, 282-284 (2008).
8. G. L. Wang, X. Y. Wang, Y. Zhou, C. M. Li, Y. Zhu, Z. Y. Xu, and C. T. Chen, “High-efficiency
frequency conversion in deep ultraviolet with a KBBF prism-coupled device,” Appl. Opt. 47, 486-488
(2008).
9. G. L. Wang, X. Y. Wang, Y. Zhou, Y. Chen, C. Li, Y. Zhu, Z. Y. Xu, and C. T. Chen, “12.95 mW sixth
harmonic generation with KBBF crystal,” Appl. Phys. B 91, 95-97 (2008).
10. Y. Zhu, G. L. Wang, C. M. Li, Q. J. Peng, D. F. Cui, Z. Y. Xu, X. Y. Wang, Y. Zhu, and C. T. Chen, “Sixth
harmonic of a Nd:YA4 laser generation in KBBF for ARPES,” Chin. Phy. Lett. 25, 963-965 (2008).
11. H. Zhang, G. L. Wang, L. Guo, A. Geng, Y. Bo, D. Cui, Z.Y. Xu, R. Li, X. Wang, and C. T. Chen, “175 to
210 nm widely tunable deep-ultraviolet light generation based on KBBF crystal,” Appl. Phys. B 93, 317-
320 (2008).

#108967 - $15.00 USD Received 19 Mar 2009; revised 21 Apr 2009; accepted 23 Apr 2009; published 29 Apr 2009
(C) 2009 OSA 11 May 2009 / Vol. 17, No. 10 / OPTICS EXPRESS 8119
12. E. S. Polzik and H. J. Kimble, “Frequency doubling with KNbO3 in an external cavity,” Opt. Lett. 16, 1400-
1402 (1991).
13. M. Watanabe, R. Ohmukai, K. Hayasaka, H. Imajo, and S. Urabe, “High-power second harmonic
generation with picosecond and hundreds –of –picosecond pulses of a cw mode-locked Ti:sapphire laser,”
Opt. Lett. 19, 637-639 (1994).
14. T. Togashi, T. Kanai, T. Sekikawa, S. Watanabe, C. T. Chen, C. Q. Zhang, Z. Y. Xu, and J. Y. Wang,
“Generation of vacuum-ultraviolet light by an optically contacted prism-coupled KBBF crystal,” Opt. Lett.
28, 254-256 (2003).
15. K. Hayasaka, Y. Zhang, and K. Kasai, “Generation of 22.8 mW single-frequency green light by frequency
doubling of a 50-mW diode laser,” Opt. Express 12, 3567-3572 (2004)

1. Introduction
Vacuum-ultraviolet (VUV) coherent light has been proposed for different applications, as
optical data storage, metrology, biomedical application, fundamental spectroscopic research,
and laser lithography. Conventional coherent VUV laser sources at this wavelength are
excimer lasers (ArF and KrF). These lasers generate a high output power of more than 100 W,
however their structures are huge and complex. Moreover, high manufacturing and
maintenance costs are required. The low repetition rate (usually several kHz) of such excimer
laser systems also restricts the applicability of many shot statistics during data acquisition in
spectroscopic measurements. Another way to produce coherent radiation in this wavelength
range consists in the optical frequency conversion of solid-state laser in nonlinear optical
crystals, which are transparent in the VUV, such as BBO, LBO or other borates. The
development of coherent radiation with high-repetition rate (quasi-continuous wave) or
continuous wave (cw) in VUV has to be based on the frequency conversion or exploitation of
new laser materials. At present, the later is a more challenging topic and the former is the
unique way to generate VUV laser light with high-repetition rate or continuous wave.
Generally, techniques of optical frequency conversion for generating short wavelength light
are second harmonic generation (SHG) and sum frequency generation (SFG) process [1].
There are several commercial nonlinear optical crystals that are capable of producing VUV
coherent light, such as LBO and BBO, through SFG [2]. But these crystals cannot be used to
produce light at wavelength shorter than 200 nm through SHG because it is impossible to
achieve the phase-matching angle of SHG. Compare to the setup of SHG, the optical setup of
SFG required multiple laser sources; consequently, the system was slightly complex and
multiple optical beams coupling also makes system at low efficiency.
Thanks the development of material science, the new excellent nonlinear material, which
is named KBe2BO3F2 (KBBF) [3], makes it possible to generate light at wavelength shorter
than 200 nm through SHG. Based on Sellmeier equations, it was predicted that VUV light at
wavelength as short as 165 nm could be generated through direct SHG with KBBF as
nonlinear crystal [4]. In experiment, VUV light at wavelength of 170 nm has been obtained
through a direct SHG with a KBBF crystal [5]. Meanwhile, VUV light at wavelength of as
short as 157 nm has also been produced by fifth-harmonic generation of a Ti:sapphire laser
with a KBBF as nonlinear crystal [6]. In the past decade, both the generated average-power
and conversion efficiency have been significant improved [7]. However, the quality of the
KBBF crystal severely restricted the laser sources to pulse lasers especially with low
repetition rate benefit from the high-peak power of the each pulse. Recently, the research
began to focus in the generation of quasi continuous wave VUV light with the development of
the KBBF crystal and prism coupling technique (PCT) [8-11].
Enhanced resonant cavity is another most important technique for an increase in frequency
conversion efficiency, especially for continuous wave lasers [12] and mode-locked pulse
lasers [13]. Without additional costs in terms of laser power, the intensity power can be
coherently boosted in the cavity; thus providing much average or higher peak power for more
efficient energy conversion. Conversion efficiencies of more than 70% have been obtained for
frequency doubling both the continuous wave laser and mode-locked pulse lasers by
exploiting an enhanced cavity. Our group is now involving the generation of more than 100
mW quasi-continuous wave VUV laser light with two successive SHG stages in the enhanced

#108967 - $15.00 USD Received 19 Mar 2009; revised 21 Apr 2009; accepted 23 Apr 2009; published 29 Apr 2009
(C) 2009 OSA 11 May 2009 / Vol. 17, No. 10 / OPTICS EXPRESS 8120
resonant cavity, in which LBO and KBBF are used as the nonlinear medium respectively,
from a mode-locked Ti:sapphire laser. In this paper, we report on the first step of this project,
namely, the realization of frequency doubling of Ti:sapphire laser with a enhanced resonant
cavity and generation of VUV coherent light by a single pass through SHG system with
KBBF as nonlinear material. Best to our knowledge, quasi-continuous-wave VUV with power
of about 25 mW is the highest power at present by means of frequency conversion.
2. Experimental setup

Ti:sapphire laser
82 MHz, 1.5 ps

PZT M1

M2
QWP
M4 M3
D1
LBO

PBS
D2

KBBF-PCT device DA
Power Meter
Servo Loop
Fig. 1. Experimental setup for generation VUV with KBBF crystal.

In Fig. 1 we show a schematic diagram of our experiment setup. The fundamental beam is
provided by a mode-locked pulsed Ti:sapphire laser, whose center wavelength was 798 nm.
The average output power was typically about 1.7 W. The output of Ti:sapphire laser is
passed through a lenses, which was used to manipulate the q-parameter of the fundamental
beam before the resonant cavity, and introduced into the enhanced SHG cavity. There is a
general limitation when one uses a mode-locked laser. The free spectral range of the
enhancement cavity must match (or be a subharmonic of) the free spectral range of the laser
cavity; it is not a free parameter. In our experiment we used a laser that emitted 1.5 ps pulses
at an 82 MHz repetition rate. So the cavity length was first set at approximately 3656 mm, and
then finely adjusted to the point of maximum enhancement. The cavity, as the same as the
enhanced cavity in Ref. 13, consists of two flat mirrors (M1 and M2) and two curve mirrors
(M3 and M4). M1 was a coupling mirror, whereas M2-M4 were coating for high reflectivity
at the fundamental wavelength. Several different input couplers, with transmission values
between 5% and 30%, were used to couple the fundamental light into the cavity. In addition,
M4 was also highly transmitting (80%) at the SH of 399 nm. The radius of curvature of M3
and M4 was 200 mm, and the circulating beam inside the ring cavity was focused between
them. At the focus between M3 and M4, a 15 mm lithium triborate (LBO) crystal was chosen
as the SHG martial because of its large nonlinear coefficient and low walk-off effect. A small
leak through M3 is monitored by a calibrated photodiode (not shown) to measure the power
Pc circulating in the cavity; this photodiode is also used to align the cavity. To keep the
external doubler cavity enhancement of the laser pulses in a simple and stable system, we had
to control the external cavity round-trip length to the longitudinal modes of the Ti:sapphire
laser cavity. For this purpose, the mirror, M2, was mounted on a piezoelectric transducer
(PZT). The initial adjustment of the round-trip length was done with a translational stage. We
did this by monitoring the temporal resonance and the resultant enhancement of the
circulating beam while scanning the PZT-mounted mirror. Locking was performed by a
Hansch-Couillaud locking scheme, which was performed by a quarter-wave plate (QWP), a

#108967 - $15.00 USD Received 19 Mar 2009; revised 21 Apr 2009; accepted 23 Apr 2009; published 29 Apr 2009
(C) 2009 OSA 11 May 2009 / Vol. 17, No. 10 / OPTICS EXPRESS 8121
polarization beam splitter (PBS), and a pair of detectors D1, and D2. The error signals were
generated with a differential amplifier and feedback to a PZT driver through servo loop.
At the second stage, we generate the VUV laser light by single-pass configuration SHG, in
which an enhanced resonant cavity is not employed. The generated UV light from enhanced
SHG cavity, as shown in Fig. 1, is focused into a KBBF crystal through a lens. Output power
of the VUV light is measured by a power meter. KBBF crystal is an excellent VUV nonlinear
crystal. However, the KBBF crystal has a plate like form along the z-axis of the crystal and its
thickness is extremely limited around 2~3 mm. Hence, a special prism coupling technique
[14] is necessary to avoid cutting the crystal along the phase-matching direction for SHG. The
KBBF crystal is sandwiched between two fused-silica prisms, which have almost the same
refractive indices with KBBF in UV region, through optical contact. The whole device has
been named KBBF PCT device and they have been successfully used in the VUV laser
generation in recent years [8-11]. In our experiment, a PCT device, which consists of a KBBF
with thickness of 2.3 mm and a pair of fused-silica prisms with an angle of 55 degree, is
employed. Phase matching angle of KBBF in type-I SHG depends the fundamental
wavelength. To obtain the perfect phase matching angle, the incident angle of fundamental is
slightly adjusted from zero degree since the PCT device is designed for SHG performance
near 400 nm. In experiment, it was realized by mounting the PCT device on a rotating stage.
The incident angle can be adjusted by rotating the stage. Generated VUV and fundamental
UV wave can be separated though the rear prism based on the different refractive indices.
3. Experimental results
1.4

1.2
SHG power (W)

1.0

0.8

0.6

0.4

0.2

0.0

0 20 40 60 80 100
Reflectivity of input coupler (%)

Fig. 2. Second-harmonic output power as a function of the reflectivity of the input coupler
mirror. The experimental and theoretical values are shown by filled square and the solid curve.

In experiment, it is impossible to get a perfect mode matching for the injected fundamental
wave; we extended the standard equations to account for the mode-matching factor m
( 0 < m < 1 ) [15]. The SH output power ( P2 ), reflected fundamental power ( Pr ), and the
circulating fundamental power ( Pc ) inside the cavity are given by,
2
P2 = T2 E NL Pc ,
Pc 4Tm
= ,
P1 (T + L + E NL Pc ) 2
Pr (T − L − E NL Pc ) 2
= (1 − T )(1 − m) + m,
P1 (T + L + E NL Pc ) 2
where P1 is the fundamental power incident on the cavity, ENL is the single-pass conversion
efficiency, L is the round-trip intracavity loss and T is the transmission of the input coupler.
ENL can be measured by removing the input coupler of the enhanced cavity, with a result that
ENL = 0.12 ± 0.02 W −1 . We have tested three different cavities with different input coupler
(T=5%, 10%, and 30%). Using the above formulas, the intracavity loss and mode-matching

#108967 - $15.00 USD Received 19 Mar 2009; revised 21 Apr 2009; accepted 23 Apr 2009; published 29 Apr 2009
(C) 2009 OSA 11 May 2009 / Vol. 17, No. 10 / OPTICS EXPRESS 8122
factor can be calculated. It resulted in estimation of L = (1.3 ± 0.2)% and m = 0.70 ± 0.04 ,
respectively. Figure 2 gives the output SH average power as a function of the reflectivity of
the input coupler. The solid curve is the theoretical prediction using the above-mentioned
parameters. The generated UV with different input coupler is also plotted in the figure. There
is good agreement between the experiment and theory, resulting that the estimation of
intracavity loss and mode matching factor were reasonable.
Except for the above-mentioned parameters, ENL , L , and m , the performance of the
doubling cavity also depends essentially on the impedance matching and power of
fundamental wave. The impedance matching can be achieved when the input coupler
transmission equals the sum of all other losses, including the power-dependent conversion
losses, Topt = L + ENL Pc . For our estimated parameter values Topt is calculated to be 37% at the
fundamentals power of 1.7 W. Since the enhanced cavity is a low finesses cavity, the output
SH power is insensitive the transmission of the input coupler near the impedance matching
condition. It also be indicated in Fig. 2. The difference of output SHG power is less than 10%
when the transmission of input coupler belongs between 30% and 40%. In our following
experiment, an input coupler with transmission of 30% is employed.

Fig. 3. Second-harmonic output power P2 at 399 nm and conversion efficiency versus input
fundamental power P1 at 798 nm. The solid curve is the theoretical prediction based on the
independently measured single-pass conversion efficiency and estimated intracavity losses and
mode-matching factor. The experimental points labeled by filled circles are output power and
the points labeled by triangles are the conversion efficiency.

Figure 3 gives the input-output characteristics and conversion efficiency of the external
resonant doubler from 798 nm to 399 nm, also plotted in the figure are our experimental
results for P2 versus P1 when the external cavity was locked to an incident fundamental
resonance. From the measured values for ENL , and estimated values for m , and L , the
theoretical prediction for the harmonic output P2 and conversion efficiency are also shown in
Fig. 3 as the fundamental power P1 . More than 1.1 W of output at 399 nm was observed with
the maximum input of 1.7 W and a maximum efficiency was 65% . If the reflective loss
(20%) at the output mirror was accounted for, the power generated in the crystal was
estimated to be 1.37 W, corresponding to 80% conversion efficiency. It shows good
agreement between the experiment and theory. Hence it is reasonable to say that the
impedance matching was nearly optimized for this experiment. At present, the conversion
efficiency, we believe, was limited only by the mode-matching factor. With a perfect mode
matching ( m → 1 ) we would expect to achieve more than 90% conversion efficiency. The
output beam pattern at 399 nm was elliptic suffering from the walk-off effect in nonlinear
crystal. In the far field, the beam pattern has an aspect ratio of approximately 1:6. The

#108967 - $15.00 USD Received 19 Mar 2009; revised 21 Apr 2009; accepted 23 Apr 2009; published 29 Apr 2009
(C) 2009 OSA 11 May 2009 / Vol. 17, No. 10 / OPTICS EXPRESS 8123
compensation of walk-off and improvements of the generated beam pattern are in considering
and processing.
Figure 4(a) gives output power curves of VUV laser light, which was generated by single-
pass through SHG using KBBF as nonlinear crystal from 399 nm to 199.5 nm with a focused
lens of focused length 200 mm; and Fig. 4(b) shows the conversion efficiency. A maximum
output power of about 25 mW, corresponding to conversion efficiency of 2.2%, was obtained
at the input UV fundamental power of 1.1 W. We did here the same behavior: the curve in
Fig. 4(a) was well fitted by Pout = E NL 2 Pin 2 with an ENL 2 = 0.02 W −1 . Meanwhile, a linear
fitting the conversion efficiency gives an ENL 2 of 0.019 W −1 , which is agree well with the
above result, in Fig. 4(b). From Fig. 4(b), it also shows that the conversion efficiency is not
saturated yet, and higher conversion efficiency and output power could be obtained if more
fundamental power at 399 nm is applied. A direct conversion efficiency of 1.5% from
Ti:sapphire laser is obtained. The further increase of conversion efficiency and output power
could be realized by establishing a resonant enhanced cavity.
We also obtained VUV output at 193.5 nm from the fourth-harmonic generation of
Ti:sapphire laser. To do this, the wavelength of Ti:sapphire laser was tuned from 798 nm to
774 nm. The generated 387 nm UV laser light was as the fundamental input incident into the
KBBF crystal. Once again, the phase matching angle for SHG can be realized by slightly
adjusting the angle of incidence. Since the enhanced resonant cavity is designed for SHG from
798 nm to 399 nm, external losses of the resonant enhanced cavity at 774 nm became larger
and only 400 mW of output power at 387 nm was obtained. The VUV light at 193.5 nm with a
output power of 4.7 mW (corresponding to conversion efficiency of 1.1% ) was obtained with
fundamental input power of 400 mW at 387 nm.

Fig. 4. VUV output power (a) and conversion efficiency (b) at 199.5 nm by an SHG with a
KBBF crystal. The filled circles are experimental data and curve is best fitting with
2
Pout = E NL 2 Pin , giving E NL 2 = 0.02 W −1 . The triangles are the measured conversion efficiency
and solid line is linear fitting, give a E NL 2 = 0.019 W −1 .

4. Summary
In summary, we have demonstrated an all-solid-state VUV coherent light source by two
successive second harmonic generation from a Ti:sapphire laser system. With an enhanced
cavity, about 80% SHG conversion efficiency was achieved for quasi-continuous-wave laser
light. Maximum output power of 25 mW at 199.5 nm was also obtained by single-pass
through SHG with a KBBF as nonlinear crystal at fundamental power of 1.1 W.

#108967 - $15.00 USD Received 19 Mar 2009; revised 21 Apr 2009; accepted 23 Apr 2009; published 29 Apr 2009
(C) 2009 OSA 11 May 2009 / Vol. 17, No. 10 / OPTICS EXPRESS 8124

You might also like