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

Pulse shaping for mode-selective ultrafast coherent Raman spectroscopy of highly scattering solids

Open Access Open Access

Abstract

We report on mode-selective ultrafast coherent anti-Stokes Raman spectroscopy (CARS) of a powder of sodium dipicolinate. We produce a pair of stretched laser pulses with precisely adjusted matching chirp rates. We use the pulses for frequency-resolved excitation of coherent molecular vibrations in this highly scattering medium. The induced oscillations are probed with the third delayed ultrashort laser pulse. Since the attained spectral width of the pump-Stokes excitation band is on the order of the spacing between the Raman lines, time-resolved CARS measurements reveal single-mode as well as double-mode coherence decay dynamics, depending on the timing between the pump and Stokes pulses. For a fixed probe pulse delay, the sweeping of the arrival time for one of the preparation pulses maps out the CARS spectrum of the analyte.

© 2008 Optical Society of America

1. INTRODUCTION

Recently, a great deal of attention has been drawn to open-loop and adaptive pulse shaping techniques [1, 2, 3, 4]. The inherently large spectral bandwidth of ultrashort laser pulses has been utilized for various coherent control schemes, where the process outcome is manipulated through constructive and destructive interferences of possible excitation paths. The cancellation of the two-photon absorption [2], control of chemical reaction branching ratios [1], or vibrational wave packet dynamics [3, 4] are well-established examples of coherent control. Another important application of pulse shaping, relevant to the subject of this work, is the restoration of the spectral selectivity.

In coherent anti-Stokes Raman scattering (CARS) spectroscopy [5], two laser pulses, at carrier frequencies ω1 and ω2, initiate coherent molecular vibrations, and the third pulse (at carrier frequency ω3) scatters inelastically off them [see Fig. 1a ]. For this four-wave mixing (FWM) process, the specificity can be gained through the mode-selective excitation of Raman transitions, i.e., at the preparation stage, as well as through the shaping of the probe pulse. In a series of experiments, Silberberg's group demonstrated the use of periodic modulation [6] and steplike jumps [7, 8, 9] of the pulse phase for this purpose. An attractive alternative is to use linearly chirped pulses [10, 11, 12, 13, 14]. This simplest pulse shape is easy to produce and control. It also offers a very intuitive picture of how pulse shaping works.

We utilize a pair of broadband but linearly chirped laser pulses for selective excitation of adjacent Raman modes in a powder of sodium dipicolinate (NaDPA). Pulse chirp leads to the effective reduction of its instantaneous bandwidth. Provided that the preparation pulses have the same amount of chirp, as in Fig. 1b, their instantaneous frequency difference can be kept constant over the whole region where the pump and Stokes pulses are overlapped. The excitation bandwidth is then determined by the effective (instantaneous) spectral widths of the pump and Stokes pulses. The frequency difference depends on the time delay between the two pulses and can be conveniently adjusted.

Our approach is analogous to those described in [12, 13], but the probing of the induced molecular vibrations differs. We use a third, time-delayed ultrashort pulse to generate the background-free CARS signal. By varying the probe pulse delay, we can observe the coherence relaxation dynamics or even quantum beats if several Raman transitions are excited [15]. We can also directly map out the CARS signal intensity versus the Raman shift by scanning the timing between the shaped pump and Stokes pulses.

Note that the pulse configuration used here is conceptually opposite to the one employed in the hybrid CARS scheme (see [16]; known also as fs/ps CARS [17]), where broadband excitation of multiple Raman transitions via a pair of ultrashort transform-limited pulses is combined with their time-delayed but frequency-resolved probing. At last, the choice of NaDPA powder as an analyte is not arbitrary. NaDPA is an easy-to-make substitute for calcium dipicolinate (CaDPA), which in turn is a marker molecule for bacterial spores [18, 19]. To this end, the goal of this work [16, 20, 21, 22] is to develop a technique that would facilitate the interrogation of powderlike opaque solids and detection of harmful agents, such as B. anthracis spores, by means of CARS spectroscopy.

2. MATERIALS AND METHODS

We use a commercially available regenerative Ti:sapphire-based amplifier system (Legend, Coherent, Inc.), evenly pumping two optical parametric amplifiers (OPAs), to produce three synchronized laser pulses of different colors: pump (λ1=722.5nm, Δν1Δω1(2πc)287cm1), Stokes (λ2=804nm, Δν2=464cm1), and probe (λ3=579nm, Δν3=210cm1). Pulse shaping is done by sending the initially transform-limited preparation pulses through a 4cm slab of SF11 glass. Additionally, the Stokes pulse is guided through a commercially available pulse shaper (Silhouette, Coherent, Inc.). The pulse shaper is used for fine tuning of the Stokes pulse phase to match its chirp to the chirp of the pump pulse. After the computer-controlled delay stages, the three beams are focused onto a single spot on the surface of a sample, which is a rotated pellet of NaDPA powder.

The schematic layout of the setup and the pulse timing is given in Fig. 1c. The scattered CARS signal is collected in the backward direction by a 2in. (1in.=2.54cm) concave mirror, slightly offset from the main axis. The light is filtered from the dominant pump, Stokes, and probe photons and refocused on the entrance slit of an imaging spectrometer (Chromex-250is) with a liquid-nitrogen-cooled charge-coupled device (LN2-cooled CCD; Spec-10, Princeton Instruments). CARS spectrum is recorded as a function of the probe or pump pulse delay, depending on the experiment.

3. SHAPING OF THE PREPARATION PULSES

As was mentioned above, we shape the initially transform-limited pump and Stokes pulses by sending them through a slab of SF-11 glass. To the first approximation, the dispersion of the medium produces a linear chirp of the pulses. Note that since the beams are not focused inside the glass, pulse propagation is well described within the linear model, i.e., without taking into account possible pulse reshaping due to the field-induced nonlinearities.

For the sake of simplicity, we assume the input pulses to have a Gaussian temporal profile, i.e., E(t)=E0exp(Γ0t2+iω0t), where E0 is the complex amplitude of the electric field, ω0 is the laser pulse carrier frequency, and Γ0a0ib0 is a complex coefficient related to the full-width-at-half-maximum (FWHM) duration of the pulse and its FWHM spectral width. Explicitly,

τ=2ln(2)a0,Δω=22ln(2)a0(1+b02a02).
If the frequency-dependent wavenumber is approximated by the first three terms of its Taylor series expansion near ω0, i.e.,
k(ω)ωn(ω)ck(ω0)+k(ω0)(ωω0)+k(ω0)(ωω0)22,
the output laser field amplitude, after the pulse travels distance L through the dispersive medium with negligible absorption, is given by
E(t)=E0exp{iω0[tLVϕ(ω0)]}exp{Γ(L)[tLVg(ω0)]2},
where Vϕ(ω0)ω0k(ω0) is the phase velocity, and Vg(ω0)1k(ω0) is the pulse group velocity [23]. The pulse acquires a linear chirp with
1Γ(L)1aib=1Γ0+2ik(ω0)L.

From the measured spectrum of the pump pulse and the known dispersion properties of the glass, we estimate the input pulse duration, τ1in, to be 51fs, and the output, after 4cm of SF11 glass, τ1out=0.48ps. The last number agrees well with the measured FWHM of the cross-correlation profile between the stretched pump and ultrashort transform-limited probe pulses [2ω3ω1 process, see Fig. 2a ], which is 0.47ps. The calculated linear chirp, 2b, is equal to 0.11mradfs2 (0.58cm1fs1). This corresponds to dλ1dt31nmps and results in 15±2nmps slope for the recorded cross-correlation spectrogram, since dλFWMdt(λFWMλ1)2dλ1dt14nmps. The effective bandwidth of the pump pulse, Δν1eff=(τ1inτ1out)Δν130cm1, is reduced by almost an order of magnitude.

Because of slightly different dispersion of SF-11 glass at the Stokes wavelength and the different initial bandwidth of the Stokes pulse, its propagation through the same glass thickness does not result in the same chirp. Indeed, direct calculations give the linear chirp 0.13mradfs2 (0.70cm1fs1). To match the chirp of the two preparation pulses, we utilize the pulse shaper Silhouette, mentioned above. We use the multiphoton intrapulse interference phase scan (MIIPS) method [24, 25] to compensate for the phase distortions due to optical elements other than SF-11 glass (with the slab removed from the beam path), and then put the glass bar back and impose a parabolic phase on top of the found spectral compensation mask. This is equivalent to merely changing the length of the glass slab.

With a thin microscope cover slide placed in the focus of the overlapped beams, we monitor the spectrum of the FWM due to the pump, Stokes, and probe pulses (ω1ω2+ω3 process) as a function of the probe pulse delay to check for the center frequency variation in the FWM signal [see Fig. 2b]. Obviously, the absence of such variation is a direct indication that the two pulses have the same chirp. Note that the spectral bandwidth of the recorded signal is determined by the bandwidth of the probe pulse rather than of the preparation pulses. Assuming a perfect match of the pump and Stokes linear chirps, we estimate the output Stokes pulse duration as 0.79ps and the effective bandwidth Δν2eff18.5cm1. The resulting spectral bandwidth of the pump-Stokes convolution, responsible for Raman excitation, is therefore Δν12eff=[(Δν1eff)2+(Δν2eff)2]1235cm1, i.e., less than the frequency difference between the two Raman modes of NaDPA powder, 1395 and 1442cm1, shown in the inset of Fig. 1a.

4. SELECTIVE EXCITATION AND TIME-DELAYED PROBING

According to the spontaneous Raman spectrum of NaDPA [see the inset in Fig. 1a], the vibrational transitions of interest have a wavenumber difference of 47cm1, i.e., they are well within the bandwidth of the probe pulse. This makes it difficult to resolve the two Raman transitions by means of a straightforward CARS spectrum acquisition. Fortunately, one can still gain the required resolution through the time-resolved measurements, and there are two complementary ways to do it.

The first one is demonstrated in Fig. 3 . CARS spectra are recorded for fixed pump-Stokes timing as a function of the probe pulse delay, as it is typically done in femtosecond CARS measurements [26]. For such spectrograms, the response at zero probe delay is usually overwhelmed by the nonresonant (NR) contribution due to the instantaneous electronic response and off-resonant Raman modes. The Raman-resonant contribution results in a long-living but exponentially decaying signal at the positive probe delays. The excitation of multiple Raman modes, with frequency differences within the probe pulse spectral bandwidth, manifests itself through the observation of quantum beats [15], which are amplitude modulations of the recorded time-resolved CARS signal at the beat frequencies of Raman-mode pairs.

The set of spectrograms in Fig. 3 shows a gradual change from a single-mode to double-mode excitation, and then back to the single-mode dynamics but for the other Raman transition. When the effective pump-Stokes frequency difference is far detuned from any particular Raman mode, as in Figs. 3a, 3h, the spectrograms exhibit an asymmetric FWM profile due to solely the NR contribution. The asymmetry along the probe pulse delay is introduced by the multiple scattering of incoming photons in the powder, which favors negative probe delays (when the probe pulse is ahead of the pump-Stokes pair) over positive ones for the FWM to occur. Tuning the pump-Stokes frequency difference closer to one of the Raman modes, as in Figs. 3b, 3g, results in the appearance of an exponentially decaying CARS signal at positive probe delays. Finally, the double-mode excitation [see Figs. 3c, 3d, 3e, 3f] leads to the signal modulation as a function of the probe pulse timing. The variation in the contrast of the beating pattern is an indication of the selective excitation of the two Raman modes. Indeed, the contrast is expected to be maximal when the mode contributions into the CARS signal are equal. It deteriorates otherwise, and one observes purely exponential decay (without oscillations) if only a single mode is excited.

When the beating is observed, the difference frequency can be retrieved (see Fig. 4 ). In particular, we get 46±3cm1 from the fast-Fourier transform (FFT) of the recorded signal, corrected for its exponential decay. The decay times for 1395 and 1442cm1 transitions are found to be 0.7±0.1ps (Δν=7.6±1.1cm1) and 0.8±0.2ps (Δν=6.6±1.6cm1), respectively. They are in agreement with the linewidths estimated from the spontaneous Raman spectrum of NaDPA powder, 6.7 and 7.5cm1.

The second approach is to fix the probe pulse timing and record the CARS signal as a function of the pump or Stokes pulse delay. Obviously, it is advantageous to have the probe pulse delayed with respect to the preparation pulses in order to suppress the NR contribution into the generated signal. The resulting profiles of the spectrally integrated CARS signal, when the probe pulse timing corresponds to the quantum-beat peaks, are shown in Fig. 5 . CARS spectrum (with the spectral resolution defined by the spectral width of the convoluted pump-Stokes pair) is mapped out along the pump pulse delay. From the known pulse chirp, one can determine the frequency difference between the two Raman transitions. We estimate it to be 62±14cm1, which is somewhat higher than the value derived from spontaneous Raman measurements. The inset in the right top corner of Fig. 5 is an example of the original CARS spectrograms, before the integration over the spectrum. One can see a slight upward shift of the peak CARS wavelength when the 1442cm1 vibrational mode is superseded by the Raman transition at 1395cm1.

5. CONCLUSION

A combination of linearly chirped preparation pulses and time-delayed ultrafast probing conveniently brings together the excitation selectivity and background-free acquisition of the Raman-resonant CARS signal. The two time scales allow for the standard time-resolved measurements of the coherence decay as well as the frequency-resolved mapping of the CARS spectrum as a function of the relative pump-Stokes delay. The described approach is shown to work even for highly scattering solids, such as powders, when the input laser pulses and the generated CARS signal are subject to multiple scattering inside the sample. Finally, the linear chirp is the simplest pulse shape to produce and manipulate, which makes the technique attractive for practical applications. One possibility is to use broadband (chirped) laser pulses for adiabatic preparation of a highly coherent molecular supersposition state, which has previously been done with narrow-linewidth lasers slightly detuned from a Raman resonance [27, 28].

ACKNOWLEDGMENTS

We thank Marlan O. Scully, Yuri V. Rostovtsev, and George R. Welch for motivating our work, for stimulating discussions of the results, and for support of the project; we are grateful to Jaan Laane and his group for their help with spontaneous Raman measurements on NaDPA powder, and to Coherent Inc. for lending us commercially available pulse shaper Silhouette. The work is sponsored by the Defense Advanced Research Projects Agency, the National Science Foundation (grant PHY-0354897), an Award from Research Corporation, and the Robert A. Welch Foundation (grant A-1547).

 figure: Fig. 1

Fig. 1 Technique and experiment schematics: (a) CARS energy level diagram. The two broadband but shaped preparation pulses, pump (ω1) and Stokes (ω2), excite Raman-active vibrational modes of the sampled molecules. The third pulse (ω3) probes the initiated coherent molecular vibrations. Inset: spontaneous Raman spectrum of NaDPA powder in the range of interest, acquired with a CW laser operating at 532nm. (b) Time-frequency diagram of the selective Raman excitation with linearly chirped laser pulses. The difference frequency Δωω1ω2 depends on the relative timing τ12 between the preparation pulses. (c) Experimental setup layout. The pump and Stokes pulses are sent through 4cm pieces of SF-11 glass. The Stokes pulse also passes through a commercially available pulse shaper (Silhouette, Coherent), where a parabolic phase mask is added to compensate for the difference in the chirp, produced by the glass slabs. The pump and probe time delays are adjusted relative to the Stokes pulse. The three beams are focused (with 40 to 50cm focal length lenses) on the sample, a pellet of NaDPA powder. The generated and scattered CARS photons are collected with a 2in. spherical mirror (f=20cm) in the backward direction, at an angle of 30° to the main axis. The collected light is filtered and refocused on the entrance slit of an imaging spectrometer (Chromex-250is) with a LN2-cooled CCD; CCD - charge coupled device.

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 Pulse shaping characterization: (a) Cross-correlation spectrogram between the chirped pump and transform-limited probe pulses. The spectrum of FWM signal (2ω3ω1 process), generated on a cover glass slide, is recorded as a function of the probe pulse delay; (b) Cross-correlation spectrogram between the linearly chirped pump, Stokes, and ultrashort probe pulses. Again, the spectrum of the FWM signal (ω1ω2+ω3 process) from a cover glass slide is acquired as a function of the probe pulse delay.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Selective excitation of Raman modes in NaDPA powder, 1442cm1 and 1395cm1. The relative timing τ12t1t2 between the two linearly chirped preparation pulses, pump (λ1=722.5nm) and (λ2=804nm), is set as (a) 333, (b) 133, (c) 100, (d) 67, (e) 0, (f) +67, (g) +100, and (h) +300fs. The induced molecular vibrations are probed with an ultrashort pulse at λ3=579nm. CARS spectrum as a function of the probe pulse delay is recorded. Timing of the pump-Stokes pulses leads to consecutive excitation of a single Raman mode at 1442cm1; both Raman modes, as it can be inferred from the beating; a single Raman mode at 1395cm1. The pump, Stokes, and probe pulse energies are 2.8, 1.1, and 0.39μJ, respectively. The integration time is 0.2s per step.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Cross section of the spectrogram in Fig. 3e at λ=535nm. The beat frequency at positive probe delays corresponds to the frequency difference between the two excited Raman modes, 47cm1. Inset: FFT of the recorded modulation, corrected for the exponential decay.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Spectrally-integrated CARS signal as a function of the pump pulse timing, τ12. The probe pulse delay is set as (a) 1.4, (b) 2.2, and (c) 2.9ps, i.e., close to the peaks of the quantum beat profile in Fig. 4, when the two Raman modes are excited. The pump, Stokes, and probe pulse energies are 3.3, 1.2, and 0.37μJ, respectively. The integration time is 0.2s per step. Inset: CARS spectrogram recorded in case (b).

Download Full Size | PDF

1. A. Assion, T. Baumert, M. Bergt, T. Brixner, B. Kiefer, V. Seyfried, M. Strehle, and G. Gerber, “Control of chemical reactions by feedback-optimized phase-shaped femtosecond laser pulses,” Science 282, 919–922 (1998). [CrossRef]   [PubMed]  

2. D. Meshulach and Y. Silberberg, “Coherent quantum control of multiphoton transitions by shaped ultrashort optical pulses,” Phys. Rev. A 60, 1287–1292 (1999). [CrossRef]  

3. V. V. Lozovoy, B. I. Grimberg, E. J. Brown, I. Pastirk, and M. Dantus, “Femtosecond spectrally dispersed three-pulse four-wave mixing: the role of sequence and chirp in controlling intramolecular dynamics,” J. Raman Spectrosc. 31, 41–49 (2000). [CrossRef]  

4. R. A. Bartels, T. C. Weinacht, S. R. Leone, H. C. Kapteyn, and M. M. Murnane, “Nonresonant control of multimode molecular wave packets at room temperature,” Phys. Rev. Lett. 88, 033001 (2002). [CrossRef]   [PubMed]  

5. Y. R. Shen, The Principles of Nonlinear Optics, Wiley Classics Library ed. (Wiley, 2003).

6. N. Dudovich, D. Oron, and Y. Silberberg, “Single-pulse coherently controlled nonlinear Raman spectroscopy and microscopy,” Nature 418, 512–514 (2002). [CrossRef]   [PubMed]  

7. D. Oron, N. Dudovich, D. Yelin, and Y. Silberberg, “Quantum control of coherent anti-Stokes Raman processes,” Phys. Rev. A 65, 043408 (2002). [CrossRef]  

8. D. Oron, N. Dudovich, and Y. Silberberg, “Femtosecond phase-and-polarization control for background-free coherent anti-Stokes Raman spectroscopy,” Phys. Rev. Lett. 90, 213902 (2003). [CrossRef]   [PubMed]  

9. D. Oron, N. Dudovich, D. Yelin, and Y. Silberberg, “Narrow-band coherent anti-Stokes Raman signals from broad-band pulses,” Phys. Rev. Lett. 88, 063004 (2002). [CrossRef]   [PubMed]  

10. A. M. Zheltikov and A. N. Naumov, “High-resolution four-photon spectroscopy with chirped pulses,” Quantum Electron. 30, 606–610 (2000). [CrossRef]  

11. A. S. Weling, B. B. Hu, N. M. Froberg, and D. H. Auston, “Generation of tunable narrow-band Thz radiation from large-aperture photoconducting antennas,” Appl. Phys. Lett. 64, 137–139 (1994). [CrossRef]  

12. E. Gershgoren, R. A. Bartels, J. T. Fourkas, R. Tobey, M. M. Murnane, and H. C. Kapteyn, “Simplified setup for high-resolution spectroscopy that uses ultrashort pulses,” Opt. Lett. 28, 361–363 (2003). [CrossRef]   [PubMed]  

13. T. Hellerer, A. M. K. Enejder, and A. Zumbusch, “Spectral focusing: high spectral resolution spectroscopy with broad-bandwidth laser pulses,” Appl. Phys. Lett. 85, 25–27 (2004). [CrossRef]  

14. G. W. Jones, D. L. Marks, C. Vinegoni, and S. A. Boppart, “High-spectral-resolution coherent anti-Stokes Raman scattering with interferometrically detected broadband chirped pulses,” Opt. Lett. 31, 1543–1545 (2006). [CrossRef]   [PubMed]  

15. R. Leonhardt, W. Holzapfel, W. Zinth, and W. Kaiser, “Terahertz quantum beats in molecular liquids,” Chem. Phys. Lett. 133, 373–377 (1987). [CrossRef]  

16. D. Pestov, R. K. Murawski, G. O. Ariunbold, X. Wang, M. C. Zhi, A. V. Sokolov, V. A. Sautenkov, Y. V. Rostovtsev, A. Dogariu, Y. Huang, and M. O. Scully, “Optimizing the laser-pulse configuration for coherent Raman spectroscopy,” Science 316, 265–268 (2007). [CrossRef]   [PubMed]  

17. B. D. Prince, A. Chakraborty, B. M. Prince, and H. U. Stauffer, “Development of simultaneous frequency- and time-resolved coherent anti-Stokes Raman scattering for ultrafast detection of molecular Raman spectra,” J. Chem. Phys. 125, 044502 (2006). [CrossRef]  

18. J. F. Powell, “Isolation of dipicolinic acid (pyridine-2-6-dicarboxylic acid) from spores of bacillus-megatherium,” Biochem. J. 54, 210–211 (1953). [PubMed]  

19. B. D. Church and H. Halvorson, “Dependence of the heat resistance of bacterial endospores on their dipicolinic acid content,” Nature 183, 124–125 (1959). [CrossRef]   [PubMed]  

20. M. O. Scully, G. W. Kattawar, R. P. Lucht, T. Opatrny, H. Pilloff, A. Rebane, A. V. Sokolov, and M. S. Zubairy, “FAST CARS: engineering a laser spectroscopic technique for rapid identification of bacterial spores,” Proc. Natl. Acad. Sci. U.S.A. 99, 10994–11001 (2002). [CrossRef]   [PubMed]  

21. D. Pestov, M. C. Zhi, Z. E. Sariyanni, N. G. Kalugin, A. A. Kolomenskii, R. Murawski, G. G. Paulus, V. A. Sautenkov, H. Schuessler, A. V. Sokolov, G. R. Welch, Y. V. Rostovtsev, T. Siebert, D. A. Akimov, S. Graefe, W. Kiefer, and M. O. Scully, “Visible and UV coherent Raman spectroscopy of dipicolinic acid,” Proc. Natl. Acad. Sci. U.S.A. 102, 14976–14981 (2005). [CrossRef]   [PubMed]  

22. M. C. Zhi, D. Pestov, X. Wang, R. K. Murawski, Y. V. Rostovtsev, Z. E. Sariyanni, V. A. Sautenkov, N. G. Kalugin, and A. V. Sokolov, “Concentration dependence of femtosecond coherent anti-Stokes Raman scattering in the presence of strong absorption,” J. Opt. Soc. Am. B 24, 1181–1186 (2007). [CrossRef]  

23. A. E. Siegman, Lasers (University Science Books, 1986).

24. V. V. Lozovoy, I. Pastirk, and M. Dantus, “Multiphoton intrapulse interference. IV. Ultrashort laser pulse spectral phase characterization and compensation,” Opt. Lett. 29, 775–777 (2004). [CrossRef]   [PubMed]  

25. B. W. Xu, J. M. Gunn, J. M. Dela Cruz, V. V. Lozovoy, and M. Dantus, “Quantitative investigation of the multiphoton intrapulse interference phase scan method for simultaneous phase measurement and compensation of femtosecond laser pulses,” J. Opt. Soc. Am. B 23, 750–759 (2006). [CrossRef]  

26. A. Materny, T. Chen, M. Schmitt, T. Siebert, A. Vierheilig, V. Engel, and W. Kiefer, “Wave packet dynamics in different electronic states investigated by femtosecond time-resolved four-wave-mixing spectroscopy,” Appl. Phys. B 71, 299–317 (2000). [CrossRef]  

27. A. V. Sokolov and S. E. Harris, “Ultrashort pulse generation by molecular modulation,” J. Opt. B: Quantum Semiclassical Opt. 5, R1–R26 (2003). [CrossRef]  

28. A. V. Sokolov, M. Y. Shverdin, D. R. Walker, D. D. Yavuz, A. M. Burzo, G. Y. Yin, and S. E. Harris, “Generation and control of femtosecond pulses by molecular modulation,” J. Mod. Opt. 52, 285–304 (2005). [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 (5)

Fig. 1
Fig. 1 Technique and experiment schematics: (a) CARS energy level diagram. The two broadband but shaped preparation pulses, pump ( ω 1 ) and Stokes ( ω 2 ) , excite Raman-active vibrational modes of the sampled molecules. The third pulse ( ω 3 ) probes the initiated coherent molecular vibrations. Inset: spontaneous Raman spectrum of NaDPA powder in the range of interest, acquired with a CW laser operating at 532 nm . (b) Time-frequency diagram of the selective Raman excitation with linearly chirped laser pulses. The difference frequency Δ ω ω 1 ω 2 depends on the relative timing τ 12 between the preparation pulses. (c) Experimental setup layout. The pump and Stokes pulses are sent through 4 cm pieces of SF-11 glass. The Stokes pulse also passes through a commercially available pulse shaper (Silhouette, Coherent), where a parabolic phase mask is added to compensate for the difference in the chirp, produced by the glass slabs. The pump and probe time delays are adjusted relative to the Stokes pulse. The three beams are focused (with 40 to 50 cm focal length lenses) on the sample, a pellet of NaDPA powder. The generated and scattered CARS photons are collected with a 2 in. spherical mirror ( f = 20 cm ) in the backward direction, at an angle of 30 ° to the main axis. The collected light is filtered and refocused on the entrance slit of an imaging spectrometer (Chromex-250is) with a LN2-cooled CCD; CCD - charge coupled device.
Fig. 2
Fig. 2 Pulse shaping characterization: (a) Cross-correlation spectrogram between the chirped pump and transform-limited probe pulses. The spectrum of FWM signal ( 2 ω 3 ω 1 process), generated on a cover glass slide, is recorded as a function of the probe pulse delay; (b) Cross-correlation spectrogram between the linearly chirped pump, Stokes, and ultrashort probe pulses. Again, the spectrum of the FWM signal ( ω 1 ω 2 + ω 3 process) from a cover glass slide is acquired as a function of the probe pulse delay.
Fig. 3
Fig. 3 Selective excitation of Raman modes in NaDPA powder, 1442 cm 1 and 1395 cm 1 . The relative timing τ 12 t 1 t 2 between the two linearly chirped preparation pulses, pump ( λ 1 = 722.5 nm ) and ( λ 2 = 804 nm ) , is set as (a) 333 , (b) 133 , (c) 100 , (d) 67 , (e) 0, (f) + 67 , (g) + 100 , and (h) + 300 fs . The induced molecular vibrations are probed with an ultrashort pulse at λ 3 = 579 nm . CARS spectrum as a function of the probe pulse delay is recorded. Timing of the pump-Stokes pulses leads to consecutive excitation of a single Raman mode at 1442 cm 1 ; both Raman modes, as it can be inferred from the beating; a single Raman mode at 1395 cm 1 . The pump, Stokes, and probe pulse energies are 2.8, 1.1, and 0.39 μ J , respectively. The integration time is 0.2 s per step.
Fig. 4
Fig. 4 Cross section of the spectrogram in Fig. 3e at λ = 535 nm . The beat frequency at positive probe delays corresponds to the frequency difference between the two excited Raman modes, 47 cm 1 . Inset: FFT of the recorded modulation, corrected for the exponential decay.
Fig. 5
Fig. 5 Spectrally-integrated CARS signal as a function of the pump pulse timing, τ 12 . The probe pulse delay is set as (a) 1.4, (b) 2.2, and (c) 2.9 ps , i.e., close to the peaks of the quantum beat profile in Fig. 4, when the two Raman modes are excited. The pump, Stokes, and probe pulse energies are 3.3, 1.2, and 0.37 μ J , respectively. The integration time is 0.2 s per step. Inset: CARS spectrogram recorded in case (b).

Equations (4)

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

τ = 2 ln ( 2 ) a 0 , Δ ω = 2 2 ln ( 2 ) a 0 ( 1 + b 0 2 a 0 2 ) .
k ( ω ) ω n ( ω ) c k ( ω 0 ) + k ( ω 0 ) ( ω ω 0 ) + k ( ω 0 ) ( ω ω 0 ) 2 2 ,
E ( t ) = E 0 exp { i ω 0 [ t L V ϕ ( ω 0 ) ] } exp { Γ ( L ) [ t L V g ( ω 0 ) ] 2 } ,
1 Γ ( L ) 1 a i b = 1 Γ 0 + 2 i k ( ω 0 ) L .
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.