Sie sind auf Seite 1von 5

Journal of the Korean Physical Society, Vol. 52, No. 4, April 2008, pp.

10431047

A Sub-10-fs Ti:Sapphire Oscillator with a Simple Four-Mirror Cavity


Baatarchuluun Tsermaa, Byeong Kwan Yang, Jae Myung Seo and Jin Seung Kim
Institute of Photonics and Information Technology, Chonbuk National University, Jeonju 561-756 (Received 21 December 2007) Optical pulses of 9.4 fs are stably generated from a Kerr-lens mode-locked Ti:sapphire laser oscillator containing only four mirrors, including two double-chirped mirrors. To the best of our knowledge, the oscillator has the simplest cavity design compared with other sub-10-fs oscillators reported yet. The balanced distribution of the negative group delay dispersion within the cavity yields improved stability against intensity uctuations.
PACS numbers: 42.55.Rz, 42.60.By, 42.65.Re Keywords: Sub-10-fs, Ti:sapphire laser, Kerr-lens mode-locked, SHG-FROG

I. INTRODUCTION Recent technological advances in the development of highly doped, broadband gain medium and dispersionengineered chirped mirrors have led to better performances of Kerr-lens mode-locked (KLM) sub-10-fs laser oscillators [1, 2]. Both prism-controlled and mirrorcontrolled dispersion compensation methods are used for generating sub-10-fs pulses from KLM lasers [39]. The early designs mainly used prisms that provided ne tuning for the intracavity net dispersion. However, they have drawbacks including uncompensated higher-order dispersion components of the prism substrate, a cavity alignment sensitive stability, and a low repetition rate. The development of dispersion-engineered chirped multilayer dielectric mirrors [10, 11] has opened the possibility to overcome the limitations inherent in prismcontrolled dispersion compensation. The main advantages of mirror-dispersion-controlled (MDC) oscillators are simplicity, compactness, and stability of the cavity, which are essential for time-resolved experiments, carrier-envelope phase stabilization, and biomedical optical coherence tomography (OCT) applications [12,13]. One of the objectives of ultrashort pulsed laser research is the development of operationally eective, simple, and cost-eective laser oscillators. In order to further improve the current performance of ultrashort lasers, we have constructed a Ti:sapphire oscillator with a reduced component count that combines greater simplicity with improved performance. This paper reports generation of sub-10-fs optical pulses from a mirrordispersion-controlled KLM oscillator which employs only 4 mirrors: a pair of double-chirped mirrors, a totally reecting end mirror, and an output coupler. Even though four-mirror oscillators are not new [14], to the best of
E-mail:

our knowledge, this is the simplest sub-10-fs oscillator which uses only a pair of negative group delay dispersion (GDD)-oscillation-compensated mirrors. In order to tailor the required negative GDD in the sub-10-fs KLM laser cavity, more than two chirped mirrors are utilized, which increases the degrees of freedom for dispersion control [15, 16]. However, the increased number of chirped mirrors can either generate an asymmetrical distribution of dispersion that eventually leads to degraded KLM eciency [1719] or, at least, complicate the cavity geometry. Our four-mirror oscillator has fully solved the aforementioned weaknesses because it employs only two double-chirped mirrors symmetrically positioned in the two arms of the cavity. The arrangement of this paper is as follows: In Section II, the layout of the sub-10-fs Ti:sapphire oscillator is discussed. In Section III, pulse characteristics measured by using a second-harmonic-generation frequency-resolved optical gating (SHG-FROG) method are presented. In Section IV, the results are summarized.

II. EXPERIMENT The schematic of the constructed oscillator is shown in Figure 1. It is a Z-folded oscillator consisting of a Ti:sapphire crystal (X), a pair of GDD-oscillationcompensated curved double-chirped mirrors (DCMs) (M2 , M3 ), a highly reecting end mirror (M1 ) and an output coupler (OC). A highly doped (0.33 wt%), 2-mmthick Ti:sapphire crystal is oriented at Brewsters angle ( 60.4 ). M1 is a mirror with protected silver coating which has 98 % reectivity in the spectral range of 600 1200 nm and almost zero dispersion. The curved mirrors M2 and M3 (LAYERTEC GmbH) have 50-mm radii of curvatures, and the mirror pair produces the required negative dispersion inside the cavity. M2 and

bch@chonbuk.ac.kr

-1043-

-1044-

Journal of the Korean Physical Society, Vol. 52, No. 4, April 2008

Fig. 1. Schematic diagram of the mirror-dispersioncontrolled sub-10-fs oscillator (X, Brewster-oriented Ti:sapphire crystal; M1 , highly reecting end mirror; M2 and M3 , a pair of GDD-oscillation-compensated doublechirped mirrors; OC, output coupler; L, focusing lens for the pump laser; S, vertical slit; , folding angle).

M3 have a nominal GDD of 70 f s2 30 f s2 at 720 1020 nm and almost cancel out the GDD oscillations of each other over the bandwidth. Their reectivity exceeds 99.8 % in the spectral range of 630 980 nm, and almost transparent (R < 10 %) for a pump beam in the spectral range of 505 545 nm. We used a frequencydoubled, diode-pumped Nd:YVO4 laser (532 nm, 5 W) as a pump source. The pump beam is focused with a 40-mm focal-length lens (L) into the Ti:sapphire crystal. A water cooling system is used to remove the heat from the crystal. The output coupler is 1-mm thick, and it has a 5 % transmittance over the spectral range of 660 920 nm. The cavity has a round-trip length of about 2.1 m, which makes the repetition rate 142 MHz. The lengths of the cavity arms conned by M1 M2 and M3 OC have been set equal to d1 300 mm and d2 700 mm, respectively. We have done prior calculations using the ABCD matrix method [20, 21] in order to estimate the distance between the folding mirrors and the position of the gain medium to ensure stability of the resonator. Even though the oscillator can support both hard-aperture and soft-aperture KLM, we mainly used the hard-aperture approach by introducing a vertical slit (S) on the longer arm side of the cavity. The proper values of the folding angles ( = 2 = 3 15.22 ) at M2 and M3 were determined by requiring compensation of the astigmatism introduced by the tilted crystal. A more balanced distribution of the negative GDD in the two arms of the cavity allows the pulse to be short at the each end of the cavity and the center of gain medium [1]. A possible way for equal distribution of the negative GDD is use of the same negative GDD mirrors in the two arms of the cavity. When the spatial and the temporal foci of the intracavity pulse coincide in the crystal, the peak intensity may reach a high value in a narrow spatial range, which enhances the best self-focusing. Thus, a symmetrical dispersion distribution may lead to an effective saturable absorption that is twice as strong as an asymmetrical dispersion distribution, which results in substantially shorter pulses [1719]. The optimal spatial and temporal focusing yields improved stability against

uctuations of the intensity. As we used a highly doped, 2-mm-thick Ti:sapphire crystal, we had a chance to reduce the number of negative chirped mirrors of the cavity down to only a pair. Using only a pair of negative GDD mirrors as a dispersion compensator is ecient for not only simplicity but also better stability against perturbations. The latter can be attributed to the symmetrical distribution of the intra-cavity dispersion because it yields more ecient KLM. In order to achieve appropriate dispersion compensation, we estimated the dispersion characteristics of all the optical components in the cavity. The refractive indices of the Ti:sapphire gain medium and air were calculated according to Sellmeier equation [22] and results in estimates of the positive second- and third-order dispersions of the cavity. The nominal GDD of the contributing elements at the central wavelength of 800 nm are 260 f s2 for double passage through the gain medium, 40 f s2 for a 2.1-m optical path in air, 280 f s2 for 4 bounces on the negative GDD mirrors and 20 f s2 for a bounce on the output coupler [23]. The net GDD of the cavity takes a value near zero over the bandwidth of 720 1020 nm, which makes it suitable to enhance the KLM condition in the cavity. Instead of keeping the repetition rate constant, we changed the cavity length to ne tune the net intracavity dispersion and found the optimum length to satisfy the most favorable Kerr-lens mode-locking condition. After ne adjustments of the cavity mirrors and the crystal, pulses with suciently broad spectrum and an output power of 400 mW were easily generated by using a pumping power of about 4.5 W. The output retained good stability against environmental perturbations for hours in both the spectrum and the power.

III. PULSE CHARACTERIZATION For pulse characterization, we used the SHG-FROG method [24,25], together with the crystal angle-dithering technique [26]. The angle dithering was used to achieve phase matching over the whole spectrum of the pulse to be measured with a modest signal power. Figure 2 depicts the experimental setup for the pulse characterization. The pulse to be measured was split into two identical pulses by using a beam splitter, and one of them directed to a variable delay stage and then recombined with the other pulse by using another beam splitter to form a pair of pulses with some relative delay but moving parallel to each other. The pulse pair was focused by using an o-axis parabolic mirror (f = 75 mm) into the 100-m-thick BBO crystal, where their second harmonic is generated. The second harmonic signal was spectrally resolved by using a spectrometer (HR2000, Ocean Optics) and was recorded at every delay step of 1.05 fs. The movement of the delay arm mirrors was performed by using a piezo-actuator, which was attached to one of the

A Sub-10-fs Ti:Sapphire Oscillator with a Simple Baatarchuluun Tsermaa et al.

-1045-

Fig. 3. (a) Measured and (b) retrieved FROG traces.

4 1 0 .9 0 .8 0 .7 0 .6 0 .5 0 .4 0 .3 0 .2 0 .1 0 -4 0 -3 0 -2 0 -1 0 0 T im e , f s 10 20 30 40
4 1 0 .9 3

Temporal Intensity, arb. units

(a)

3 2

Fig. 2. Experimental layout of the SHG-FROG apparatus for characterization of sub-10-fs pulses.

Phase, rad

1 0 -1 -2

two arms of the Michelson interferometer. The interferometer provided accurate positioning of the delay arm mirrors. The phase-matching bandwidth of the 100-mthick BBO crystal was insucient for the entire bandwidth of the pulses, so the BBO crystal was mounted on a galvanometer to dither the crystal, which makes the effective phase-matching bandwidth could cover the whole bandwidth of the pulses. Transmission of the pulses through the beam splitter substrate induced substantial broadening of the pulses. To compress and recover their initial pulse shape, we estimated the GDD of the fused silica substrate of the beam splitter using Sellmeiers equation and achieved pre-compensation for this positive dispersion by using pairs of negative GDD mirrors. Figure 3 shows measured and retrieved SHG-FROG traces of 128 by 128 grid points. The symmetry of the measured FROG trace along the time delay axis indicates that the optical components of the FROG system were well aligned. The phase was retrieved by using the principal components generalized projection algorithm [27] for the iterative procedure. The FROG error of this reconstruction procedure amounts to 0.0081. Figure 4 shows the reconstructed pulse characteristics in the temporal and the spectral domains. As is seen in Figure 4(a), the temporal width of the pulse was 9.4 fs at FWHM. The atness of the temporal phase in the central peak suggests that the pulse is nearly transform limited. Figure 4(b) shows the spectrum and the spectral phase of the retrieved pulse. The retrieved spectrum covers a spectral range of 710 980 nm with a spectral width of 100.2 nm at FWHM. The small oscillation in the spectral phase may be attributed to the residual GDD oscillation induced by the DCMs and the output coupler. The pulse characteristics of sub-10-fs oscillators critically depend on the nature of the constituent optical components. Selection of cavity mirrors well-matched

Spectral Intensity, arb.units

0 .8 0 .7 0 .6 0 .5 0 .4 0 .3 0 .2 0 .1 0 700 750 800 850 900 W a v e le n g t h , n m

(b)

2 1 0 -1 -2 -3 -4 1000

Phase, rad.

950

Fig. 4. Reconstructed pulse proles. (a) Temporal intensity (solid) and phase (dashed) proles of a 9.4-fs pulse (FWHM). (b) Spectrum (solid) and spectral phase (dashed) proles (spectral width is 100.2 nm at FWHM).

with the spectral bandwidth and the dispersion characteristics of the intracavity optical components is the key issue for construction of an ideal oscillator with a pulse duration limited only by the bandwidth of the gain medium. However, commercially available components do not allow realization of such an ideal oscillator. In order to check the consistency of the retrieved pulse parameters, we compared the retrieved results with the independently measured data. Figure 5 depicts a comparison of the retrieved and the measured pulse spectra. The retrieved spectrum fairly well reproduces its measured counterpart over the whole bandwidth. For the time-domain consistency check, we measured the intensity autocorrelation of the pulse and compared it with the intensity autocorrelation calculated from the retrieved spectrum and the spectral phase of the pulse (Figure 6).

-10461 .0 0 .9 0 .8

Journal of the Korean Physical Society, Vol. 52, No. 4, April 2008

pendent measurements. The agreements between them were fairly good.

Intensity, a.u.

0 .7 0 .6 0 .5 0 .4 0 .3 0 .2 0 .1 0 .0 7 00

ACKNOWLEDGMENTS This work is supported in part by a Korea Research Foundation grant (KRF-2006-005-J00302) and by the Nano-Bioelectronics and Systems Research Center (Seoul National University), an Engineering Research Center supported by the Korea Science and Engineering Foundation.

7 50

800

8 50

90 0

9 50

1 00 0

W avelength, nm

Fig. 5. Comparison of the retrieved (thick line) and the independently measured (thin line) fundamental pulse spectra.
1 0 .9

REFERENCES
Intensity, arb. units
0 .8 0 .7 0 .6 0 .5 0 .4 0 .3 0 .2 0 .1 0 -8 0 -6 0 -4 0 -2 0 0 D e la y , f s 20 40 60 80

Fig. 6. Comparison of the independently measured intensity autocorrelation (dashed) with the autocorrelation (solid) calculated from the reconstructed spectrum and phase.

The agreement was fairly good, except for slight discrepancies at the leading and the trailing parts, which are sensitive to residual error.

IV. SUMMARY We have designed and constructed a very simple sub10-fs Kerr-lens mode-locked, Ti:sapphire laser oscillator by using only four mirrors: a pair of DCMs for intracavity dispersion compensation, one high reector, and one output coupler. The average output power was about 400 mW for a pumping power of 4.5 W. The balanced distribution of the negative group delay dispersion within the cavity yielded good stability against environmental uctuations, and the oscillator kept working for several hours without any problem. The pulse characteristics were measured by using the SHG-FROG method combined with crystal dithering technique. The intensity and the phase of the pulses in the temporal and the spectral domains retrieved from the measured FROG traces showed that the pulse width was 9.4 fs and the spectral bandwidth was 100.2 nm. Consistency checks of the phase retrieval were also performed in the temporal and the spectral domain with other inde-

[1] F. X. Kartner, Few-cycle Laser Pulse Generation and Its Applications (Springer, Berlin, 2004). [2] J. Ye and S. T. Cundi, Femtosecond Optical Frequency Comb Technology (Springer, New-York, 2005), Chap. 23. [3] L. Xu, G. Tempia, A. Poppe, M. Lenzner, Ch. Spielman, F. Krausz, A. Stingl and K. Ferencz, Appl. Phys. B 65, 151 (1997). [4] G. Steinmeyer, D. H. Sutter, L. Gallmann, N. Matuschek and U. Keller, Science 286, 1507 (1999) [5] E. Slobodchikov, S. Sakabe, T. Kuge and S. Kawato, Opt. Rev. 6, 149 (1999). [6] A. Baltuska, Z. Wei, M. S. Pshenichnikov, D. A. Wiersma and R. Szipocs, Appl. Phys. B 65, 175 (1997). [7] U. Morgner, F. X. Kartner, S. H. Cho, Y. Chen, H. A. Haus, J. G. Fujimoto, E. P. Ippen, V. Scheuer, G. Angelow and T. Tschudi, Opt. Lett. 24, 411 (1999). [8] Y. S. Lim, H. S. Jeon, Y. C. Noh, K. J. Yee, D. S. Kim, J. H. Lee, J. S. Chang and J. D. Park, J. Korean Phys. Soc. 40, 837 (2002). [9] Y. W. Lee and J. H. Yi, J. Korean Phys. Soc. 49, 412 (2006). [10] R. Szipocs, K. Ferencz, C. Spielmann and F. Krausz, Opt. Lett. 19, 201 (1994). [11] N. Matuschek, F. X. Kartner and U. Keller, IEEE J. Quantum Elect. 35, 129 (1999). [12] S. T. Cundi, J. Korean Phys. Soc. 46, 1181 (2006). [13] T. Brabec and F. Krausz, Rev. Mod. Phys. 72, 545, (2000). [14] A. Penzkofer, M. Wittmann, M. Lorenz, E. Siegert and S. Macnamara, Opt. Quantum Elect. 28, 423 (1996). [15] A. Stingl, M. Lenzner, Ch. Spielmann and F. Krausz, Opt. Lett. 20, 602 (1995). [16] T. Fuji, A. Unterhuber, V. S. Yakovlev, G. Tempia, A. Stingl, F. Krausz and W. Drexler, Appl. Phys. B 77, 125 (2003). [17] P. Christov and V. D. Stoev, Opt. Lett. 20, 2111 (1995). [18] I. P. Christov, V. Stoev, M. Murnane and H. Kapteyn, Opt. Lett. 21, 1493 (1996). [19] Y. H. Cha, K. Lee, J. M. Han and Y. J. Rhee, J. Opt. Soc. Am. B 20, 1369 (2003). [20] Y. W. Lee, J. H. Yi, Y. H. Cha, Y. J. Rhee, B. C. Lee and B. D. Yoo, J. Korean Phys. Soc. 46, 1131 (2005). [21] J. Xia and M. H. Lee, Appl. Opt. 41, 453 (2002).

A Sub-10-fs Ti:Sapphire Oscillator with a Simple Baatarchuluun Tsermaa et al. [22] M. Bass, Handbook of Optics Vol. II, 2nd ed. (McGraw Hill, New-York, 1995). [23] B. Tsermaa, Ph.D dissertation, Chonbuk National University, 2007. [24] R. Trebino, Frequency-resolved Optical Gating (Kluwer, Boston, 2002). [25] B. K. Yang and J. S. Kim, Hankook Kwanghak Hoeji

-1047-

15, 278 (2004). [26] P. OShea, M. Kimmel and R. Trebino, Opt. Expr. 7, 342 (2000). [27] D. J. Kane, IEEE J. Sel. Top. Quantum Elect. 4, 278 (1998).

Das könnte Ihnen auch gefallen