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

Supercontinuum-based three-color three-pulse time-resolved coherent anti-Stokes Raman scattering

Open Access Open Access

Abstract

We demonstrate the use of a photonic crystal fiber (PCF) as a compact three-color fs laser system operating at 76 MHz, limited only by the repetition rate of the pump laser. The system is suitable for background-free time-resolved four-wave mixing measurements, which arguably reach fundamental limits in signal detectivity. We give a detailed characterization of the near transform-limited multi-color pulses that are extracted from the PCF, and prove the system through time-resolved coherent anti-Stokes Raman scattering measurements in bipyridyl ethylene and styrene.

© 2015 Optical Society of America

1. Introduction

Parametric four-wave mixing (4WM) processes represent one of the more general families of ultrafast time-resolved nonlinear optical spectroscopies [1], which acquire their full capability when executed with three incident pulses in three colors that can be independently controlled [2]. Beyond flexibility, 4WM in three colors achieves ultimate sensitivity by allowing background free detection. In this limit, detectivity is determined by the noise equivalent count rate, R, of signal photons, which, for spontaneous scattering on a single center, is given quite generally by the product,

R=(IL/ω)σηfδt,
of incident local intensity, IL, scattering cross-section, σ, collection efficiency, η, and fill factor, ff=fδt, product of pulse width, δt, and laser repetition rate, f. There are fundamental limitations on the attainable count rate. To remain in the weak-field limit of perturbative parametric mixing processes, the Keldysh criterion [3] limits the maximum tolerable local field intensity to IL,max~1012 W/cm2 (EL,max~3 V/nm). This constitutes the field ionization driven damage threshold in condensed matter, which is easily attained in tightly focused fs laser pulses (1 nJ, 100 fs pulse, focused to a spot of 1 μm2). σ is a material property, therefore under optimized collection conditions (η~0.1), for ultrafast measurements with time resolution set by δt, it is necessary to operate at as high a repetition rate as allowed by the recovery time of the system under interrogation. For molecular matter, photon anti-bunching sets the ultimate limit of R~25 MHz [4], which applies in the limit where the scattering probability is saturated. Otherwise, thermal relaxation rates determine the upper limit in ff. We will conclude below that optimal source requirements for measurements on single molecules is fairly narrowly defined: three independent colors at f100 MHz and average power of < 1 mW (δt=100 fs). Here, we show an efficient and compact system based on photonic crystal fibers (PCF) that meet these requirements, and we demonstrate this through three-color time-resolved coherent anti-Stokes Raman scattering (tr-CARS) measurements.

Various approaches have been previously used for non-resonant, background-free CARS spectro-microscopy [5–30]. These include second harmonic based [7] or supercontinuum (SC) based techniques [8–21], or reliance on a single broadband pulse [22–28], along with approaches requiring kHz amplifiers [29,30]. In [29], a multiplexed three-color three-pulse tr-CARS scheme was developed to acquire non-resonant background-free spectra with high spectral resolution and relatively broad spectral range (~900 cm−1). However, this approach requires high intensity laser systems, including a kHz amplifier and optical parametric amplifier. The more efficient approach to achieve a large Raman frequency range (> 2000 cm−1) from an unamplified and high repetition-rate source (> 100 MHz) is to use SC generated in nonlinear fibers [8–21]. This has been implemented in frequency domain two-color CARS, with degenerate pump and probe excitations driven by the narrowband source and the Stokes induced by the continuum [8]. Microscopy with two-pulse three-color CARS, in which two different frequency components of the continuum act as pump and Stokes and the narrowband pulse acts as probe [9], has also been effectively implemented to obtain images tagged by vibrational dephasing times [10]. This approach takes advantage of the time-delayed probe to separate the resonant signal from the nonresonant background by virtue of the contrast in coherence times of the nonresonant background (< 10 fs) and vibrational resonances (0.5 ps-10 ps). Approaches based on amplified laser systems to pump multiple optical parametric amplifiers to generate multiple fs pulses are limited in repetition rate, typically to < 10 MHz. In contrast, oscillator pumped fiber-based setups have the advantage of higher repetition rate (> 100 MHz) at significantly reduced cost and complexity, signal-to-noise ratio improvement by higher data acquisition rates, and peak powers reduced by the fill factor to prevent sample damage. Despite these promising advantages, reports of tdomain 4WM via SC generation in fibers have been limited to impulsive application of the pump and dump (Stokes) pulses. Here, we demonstrate three-color, three-pulse, tr-CARS based on SC generation in a commercially available PCF and subsequent slicing of the spectrum with filters to generate the desired pump, Stokes, and probe pulses with independently controllable time-frequency profiles. The demonstration suggests that the approach is well suited to address the challenges of single-molecule ultrafast coherent spectroscopy [31–33].

2. Experiment

2.1. Methods

A detailed schematic of the three-color, three-pulse tr-CARS setup is shown in Fig. 1. A mode-locked Ti:sapph laser (Coherent Mira Seed) operating at 798 nm (average power = 700 mW, f=76 MHz, δt=40 fs) is used to pump a 12 cm long, polarization maintaining PCF (NKT Photonics NL-PM-750). The fiber is mounted in a protective module (FemtoWHITE 800), and housed in a commercial intensity control/fiber-coupling unit (Newport Wavelength Extender (WE) SCG-02). The WE splits the pump beam using a polarizing cube (PCBS) into two, and provides variable attenuators consisting of half waveplates and Glan-laser polarizers to control the intensities in the two arms. The first arm passes through a Faraday isolator (FI) and is used to pump the PCF with 50% coupling efficiency. The second arm, which consists of the remaining Ti:sapph output, is utilized as either the reference pulse in cross-correlation measurements or as the Stokes pulse in the CARS measurements (see below). The spectrally broad output of the PCF (inset of Fig. 1) is further split into red and blue arms using a 660 nm long wave pass dichroic mirror (Semrock FF660-Di02), followed by bandpass filters in each arm to select the desired colors. In the present, 10 nm FWHM bandpass filters centered at 610 nm (Thorlabs FB610-10) and 710 nm (Thorlabs FB710-10) are used in the blue and red arms, respectively. The sliced pulses are then recombined along with the leftover fundamental in a collinear geometry using a 757 nm long wave pass dichroic mirror (Semrock FF757-Di01). Separate delay lines and attenuators are used to introduce relative delay between the pulses and for intensity control. The collinear CARS measurements are carried out on liquid trans-1,2-bis-(4-pyridyl) ethylene (BPE) dissolved in methanol, or in pure styrene, contained in 1 mm quartz cuvettes. A 0.25 NA objective lens is used to focus the three beams to a spot size of ~10 μm. In our tr-CARS configuration, the sliced pulses at 708 nm and 612 nm serve as pump (Pu) and probe (Pr), while the leftover fundamental at 798 nm serves as the Stokes (St), to generate the anti-Stokes signal ωPuωSt+ωPr=ωAS centered at 557 nm. The transmitted excitation pulses, along with the CARS signal, are collimated and spectrally filtered (Semrock FF01-554/23-25) to isolate the anti-Stokes signal. The signal is collected using an avalanche photodiode (APD Excelitas SPCM-AQRH-16-FC), and recorded using a dual-channel photon counter (Stanford Research Systems SR400). The pulses are characterized through sum-frequency generation (SFG) cross-correlation using a 1 mm BBO crystal placed in the sample position. The three-pulse instrument response is characterized through third-order cross-correlation frequency-resolved optical gating (XFROG) [34] by replacing the sample with pure carbon disulfide (CS2) and using a spectrograph (Andor iDus 401/Shamrock 303i) in place of the APD. The pulses are not precompensated for group-delay dispersion (GDD) of the objective lens, which is estimated to be 2200 fs2 at 612 nm (4 cm, fused silica).

 figure: Fig. 1

Fig. 1 Experimental apparatus for three-color three-pulse CARS (λ/2 – half waveplate, PCBS – polarizing cube beamsplitter, FI – Faraday isolator, AT1, AT2 – attenuator assemblies consisting of half waveplates and Glan-laser polarizers, L1, L2 – aspheric lenses for focusing and collimation, PCF – photonic crystal fiber, WE – wavelength extender, DM1, DM2 – 660 nm and 757 nm long wave pass dichroic beamsplitters, BPF1, BPF2, BPF3, BPF4 – bandpass filters centered at 800 nm, 710 nm, 610 nm, and 554 nm respectively, SPF – 390 nm short wave pass filter, APD – avalanche photodiode, Spec – spectrograph, PMT – photomultiplier tube. Inset shows the SC spectrum on log scale.

Download Full Size | PDF

2.2. Characterization of the sliced PCF pulses

SC generation in the NL-PM-750 PCF, and applications therein, are available in prior reports [35–37]. The PCF displays anomalous dispersion with two zero crossings at 750 nm and 1260 nm. The output characteristics of the highly nonlinear fiber depend on the intensity, wavelength, and bandwidth of the pump beam. When directly pumped with the 40 fs (25 nm FWHM) Ti:sapph laser at 0.5 W average power, a fairly flat SC is generated across the wavelength range 400-900 nm, as illustrated in the inset of Fig. 1. Note that the red edge of the displayed spectrum is limited by the spectral response of our spectrometer, otherwise it stretches out to 1600 nm [35,36]. Under this pumping condition, the sliced pulses of average power ~5 mW stretch out to ~1 ps. It is in principle possible to compress these output pulses. Instead, we generate nearly transform-limited output pulses directly by limiting both bandwidth and intensity of the pump laser to counter-balance the interplay of self-phase modulation and dispersion. The generation of transform-limited SC components at these low intensities is thought to principally arise from soliton fission [13,38]. With the bandwidth of the pump laser limited to 10 nm using a bandpass filter (Fig. 1, BPF1), by adjusting the pump intensity, near transform-limited 100 fs pulses can be extracted from the PCF. Under these conditions, the SC is strongly colored, with spectral structure that tunes as a function of pump power [13,38]. This is illustrated in Fig. 2(a). For the set of spectral slices pre-specified by the bandpass filters to λ(δλ)=612(5) nm, 708(8) nm, and 798(10) nm, the shortest pulses are obtained when the PCF is pumped at 36 mW (total SC output of 18 mW). The strongly modulated SC spectrum obtained under this pumping condition is the green trace in Fig. 2(a), and the normalized spectra of the filtered slices are shown in Fig. 2(b). The filter functions inscribe the bands in the SC spectrum, to slice out sections of stationary phase, albeit with different intensities. The extracted powers at 612 nm and 708 nm are 110 μW and 25 μW, respectively. With RMS pulse-to-pulse intensity fluctuations of the pump laser of 1.4%, the RMS noise in the generated pulses at 612 nm and 708 nm is 23% and 10%, respectively.

 figure: Fig. 2

Fig. 2 (a) SC spectra with different output powers, (b) normalized spectra of the probe (612 nm), pump (708 nm), and Stokes (798 nm) pulses sliced from the SC spectrum given by the green trace in (a).

Download Full Size | PDF

To characterize the sliced pulses, we carry out cross-correlation and XFROG measurements, which are summarized in Figs. 3 and 4. Using the filtered 798 nm (94 fs) pulse as reference, the cross-correlations of the pump (708 nm) and probe (612 nm) pulses yield widths of FWHM = 150 fs and 153 fs, respectively (Fig. 3). Assuming Gaussian profiles, the cross-correlation widths δti,j=[δti2+δtj2]1/2 imply pulse widths of FWHM = 120 fs for both pump and probe slices. This is to be compared with their respective bandwidth-equivalent transform limit of 92 fs and 110 fs. The residual pulse broadening can be entirely accounted by GDD in the objective lens. Thus, for a Gaussian pulse traveling a distance z in a dispersive medium: τ/τ0=[1+(z/LD)2]1/2, where LD=τ02/β2 is the dispersive length and β2 is the second-order dispersion coefficient [39]. The observed broadening suggests that the pulses are subject to GDDβ2z=τ0(τ2τ02)1/2~2400 fs2 and ~2000 fs2 at 612 nm and 708 nm, which is close to the value of GDD=2200 fs2 and 1800 fs2 that we estimate for the objective lens at the same wavelengths (z=4 cm, fused silica). In effect, the sliced pulses out of the PCF appear to be transform-limited, with pulse widths limited by the pump laser.

 figure: Fig. 3

Fig. 3 Cross-correlations of pump (black dotted) and probe (open orange circles) pulses with the 798 nm reference pulse.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 (a) Measured XFROG trace, (b) retrieved temporal (black solid) and phase profiles (open circles), together with a quadratic fit (green solid).

Download Full Size | PDF

The third-order XFROG measurement, which is shown in Fig. 4, captures the instrument response function for 4WM measurements and confirms that the residual chirp in the pulses due to GDD in the objective is linear (quadratic phase). The measurement carried out in liquid CS2 consists of recording the anti-Stokes spectrum as a function of delay of the 610 nm probe pulse, and with coincident pump and Stokes pulses that are fixed in time. In effect, this is a spectrally-resolved tr-CARS measurement in the absence of vibrational resonances; therefore, it measures the instantaneous response of the system. The XFROG trace is shown in Fig. 4(a), and the retrieved temporal intensity (black solid) and phase (open circles) profiles using the Femtosoft XFROG software (version 3.2.4), are shown in Fig. 4(b). The temporal phase is predominantly quadratic, as shown by the fit (in green). The fit yields GDD=4300 fs2, in excellent agreement with the additive GDD values (2400 fs2 + 2000 fs2) extracted above from the cross-correlation measurements. Thus, near transform-limited 100 fs pulses, in three colors, at an average power of ~100 μW in each sliced pulse, are obtained out of the PCF at 76 MHz repetition rate.

3. CARS measurements and discussion

The time-frequency profiles of the filtered pulses were designed to interrogate quantum beats among ubiquitous, aromatic C=C stretching modes of organic molecules, which occur near 1600 cm−1. These are the prominent lines seen in the Raman spectra of BPE and styrene, which are shown in Fig. 5, along with the measured tr-CARS time traces. Note that both Raman and CARS are determined by the third order polarization, P(3). The first measures the imaginary part, Im[P(3)(ω)], in the frequency domain, while the latter measures |P(3)(t)|2 in the time domain [1]. As such, for an ensemble measurement with homogeneously broadened lines, the Raman spectrum contains all of the information necessary to predict the tr-CARS response. With the 708 nm pulse acting as pump and the 798 nm pulse acting as Stokes, the polarization that can be prepared is delimited by the Raman window given by the spectral convolution:

WR(ω)=IPu(ω')ISt(ω'ω)dω'.
The window, which is centered at ωPuωSt=1593 cm−1 and has a FWHM of 224 cm−1, is superimposed on the Raman spectra (insets in Fig. 5). The tr-CARS traces are retrieved from the inverse Fourier transform of the window-weighted Raman spectrum:
SAS(t)=|WR(ω)IR(ω)eiωtdω|2.
The comparison of the observed tr-CARS signal and its simulation for BPE and styrene are presented in Figs. 5(a) and 5(b). Beside the main beats, which occur with a period of 1 ps and 1.3 ps in BPE and styrene, respectively, the styrene time profile in Fig. 5(b) exhibits a small-amplitude higher-frequency modulation due to the weaker vibrations that appear in the Raman window. Formally, it is the third-order polarization that is detected through the anti-Stokes bandpass filter (BPF4 in Fig. 1), and the convolution with probe pulse can in principle diminish the depth of modulation in the time trace [40]. Simulations of the effect of probe convolution show that this is a minor effect, consistent with the instrumental response captured by the third order XFROG in Fig. 4.

 figure: Fig. 5

Fig. 5 Measured (black) and simulated (red) tr-CARS traces for (a) BPE and (b) styrene. The simulations are based on the windowed inverse Fourier transform of the Raman spectra (Eq. (3) of text). The Raman spectra and window function (blue) are shown in the insets.

Download Full Size | PDF

A principle advantage of the three-color 4WM was advanced as a background-free detection method, yet the time trace of the signal in Fig. 5(a) clearly rides over a background of 104 counts, comparable to the signal. This we have verified to arise from a two-photon induced fluorescence in BPE, which coincides with the anti-Stokes window. The baseline of 400 counts in the case of styrene [Fig. 5(b)] is due to scattered stray light, which sets the noise equivalent count rate in the setup. Otherwise, the dynamic range of the signal, S:N~100, agrees with the shot noise, N/N1/2, where N~104 is the signal photon count rate. The effect of source noise, due to pulse-to-pulse fluctuations of the nonlinear SC generation in the PCF, is not apparent.

4. Conclusion

Based on fundamental considerations, namely Eq. (1), we have argued that ultimate limits in detectivity of ultrafast time-resolved processes in molecular matter is reached through three-color fs lasers operating near or above 100 MHz. We have shown that this can be accomplished in a compact setup, through the use of a PCF, and we have proven the concept by implementing it to carry out tr-CARS measurements. The key finding is that nearly transform-limited 100-fs coherent multicolor pulses can be extracted from a single PCF by controlling the bandwidth and intensity of the pump pulse. The approach is amenable to a variety of adaptations and improvements. The repetition rate is only limited by the pump source, and, given the low requirement on pump power, the repetition rate of the system can be readily increased to 10 GHz using demonstrated passively mode-locked Ti:sapph oscillator technology [41]. And while we demonstrated the operation with prescribed fixed colors and pulse widths, it should be clear that through the use of variable bandpass filters, the entire PCF spectrum (400-1600 nm) is accessible to generate multiple compressible pulses of different color. To this end, the combination of the PCF and a pulse shaper [42] would greatly add to flexibility and compactness.

The development we presented is part of a program aimed at preparing and manipulating quantum coherences on individual molecules. Through the use of Eq. (1), it is easy to show that such measurements are impractical on bare molecules: R~1 Hz, ff=10−5, for a molecule with a very large Raman cross-section of 3×10−25 cm2. To reach single-molecule sensitivity, one takes advantage of the surface-enhanced Raman scattering (SERS) effect [32,33] through the use of plasmonic nano-antennas [43]. The standard SERS enhancement factor, EF=β4, which results from the enhanced local field β=EL/E0, modifies Eq. (1):

R=(I0β2/ω)(β2σ)ηfδt.
The distributed enhancement factors underscore the two effects in enhanced detectivity. The first is a reduction on the demand of the incident intensity, IL=I0β2, and the second is the enhanced effective cross section, σ*=β2σ. For a typical value of β=100, the demand on incident intensity is reduced by 4 orders of magnitude, while the detectivity is increased by 4 orders of magnitude. To be clear, the damage threshold of IL=1012 W/cm2 is now reached at pulse energies of 10−13 J, or 10 μW average power (ff=10−5), which is amply provided by the described compact system.

Acknowledgments

The authors thank S. Yampolsky, D. A. Fishman and E. O. Potma for useful discussions. This work was made possible by the National Science Foundation Center for Chemical Innovation on Chemistry at the Space-Time Limit (grant CHE-1414466).

References and links

1. S. Mukamel, Principles of Nonlinear Optical Spectroscopy (Oxford University, 1995).

2. D. Segale and V. A. Apkarian, “Dissipative quantum coherent dynamics probed in phase-space: electronically resonant 5-color 4-wave mixing on I2(B) in solid Kr,” J. Chem. Phys. 135(2), 024203 (2011). [CrossRef]   [PubMed]  

3. L. V. Keldysh, “Ionization in the field of a strong electromagnetic wave,” Sov. Phys. JETP 5, 1307–1314 (1965).

4. Th. Basché, W. E. Moerner, M. Orrit, and H. Talon, “Photon antibunching in the fluorescence of a single dye molecule trapped in a solid,” Phys. Rev. Lett. 69, 1516–1519 (1992).

B. Lounis, H. A. Bechtel, D. Gerion, P. Alivisatos, and W. E. Moerner, “Photon antibunching in single CdSe/ZnS quantum dot fluorescence,” Chem. Phys. Lett. 329, 399–404 (2000).

5. J.-X. Cheng and X. S. Xie, “Coherent anti-Stokes Raman scattering microscopy: instrumentation, theory, and applications,” J. Phys. Chem. B 108(3), 827–840 (2004). [CrossRef]  

6. W. Min, C. W. Freudiger, S. Lu, and X. S. Xie, “Coherent nonlinear optical imaging: beyond fluorescence microscopy,” Annu. Rev. Phys. Chem. 62(1), 507–530 (2011). [CrossRef]   [PubMed]  

7. J. Peng, D. Pestov, M. O. Scully, and A. V. Sokolov, “Simple setup for hybrid coherent Raman microspectroscopy,” J. Raman Spectrosc. 40(7), 795–799 (2009). [CrossRef]  

8. Y. J. Lee, S. H. Parekh, Y. H. Kim, and M. T. Cicerone, “Optimized continuum from a photonic crystal fiber for broadband time-resolved coherent anti-Stokes Raman scattering,” Opt. Express 18(5), 4371–4379 (2010). [PubMed]  

9. Y. J. Lee, Y. Liu, and M. T. Cicerone, “Characterization of three-color CARS in a two-pulse broadband CARS spectrum,” Opt. Lett. 32(22), 3370–3372 (2007). [CrossRef]   [PubMed]  

10. Y. J. Lee and M. T. Cicerone, “Vibrational dephasing time imaging by time-resolved broadband coherent anti-Stokes Raman scattering microscopy,” Appl. Phys. Lett. 92(4), 041108 (2008). [CrossRef]  

11. H. Kano and H. Hamaguchi, “Femtosecond coherent anti-Stokes Raman scattering spectroscopy using supercontinuum generated from a photonic crystal fiber,” Appl. Phys. Lett. 85(19), 4298–4300 (2004). [CrossRef]  

12. K. B. Shi, P. Li, and Z. W. Liu, “Broadband coherent anti-Stokes Raman scattering spectroscopy in supercontinuum optical trap,” Appl. Phys. Lett. 90(14), 141116 (2007). [CrossRef]  

13. H. N. Paulsen, K. M. Hilligse, J. Thøgersen, S. R. Keiding, and J. J. Larsen, “Coherent anti-Stokes Raman scattering microscopy with a photonic crystal fiber based light source,” Opt. Lett. 28(13), 1123–1125 (2003). [CrossRef]   [PubMed]  

14. E. R. Andresen, H. N. Paulsen, V. Birkedal, J. Thøgersen, and S. R. Keiding, “Broadband multiplex coherent anti-Stokes Raman scattering microscopy employing photonic-crystal fibers,” J. Opt. Soc. Am. B 22(9), 1934–1938 (2005). [CrossRef]  

15. H. G. Breunig, M. Weinigel, R. Bückle, M. Kellner-Höfer, J. Lademann, M. E. Darvin, W. Sterry, and K. König, “Clinical coherent anti-Stokes Raman scattering and multiphoton tomography of human skin with a femtosecond laser and photonic crystal fiber,” Laser Phys. Lett. 10(2), 025604 (2013). [CrossRef]  

16. T. Baldacchini and R. Zadoyan, “In situ and real time monitoring of two-photon polymerization using broadband coherent anti-Stokes Raman scattering microscopy,” Opt. Express 18(18), 19219–19231 (2010). [CrossRef]   [PubMed]  

17. A. F. Pegoraro, A. Ridsdale, D. J. Moffatt, Y. Jia, J. P. Pezacki, and A. Stolow, “Optimally chirped multimodal CARS microscopy based on a single Ti:sapphire oscillator,” Opt. Express 17(4), 2984–2996 (2009). [CrossRef]   [PubMed]  

18. A. A. Ivanov, M. V. Alfimov, and A. M. Zheltikov, “Wavelength-tunable ultrashort-pulse output of a photonic-crystal fiber designed to resolve ultrafast molecular dynamics,” Opt. Lett. 31(22), 3330–3332 (2006). [CrossRef]   [PubMed]  

19. D. A. Sidorov-Biryukov, E. E. Serebryannikov, and A. M. Zheltikov, “Time-resolved coherent anti-Stokes Raman scattering with a femtosecond soliton output of a photonic-crystal fiber,” Opt. Lett. 31(15), 2323–2325 (2006). [CrossRef]   [PubMed]  

20. H. Tu and S. A. Boppart, “Coherent fiber supercontinuum for biophotonics,” Laser Photonics Rev. 7(5), 628–645 (2013). [CrossRef]   [PubMed]  

21. Y. Liu, M. D. King, H. Tu, Y. Zhao, and S. A. Boppart, “Broadband nonlinear vibrational spectroscopy by shaping a coherent fiber supercontinuum,” Opt. Express 21(7), 8269–8275 (2013). [CrossRef]   [PubMed]  

22. N. Dudovich, D. Oron, and Y. Silberberg, “Single-pulse coherently controlled nonlinear Raman spectroscopy and microscopy,” Nature 418(6897), 512–514 (2002). [CrossRef]   [PubMed]  

23. S. Kumar, T. Kamali, J. M. Levitte, O. Katz, B. Hermann, R. Werkmeister, B. Považay, W. Drexler, A. Unterhuber, and Y. Silberberg, “Single-pulse CARS based multimodal nonlinear optical microscope for bioimaging,” Opt. Express 23(10), 13082–13098 (2015). [CrossRef]   [PubMed]  

24. S. H. Lim, A. G. Caster, and S. R. Leone, “Single-pulse phase-control interferometric coherent anti-Stokes Raman scattering spectroscopy,” Phys. Rev. A 72(4), 041803 (2005). [CrossRef]  

25. B. von Vacano and M. Motzkus, “Time-resolved two color single-beam CARS employing supercontinuum and femtosecond pulse shaping,” Opt. Commun. 264(2), 488–493 (2006). [CrossRef]  

26. H. Li, D. A. Harris, B. Xu, P. J. Wrzesinski, V. V. Lozovoy, and M. Dantus, “Coherent mode-selective Raman excitation towards standoff detection,” Opt. Express 16(8), 5499–5504 (2008). [CrossRef]   [PubMed]  

27. K. Furusawa, N. Hayazawa, and S. Kawata, “Two‐beam multiplexed CARS based on a broadband oscillator,” J. Raman Spectrosc. 41(8), 840–847 (2010). [CrossRef]  

28. I. Pope, W. Langbein, P. Watson, and P. Borri, “Simultaneous hyperspectral differential-CARS, TPF and SHG microscopy with a single 5 fs Ti:Sa laser,” Opt. Express 21(6), 7096–7106 (2013). [CrossRef]   [PubMed]  

29. D. Pestov, R. K. Murawski, G. O. Ariunbold, X. Wang, M. 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(5822), 265–268 (2007). [CrossRef]   [PubMed]  

30. A. Volkmer, L. D. Book, and X. S. Xie, “Time-resolved coherent anti-Stokes Raman scattering microscopy: Imaging based on Raman free induction decay,” Appl. Phys. Lett. 80(9), 1505–1507 (2002). [CrossRef]  

31. R. Hildner, D. Brinks, and N. F. van Hulst, “Femtosecond coherence and quantum control of single molecules at room temperature,” Nat. Phys. 7(2), 172–177 (2011). [CrossRef]  

32. S. Yampolsky, D. A. Fishman, S. Dey, E. Hulkko, M. Banik, E. O. Potma, and V. A. Apkarian, “Seeing a single molecule vibrate through time-resolved coherent anti-Stokes Raman scattering,” Nat. Photonics 8(8), 650–656 (2014). [CrossRef]  

33. Y. Zhang, Y.-R. Zhen, O. Neumann, J. K. Day, P. Nordlander, and N. J. Halas, “Coherent anti-Stokes Raman scattering with single-molecule sensitivity using a plasmonic Fano resonance,” Nat. Commun. 5, 4424 (2014). [PubMed]  

34. J. Dudley, X. Gu, L. Xu, M. Kimmel, E. Zeek, P. O’Shea, R. Trebino, S. Coen, and R. Windeler, “Cross-correlation frequency resolved optical gating analysis of broadband continuum generation in photonic crystal fiber: simulations and experiments,” Opt. Express 10(21), 1215–1221 (2002). [CrossRef]   [PubMed]  

35. NKT Photonics application note, “Supercontinuum generation in photonic crystal fibers” (NKT Photonics, 2009). http://www.nktphotonics.com/files/files/Application_note_-_Supercontinuum%20-%20General.pdf

36. Newport Corporation application note, “Supercontinuum generation in SCG-800 photonic crystal fiber” (Newport Corporation, 2006). http://www.newport.com/file_store/Optics_and_Mechanics/AppsNote28.pdf

37. R. Zadoyan, T. Baldacchini, J. Carter, C.-H. Kuo, and D. Ocepek, “CARS module for multimodal microscopy,” In SPIE BiOS, 79030Z, International Society for Optics and Photonics, 2011.

38. J. Herrmann, U. Griebner, N. Zhavoronkov, A. Husakou, D. Nickel, J. C. Knight, W. J. Wadsworth, P. St. J. Russell, and G. Korn, “Experimental evidence for supercontinuum generation by fission of higher-order solitons in photonic fibers,” Phys. Rev. Lett. 88(17), 173901 (2002). [CrossRef]   [PubMed]  

39. S. A. Akhmanov, V. A. Vysloukh, and A. S. Chirkin, Optics of Femtosecond Laser Pulses (AIP, 1992).

40. R. Zadoyan and V. A. Apkarian, “Imaging the molecular rovibrational coherence through time-gated, frequency-resolved coherent anti-Stokes Raman scattering,” Chem. Phys. Lett. 326(1-2), 1–10 (2000). [CrossRef]  

41. A. Bartels, D. Heinecke, and S. A. Diddams, “Passively mode-locked 10 GHz femtosecond Ti:sapphire laser,” Opt. Lett. 33(16), 1905–1907 (2008). [CrossRef]   [PubMed]  

42. V. V. Lozovoy, I. Pastirk, and M. Dantus, “Multiphoton intrapulse interference. IV. Ultrashort laser pulse spectral phase characterization and compensation,” Opt. Lett. 29(7), 775–777 (2004). [CrossRef]   [PubMed]  

43. L. Novotny and N. F. van Hulst, “Antennas for light,” Nat. Photonics 5(2), 83–90 (2011). [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 Experimental apparatus for three-color three-pulse CARS ( λ / 2 – half waveplate, PCBS – polarizing cube beamsplitter, FI – Faraday isolator, AT1, AT2 – attenuator assemblies consisting of half waveplates and Glan-laser polarizers, L1, L2 – aspheric lenses for focusing and collimation, PCF – photonic crystal fiber, WE – wavelength extender, DM1, DM2 – 660 nm and 757 nm long wave pass dichroic beamsplitters, BPF1, BPF2, BPF3, BPF4 – bandpass filters centered at 800 nm, 710 nm, 610 nm, and 554 nm respectively, SPF – 390 nm short wave pass filter, APD – avalanche photodiode, Spec – spectrograph, PMT – photomultiplier tube. Inset shows the SC spectrum on log scale.
Fig. 2
Fig. 2 (a) SC spectra with different output powers, (b) normalized spectra of the probe (612 nm), pump (708 nm), and Stokes (798 nm) pulses sliced from the SC spectrum given by the green trace in (a).
Fig. 3
Fig. 3 Cross-correlations of pump (black dotted) and probe (open orange circles) pulses with the 798 nm reference pulse.
Fig. 4
Fig. 4 (a) Measured XFROG trace, (b) retrieved temporal (black solid) and phase profiles (open circles), together with a quadratic fit (green solid).
Fig. 5
Fig. 5 Measured (black) and simulated (red) tr-CARS traces for (a) BPE and (b) styrene. The simulations are based on the windowed inverse Fourier transform of the Raman spectra (Eq. (3) of text). The Raman spectra and window function (blue) are shown in the insets.

Equations (4)

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

R = ( I L / ω ) σ η f δ t
W R ( ω ) = I P u ( ω ' ) I S t ( ω ' ω ) d ω '
S A S ( t ) = | W R ( ω ) I R ( ω ) e i ω t d ω | 2
R = ( I 0 β 2 / ω ) ( β 2 σ ) η f δ t
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.