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

Single-beam heterodyne FAST CARS microscopy

Open Access Open Access

Abstract

We demonstrate, for the first time, single-beam heterodyne FAST CARS imaging without data post-processing and with nonresonant background subtraction in a simple setup via the real-time piezo modulation of the probe delay. Our fast signal acquisition scheme does not require a spatial light modulator in the pulse shaper, and is suitable for high-resolution imaging and time-resolved dynamics. In addition, the spectral detection of the back-scattered FAST CARS signal is incorporated into the pulse shaper, allowing for a compact and more efficient design. Such epi-detection capability is demonstrated by imaging Si and MoS2 microstructures.

© 2016 Optical Society of America

1. Introduction

Raman spectroscopy is a noninvasive technique that has found wide chemical and biological applications by providing rich information about vibrational levels of molecules and material structures [1–3]. Coherent anti-Stokes Raman scattering (CARS) provides an enhancement of the spontaneous Raman signals by generation of the vibrational coherence, offering high temporal resolution, faster acquisition speed, and avoids the one-photon fluorescence [2–5]. CARS is a third-order nonlinear process in which the pump E(ωp) and Stokes E(ωs) fields generate coherence at ΩR = ωpωs, probed by a third field E(ωpr) to generate a blue-shifted CARS signal at ωsig = ωpr + ΩR. Typically there is also a nonresonant four-wave mixing (FWM) background at ωsig which masks and distorts the CARS signal. Various methods were developed to eliminate this background, including epi-detection [6], polarization control [7], heterodyne CARS [8, 9], nonlinear interferometric vibrational imaging [10, 11], Fourier transform CARS [12], and femtosecond adaptive spectroscopic techniques for CARS (FAST CARS) [13, 14]. Heterodyne CARS mixes a local oscillator (LO) field with the CARS signal and provides an improved shot-noise-limited detection in low-concentration measurements [15]. Most of these CARS schemes require splitting and recombination of ultrashort laser pulses, which is an experimental challenge in spatial and temporal alignment.

Single-beam CARS techniques eliminate the need for complicated alignment, generating CARS signals via coherent control of a broadband femtosecond pulse in a pulse shaper with the aid of a spatial light modulator (SLM) [16]. Other single-beam CARS schemes include polarization [17,18], time-resolved two color [19], binary phase [20], and spectral notch shaping [21,22]. Also, advancement in technology has allowed fast modulation speed and broadband excitation to address a wide range of Raman modes [23–25]. Incorporation of the conventional CARS technique into the single-beam regime has been achieved for heterodyne detection and spectral focusing with improved performance [26,27]. Heterodyne single-beam multiplex CARS was also demonstrated by dividing the femtosecond pulse into a broadband low-frequency pump/Stokes, a narrowband probe, and a high-frequency LO, and the CARS signal was retrieved via double quadrature spectral interferometry (DQSI) method [28].

Previously we demonstrated a single-beam FAST CARS spectroscopic technique, where the narrowband probe was delayed relative to the broadband pump/Stokes field to achieve the non-resonant background suppression at an optimal probe delay [29]. Instead of using the SLM, we translated the probe folding mirror using motorized actuator to achieve a large probe delay. The FWM background generated by the pump/Stokes field was used as the LO, which guaranteed stability [30]. However, the phase relation between the FAST CARS signal and the LO was not precisely controlled, leading to line-shape distortions. Also the background acquisition required long-range shift of the probe mirror actuator, which was time-consuming and not suitable for imaging.

Here we implemented a single-beam heterodyne FAST CARS microscopy by exploiting the phase relation between the signal and the LO. The energy level diagram of the heterodyne FAST CARS scheme is shown in Fig. 1(a), and the spectral intensities of the probe and pump/Stokes fields are plotted in Fig. 1(b). In addition to the large-scale delay controlled by the motorized actuator, we added a piezo actuator for fine tuning the delay. By modulating the probe mirror position locally within a fraction of the wavelength using the piezo actuator, we switched the probe delay between two close values τ0 + δτ and τ0δτ, and obtained two spectra I+(ω) and I(ω), respectively. Here τ0 denotes the central delay value and δτ is the modulation amplitude. By carefully choosing τ0 and δτ, I+(ω) and I(ω) will correspond to the constructive and destructive interference between the FAST CARS signal and the LO, as shown by the simulation data in Fig. 1(c). Then by using a modified DQSI method [18]:

ICARS=I+(ω)I(ω)I+(ω)+I(ω),
fast reconstruction of the vibrational signal and subtraction of the nonresonant background was achieved, with clearly observable Raman peaks shown in Fig. 1(d). In addition, an improved epi-detection efficiency is obtained by incorporating the spectral detection into the pulse shaper, which reduces redundant optical components for re-collimating the backscattered signal and sending it to an external spectrometer [22,29]. This compact design may be used for improving nonlinear microscopy systems.

 figure: Fig. 1

Fig. 1 Schematic of the single-beam heterodyne FAST CARS. (a) Energy diagram of the FAST CARS scheme. (b) Spectral intensity of the excitation pulse with the probe and pump/Stokes parts marked with red and black lines, respectively. (c) Anti-Stokes signals I+(ω) (red) and I(ω) (blue) simulated corresponding to delays τ0 + δτ and τ0δτ, respectively, with τ0 ≈ 0. Inset shows the constructive (red) and destructive (blue) interferences between the FAST CARS signal and the FWM LO. (d) FAST CARS signal retrieved from the spectra in (c) using Eq. (1).

Download Full Size | PDF

2. Results and discussion

Results obtained from 1,1,2,2- tetrachloroethane (TeCA) are shown in Fig. 2, demonstrating the probe-delay effect on the retrieved Raman peaks. Figure 2(a) shows the spectrogram of TeCA acquired for different probe-delay times, in which we can see the tails of various Raman modes due to their long-lasting coherences. Nonresonant background was reduced at all delay values τ0, and we plot three spectra at delays of 0 ps, 0.6 ps, and 1.2 ps in Fig. 2(b), and they are offset for visibility. For comparison, we also plot the zero probe delay spectrum from our previous work [29] in Fig. 2(c), in which the resonant signal is buried in the large nonresonant background. As the delay τ0 was increased from 0 ps to τ0 ≈ 0.6 ps, the estimated height of the main peak around 353 cm−1 has increased by around 30%, and the full width at half maximum decreased from 32 cm−1 to 21 cm−1 compared to the zero probe delay signal. Other peaks showed similar behavior, and the effect is similar to the π-step-phased probe case in previous literature [31]. Also noticeable is the lineshape distortion to the retrieved peaks, with dispersive shapes forming on both sides of the 353 cm−1 peak at 0.6 ps delay. When the delay was further increased to τ0 ≈ 1.2 ps, the distortion became so large that the peak height information was lost, but the central peak became more distinct. One interesting point to note is that besides the main peak at 353 cm−1, there is also a spectral structure appearing at 370 cm−1. Its origin could be the 366 cm−1 vibrational mode of TeCA, whose spontaneous Raman spectrum is shown in Fig. 2(d). We note that at large τ0 values, the delay induces a spectral phase across the probe spectrum and modifies its interaction with the nonlinear susceptibility, which can lead to the appearance of the doublet structure that would otherwise be hidden as in the zero probe delay case.

 figure: Fig. 2

Fig. 2 Single-beam heterodyne FAST CARS of TeCA using the forward detection: (a) 2D plot of the heterodyne FAST CARS spectra at different probe delays. (b) Heterodyne FAST CARS signal at 0 (black), 0.6 (red), and 1.2 ps (blue) delay. (c) FAST CARS signal of TeCA at zero probe delay from [29]. (d) Spontaneous Raman spectrum of TeCA. Spectra in (b) are normalized to the same factor.

Download Full Size | PDF

Next we carried out measurements probing coherence lifetimes of the Raman modes using fused silica. It was previously shown that fused silica contains a broad vibrational band, denoted as ω1, due to an internal continuous random network with a short decoherence time [32]. The retrieved FAST CARS signals of fused silica at different probe delay times are shown in Fig. 3. We identify the ω1 band as the broad spectral feature in the range between 200 cm−1 and 440 cm−1. As the probe was further delayed relative to the pump/Stokes, this band intensity gradually decreased and only the narrow peak due to the four-fold ring structure at 495 cm−1 [32] was preserved. The coherence lifetime of the ω1 band is estimated to be shorter than 1 ps due to the convolution of the FAST CARS signal with the probe temporal shape, which limits the temporal resolution to ∼1 ps.

 figure: Fig. 3

Fig. 3 FAST CARS spectra of fused silica using the forward detection at different probe delays: (a) 0 ps, (b) 0.2 ps, (c) 0.4 ps, (d) 0.6 ps, (e) 0.8 ps, and (f) 1 ps.

Download Full Size | PDF

We used carbon tetrachloride (CCl4) as the test sample to verify the power and concentration dependence. Because the raw signal intensities I±(ω), generated by third-order nonlinear process, are proportional to the third power of the total input laser intensity I±(ω) ∝ (Input power)3, then Eq. (1) the retrieved signal has a 1.5th-power dependence on the input power:

ICARS(Inputpower)1.5.
And for heterodyne detection we expect a linear dependence on the concentration [8]. Figure 4(a) shows the retrieved spectrum from CCl4 at zero probe delay. Figures 4(b) and 4(c) show the signal dependence on the input power and concentration, respectively, on a log-log scale, where the input power is measured in mW and concentration in percentage. The input laser power was measured before entering the pulse shaper, and Fig. 4(c) was obtained using different concentrations of CCl4 in acetone at fixed laser power. The intensities shown in Figs. 4(b) and 4(c) correspond to the 459 cm−1 transition of CCl4. The slopes of the fitting curves in Figs. 4(b) and 4(c) are 1.49±0.02 and 1.04±0.05, respectively, which agrees with the theory. Other transitions showed similar behavior.

 figure: Fig. 4

Fig. 4 Input power and concentration dependence of the CCl4 FAST CARS signal measured in the forward detection. (a) Retrieved FAST CARS spectrum from CCl4 at zero probe delay. (b) Log-log plot of the intensity of the 459 cm−1 peak as a function of the laser power. (c) Intensity of the 459 cm−1 peak measured at different concentrations of CCl4 in acetone.

Download Full Size | PDF

Finally we demonstrate the epi-detection capability of the setup by imaging microscopic structures of the gold-coated silicon (Si) and molybdenum disulfide (MoS2) flakes. The samples were scanned using piezo-driven stages, and characteristic peak intensities using 1 ps probe delay value at each spatial location were collected to obtain the FAST CARS images. Figures 5(a) and 5(d) are the microscopic optical images of the Si and MoS2 samples, respectively, with marked scan areas. Figures 5(b) and 5(e) compare two FAST CARS spectra acquired at different pixels from the two samples, with one pixel containing the target substance and the other not. The characteristic Raman peak values were used for imaging: for Si the 520 cm−1 Raman mode was used, and for MoS2 the peak corresponding to the E2g transition around 385 cm−1 was used. The retrieved FAST CARS images are shown in Figs. 5(c) and 5(f), which are in good agreement with the optical images for both samples.

 figure: Fig. 5

Fig. 5 Epi-detection FAST CARS microscopy of Si and MoS2 microstructures. Optical images of the structured Si (a) and MoS2 flake (d) samples, with scale bars of 2μm. FAST CARS spectra of Si (b) and MoS2 (e) plotted in black lines compared to the background in red lines, with corresponding FAST CARS images in (c) and (f), respectively, using the Si peak at 520 cm−1 and the MoS2 peak at 385 cm−1. The two spectra in (b) were acquired at (6.4 μm, 6.4 μm) (black) and (4.0 μm, 4.0 μm) (red) corresponding to the image in (c), and the two spectra in (e) were acquired at (4.6 μm, 3.8 μm) (black) and (1.0 μm, 2.2 μm) (red) corresponding to (f).

Download Full Size | PDF

The appropriate power assignment to the probe and pump/Stokes field could contribute to an enhanced signal level. For the experiment on TeCA described earlier, the total power on the sample was 8.3 mW, with a probe to pump/Stokes power ratio ∼ 1:75. By placing an ND filter in front of the pump/Stokes mirror to solely attenuate this part, we modified the probe to pump/Stokes power ratio to ∼ 1:9.3, and the total power on the sample changed to 7.8 mW. We found that for TeCA the signal-to-noise ratio of the primary retrieved peak at 355 cm−1 is doubled as compared to the previous case using 8.3 mW total power but with a smaller probe to pump/Stokes power ratio. Because the LO is also dependent on the pump/Stokes field, careful consideration is needed to reach the optimal ratio between the probe and pump/Stokes power for the FAST CARS signal retrieval.

The limited sensitivity and acquisition speed of the current setup can be improved from various perspectives. Higher pulse energy lasers such as cavity-dumping oscillator [31] or high-repetition-rate regenerative amplifier system [11] could lead to an order of magnitude signal enhancement and therefore less acquisition time with better sensitivity. Also, the piezo actuator controlling the probe mirror can be operated faster than 10 kHz, so higher detection speed can be achieved by adopting fast acquisition detector like a frame-transfer CCD, scientific CMOS camera, or the recently devised TAMP array [33]. Another limitation concerns the bandwidth, that is, given a broader spectral coverage, the full fingerprint region can be addressed, therefore leading to broader applications in chemistry and biology. Currently we are limited to multiplex detection in the deep fingerprint region, but the setup can easily be modified for broadband sources such as sub-10 fs oscillators [27,28], or oscillator-pumped supercontinuum [9,25].

3. Methods

3.1. Samples

TeCA and CCl4 samples were commercial products (SigmaAldrich), and fused silica sample was from UV-grade window material (Newport). The Si sample was prepared using the e-beam lithography on Si with the PMMA patterned mask followed by coating with 50 nm of gold; the MoS2 flakes were prepared by exfoliation of bulk MoS2 using tape onto a Si substrate.

3.2. Experimental setup

The sketch of our experimental setup is shown in Fig. 6. Details of the pulse shaper have been previously described [29]. A Ti:Sapphire oscillator (KMLabs, TS kit) was pumped by a green laser (Millenia eV, Spectra Physics) to produce 25 fs pulses centered around 810 nm, with a repetition rate of 85 MHz. A variable ND filter was used for attenuation, and followed by a chirp mirror pair for dispersion compensation. Then the high-frequency part of the pulse that coincides with the CARS signal was blocked by a long-pass filter (LP02-785RU-25, Semrock), and the pulse was sent to the pulse shaper where its spectrum was dispersed across the Fourier plane [34]. A major part of the pulse on the low frequency side was reflected by the mirror M1 that served as pump/Stokes, and a narrowband portion was reflected by the mirror M2, serving as probe. A right angle mirror M3 was placed closely in front of M2 such that the edges of M2 and M3 determine the bandwidth of the probe, which was chosen as 30 cm−1. M2 was mounted on a linear stage with a motorized actuator for coarse tuning τ0 and a piezo actuator for fine tuning τ0 and the modulation δτ. The following procedure was used to determine the relative position between M1 and M2: distilled water was used as the sample, and a large range of probe delays was scanned using the motorized actuator, with spectra taken at each increment of the stepper motor. Then by plotting the intensity at a specific wavelength on the signal curve versus the probe delay we retrieved a fringe pattern similar to an autocorrelation trace, resulted from the interference between the LO and the nonresonant signal generated by the probe. The envelope of this pattern indicates the temporal shape of the probe pulse, and we can determine the zero delay position according to this envelope. Then the probe delay value τ0 can be calculated according to the motorized actuator shift from the zero delay position. Also, optimal modulation amplitude for the piezo actuator was found to be ∼190 nm to maximize the retrieved Raman peaks, which corresponds to δτ ≈ 1.3 fs.

 figure: Fig. 6

Fig. 6 Experimental setup: DC, chirp mirror pair for dispersion compensation; LPF, long-pass filter; G, grating; CM, concave mirror; M, mirror; SPF, short-pass filter; L, lens; Obj, microscope objective; S, sample on X–Y translation stage; spec, spectrometer. The excitation and FAST CARS beams are shown by red and green colors, respectively.

Download Full Size | PDF

The reflected beams from M1 and M2 were tilted so that they could be guided out collinearly and sent to the microscope, where both beams were focused on the sample by an objective (N40X-NIR, Nikon). For transparent samples, the transmitted light was collimated by another objective (LPlan50X, Nikon) and sent through a short-pass filter (TSP01-790-25x36, Semrock) before entering the spectrometer (SureSpectrum, Bruker Optics) for detection; for opaque samples, back-scattered light was collected by the same focusing objective and reflected back towards the pulse shaper. After reaching the Fourier plane, the blue-shifted signal was guided by M3 to the telescope consisting of the concave mirror CM2 and the lens L, and imaged using the CCD camera (Spec10, Princeton Instruments). The focal lengths of CM2 and L were 500 mm and 200 mm, respectively. The same short-pass filter was placed inside the telescope for filtering the residual excitation light. Both the forward arm spectrometer and the CCD in the epi-detection arm were calibrated by a spectral lamp, and the Raman shifts were calculated from the frequency shift from the probe central wavelength which was measured by the forward arm spectrometer.

TeCA, fused silica, and CCl4 were transparent samples and were measured in the forward direction; Si and MoS2 samples were imaged using epi-detection. The laser power on the sample without attenuation by the ND filter was 50 mW, and typical laser powers used on each sample were 9.8, 8.3, 19.2, 42, and 16.7 mW for CCl4, TeCA, fused silica, Si, and MoS2, respectively, except for the power dependent measurement. The CCD exposure time for all samples was fixed at 40 ms, and CCl4, TeCA, and silicon spectra were averaged 10 times, while fused silica and MoS2 spectra were averaged 100 times due to their relatively weak CARS signals. During the acquisition the CCD received a trigger frequency of 14 Hz while the piezo was driven at 7 Hz such that the two spectra were collected at constructive and destructive interference positions, from which the FAST CARS signals were retrieved using Eq. (1). The image acquisition times for Si and for MoS2 shown in Fig. 5 were 12 minutes and 2 hours, respectively. For comparison, spontaneous Raman spectra were acquired using a commercial Raman system (LabRAM HR Evolution, Horiba).

3.3. Theory

We denote E(ω) as the excitation laser field, which is separated into the probe field Epr (ω) with a bandwidth of δω, and the pump/Stokes field Eps (ω), as shown in Fig. 1(b). For the transform limited pulses, the electric fields are taken to have real values. The nonlinear polarization is given by [3]:

P(ω)[Epr(ωΩ)χ(Ω)+Eps(ωΩ)χ(Ω)]A(Ω)dΩ,
where
A(Ω)=E*(ω)E(Ω+ω)dω,
χ(Ω)=χNR(Ω)+χR(Ω)=χNR(Ω)+mAmΩRmΩ+iΓRm.
A(Ω) corresponds to the vibrational coherence due to impulsive excitation, and χ(Ω) is the nonlinear susceptibility, which has a real nonresonant part χNR and a complex resonant part χR with different Raman modes with frequencies ΩRm, transition amplitudes Am, and linewidths ΓRm. Typical Raman lines come from the imaginary part of this susceptibility Im[ χ(Ω)]. We also denote PFWMps(ω)=Eps(ωΩ)χ(Ω)A(Ω)dΩ, which is a strong FWM polarization generated by the pump/Stokes field in the CARS signal region, and is used as LO.

When a time delay τ relative to the pump/Stokes field is added to the probe, it is equivalent to adding a linear spectral phase to Epr (ω) in the frequency domain, and Eq. (3) is modified as:

P(ω)Epr(ωΩ)ei(ωΩ)τχ(Ω)A(Ω)dΩ+PFWMps(ω).
Now we consider two spectra taken at two different delays τ0 + δτ and τ0δτ, where τ0 represents central probe-delay and δτ is a small modulation on the order of fs. Because the bandwidth of the probe δω is typically tens of wavenumbers, the phase term ei(ω−Ω)δτ varies little when integrating over the probe region and can be approximated as eprδτ. We assign the following function:
f(ω)Epr(ωΩ)ei(ωΩ)τ0χ(Ω)A(Ω)dΩ,
so that the two polarizations P+(ω) and P(ω), corresponding to τ0 + δτ and τ0δτ, can be written as:
P+(ω)={Re[f(ω)]cos(ωprδτ)+Im[f(ω)]sin(ωprδτ)}+i{Im[f(ω)]cos(ωprδτ)Re[f(ω)]sin(ωprδτ)}+PFWMps(ω),
P(ω)={Re[f(ω)]cos(ωprδτ)Im[f(ω)]sin(ωprδτ)}+i{Im[f(ω)]cos(ωprδτ)+Re[f(ω)]sin(ωprδτ)}+PFWMps(ω).
For simplicity in analysis, we first assume PFWMps(ω) to be real. The detected signal intensities I+(ω) ∝ |P+(ω)|2 and I(ω) ∝ |P(ω)|2 correspond to the constructive and destructive interference between the signal and the LO, respectively, and we obtain the difference:
I+(ω)I(ω)|P+(ω)|2|P(ω)|24PFWMps(ω)Im[f(ω)]sin(ωprδτ),
in which we keep only the heterodyne term because |PFWMps(ω)||f(ω)|. Further approximating PFWMps(ω)I+(ω)+I(ω), and fixing δτ = π/(2ωpr), we obtain from Eq. (1):
ICARS=I+(ω)I(ω)I+(ω)+I(ω)Im[f(ω)].

The above equation relates the detected CARS signal to the function f(ω). A closer look at this function in Eq. (7) reveals that it is essentially the susceptibility χ(Ω) gated by the probe with phase Epr (ω − Ω)ei(ω−Ω)τ0 and a large smooth envelope function A(Ω). For ideal probe shape Epr (ω) assumes the form of a delta function with infinitely small bandwidth so that f(ω) reproduces the shape of χ(Ω). In reality Epr (ω) always has a finite bandwidth, leading to a compromised resolution. In the FAST CARS case here it is further complicated by the phase factor ei(ω−Ω)τ0. Using Eqs. (7) and (11) we can express ICARS as:

ICARSEpr(ωΩ)A(Ω){Im[χ(Ω)]cos[(ωΩ)τ0]Re[χ(Ω)]sin[(ωΩ)τ0]}dΩ.
A simple situation is when τ0 is small such that (ω − Ω)τ0 does not vary too much over the probe spectral region and can therefore be approximated by ωpr τ0. Then by choosing cos(ωpr τ0) = 1 the term containing Re[χ(Ω)] in Eq. (1) is canceled, and we see that ICARS is directly related to Im[χ(Ω)]. In the case of larger delay τ0, the same condition cos(ωpr τ0) = 1 should be satisfied, so that the nonresonant background term Re[ χNR(Ω)]sin[(ω − Ω)τ0] would see an equal contribution from the positive and negative part of the sine factor when integrating over the spectral region of the probe and yield vanishing contribution.

Next we consider the neglected imaginary part of the LO field PFWMps(ω), which is due to the contribution from χR(Ω) in the expression for PFWMps(ω). Consider the expression PFWMps(ω)=|PFWMps(ω)|eiθ(ω), where θ(ω) denotes the spectral phase of the LO field. Then the polarizations P+(ω) and P(ω) change correspondingly, leading to a modification to the difference in Eq. (10):

I+(ω)I(ω)4sin(ωprδτ){Re[PFWMps(ω)]Im[f(ω)]Im[PFWMps(ω)]Re[f(ω)]},
which results in the correction term involving Im[PFWMps(ω)]. Expanding the above equation using Eq. (7), and setting ωpr δτ = π/2, we obtain a modified ICARS expression compared to Eq. (12):
ICARSEpr(ωΩ)A(Ω){Im[χ(Ω)]cos[(ωΩ)τ0+θ(ω)]Re[χ(Ω)]sin[(ωΩ)τ0+θ(ω)]}dΩ.
In the situation where |χNR| ≫ |χR|, the spectral phase θ(ω) is small and simulations show that the first term containing Im[χ(Ω)] again provides the Raman lines, and the second term containing Re[χ(Ω)] contributes to a minor background which can be offset by adjusting the actual τ0 compared to the theoretical value by less than 0.1 fs. When the ratio χRNR increases, the oscillation due to the phase enters the signal, and causes severe lineshape distortion to the retrieved CARS spectrum— turning it into almost a dispersive shape. A possible solution could be to use the high-frequency part of the excitation pulse as LO [28], so that the LO field becomes real. This may be achieved in the current setup by replacing M3 with a variable density ND filter. Such LO is also advantageous in that its intensity can be controlled to obtain an optimal LO power level for maximizing the signal-to-noise ratio of the retrieved FAST CARS signal [15].

The simulated curves in Figs. 1(c) and 1(d) were obtained by setting the Raman modes as Ω1 = 300 cm−1, Ω2 = 650 cm−1, Ω3 = 1000 cm−1, with Γ1 = 15 cm−1, Γ2 = 10 cm−1, and Γ3 = 20 cm−1. The probe pulse width was δω = 30.5 cm−1, and the delay parameters τ0 = 0.06 fs and δτ = 0.65 fs.

4. Conclusion

In conclusion, we demonstrated single-beam heterodyne FAST CARS microscopy using the temporal control of the probe delay inside the pulse shaper. Phase shifting between the signal and LO was achieved using real-time piezo modulation, and the nonresonant background was reduced using hetero-dyne detection, allowing for imaging applications as well as the study of the dynamic evolution of FAST CARS signals at various probe delays. We also improved the epi-detection design by incorporating spectral detection into the pulse shaper, which is more efficient and compact compared to previous work [22,29].

Funding

National Science Foundation (NSF) Grants (PHY-1241032 (INSPIRE CREATIV), PHY-1068554, EEC-0540832 (MIRTHE ERC), PHY-1307153), CHE-1609608; Robert A. Welch Foundation Award (A-1261, A-1547); Office of Naval Research (ONR) Award (N00014-16-1-2578).

Acknowledgments

Yujie Shen is supported by the Herman F. Heep and Minnie Belle Heep Texas A&M University Endowed Fund held/administered by the Texas A&M Foundation. We thank Prof. Zhenrong Zhang for providing the MoS2 sample.

References and links

1. N. B. Colthup, L. H. Daly, and S. E. Wiberley, Introduction to Infrared and Raman Spectroscopy (Academic, 1990).

2. J.-X. Cheng and X. S. Xie, Coherent Raman Scattering Microscopy (CRC, 2012).

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

4. H. Li, D. A. Harris, B. Xu, P. J. Wrzesinski, V. V. Lozovoy, and M. Dantus, “Standoff and arms-length detection of chemicals with single-beam coherent anti-Stokes Raman scattering,” Appl. Opt. 48(4), B17–B22 (2009). [CrossRef]   [PubMed]  

5. 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]  

6. A. Volkmer, J.-X. Cheng, and X. S. Xie, “Vibrational imaging with high sensitivity via epidetected coherent anti-Stokes Raman scattering microscopy,” Phys. Rev. Lett. 87(2), 023901 (2001). [CrossRef]  

7. J.-X. Cheng, L. D. Book, and X. S. Xie, ”Polarization coherent anti-Stokes Raman scattering microscopy,” Opt. Lett. 26(17), 1341–1343 (2001). [CrossRef]  

8. E. O. Potma, C. L. Evans, and X. S. Xie, ”Heterodyne coherent anti-Stokes Raman scattering (CARS) imaging,” Opt. Lett. 31(2), 241–243 (2006). [CrossRef]  

9. C. H. Camp Jr, Y. J. Lee, J. M. Heddleston, C. M. Hartshorn, A. R. Hight Walker, J. N. Rich, J. D. Lathia, and M. T. Cicerone, ”High-speed coherent Raman fingerprint imaging of biological tissues,” Nature Photonics 8(8), 627–634 (2014). [CrossRef]  

10. D. L. Marks and S. A. Boppart, “Nonlinear interferometric vibrational imaging,” Phys. Rev. Lett. 92(12), 123905 (2004). [CrossRef]   [PubMed]  

11. P. D. Chowdary, Z. Jiang, E. J. Chaney, W. A. Benalcazar, D. L. Marks, M. Gruebele, and S. A. Boppart, “Molecular histopathology by spectrally reconstructed nonlinear interferometric vibrational imaging,” Cancer research 70(23), 9562–9569 (2010). [CrossRef]   [PubMed]  

12. J. P. Ogilvie, E. Beaurepaire, A. Alexandrou, and M. Joffre, “Fourier-transform coherent anti-Stokes Raman scattering microscopy,” Opt. Lett. 31(4), 480–482 (2006). [CrossRef]  

13. M. O. Scully, G. W. Kattawar, R. P. Lucht, T. Opatrný, 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(17), 10994–11001 (2002). [CrossRef]   [PubMed]  

14. 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]  

15. M. Jurna, J. P. Korterik, C. Otto, and H. L. Offerhaus, ”Shot noise limited heterodyne detection of CARS signals,” Opt. Express 15(23), 15207–15213 (2007). [CrossRef]   [PubMed]  

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

17. D. Oron, N. Dudovich, and Y. Silberberg, “Femtosecond phase-and-polarization control for background-free coherent anti-Stokes Raman spectroscopy,” Phys. Rev. Lett. 90(21), 213902 (2003). [CrossRef]   [PubMed]  

18. S.-H. Lim, A. G. Caster, O. Nicolet, and S. R. Leone, ”Chemical imaging by single pulse interferometric coherent anti-Stokes Raman scattering microscopy,” J. Phys. Chem. B 110(11), 5196–5204 (2006). [CrossRef]   [PubMed]  

19. 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]  

20. P. J. Wrzesinski, D. Pestov, V. V. Lozovoy, B. Xu, S. Roy, J. R. Gord, and M. Dantus, ”Binary phase shaping for selective single-beam CARS spectroscopy and imaging of gas-phase molecules,” J. Raman Spectrosc. 42(3), 393–398 (2011). [CrossRef]  

21. 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]  

22. Y. Shen, D. V. Voronine, A. V. Sokolov, and M. O. Scully, ”Low-wavenumber efficient single-beam coherent anti-Stokes Raman scattering using a spectral hole,” Opt. Lett. 40(7), 1223–1226 (2015). [CrossRef]   [PubMed]  

23. J. M. Levitt, O. Katz, and Y. Silberberg, “Frequency-encoded multiplexed CARS microscopy by rapid pulse shaping,” J. Mod. Opt. 61(10), 872–876 (2014). [CrossRef]  

24. K. Isobe, A. Suda, M. Tanaka, H. Hashimoto, F. Kannari, H. Kawano, H. Mizuno, A. Miyawaki, and K. Midorikawa, “Single-pulse coherent anti-Stokes Raman scattering microscopy employing an octave spanning pulse,” Opt. Express 17(14), 11259–11266 (2009). [CrossRef]   [PubMed]  

25. 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]  

26. C. Müller, T. Buckup, B. von Vacano, and M. Motzkus, ”Heterodyne single-beam CARS microscopy,” J. Raman Spectrosc. 40(7), 809–816 (2009). [CrossRef]  

27. L. Brückner, T. Buckup, and M. Motzkus, “Enhancement of coherent anti-Stokes Raman signal via tailored probing in spectral focusing,” Opt. Lett. 40(22) 5204–5207 (2015). [CrossRef]   [PubMed]  

28. A. Wipfler, T. Buckup, and M. Motzkus, ”Multiplexing single-beam coherent anti-Stokes Raman spectroscopy with heterodyne detection,” Appl. Phys. Lett. 100(7), 071102 (2012). [CrossRef]  

29. Y. Shen, D. V. Voronine, A. V. Sokolov, and M. O. Scully, ”A versatile setup using femtosecond adaptive spectroscopic technique for coherent anti-Stokes Raman scattering,” Rev. Sci. Instrum. 86(8), 083107 (2015). [CrossRef]   [PubMed]  

30. B. Li, W. S. Warren, and M. C. Fischer, ”Phase-cycling coherent anti-Stokes Raman scattering using shaped femtosecond laser pulses,” Opt. Express 18(25), 25825–25832 (2010). [CrossRef]   [PubMed]  

31. B.-C. Chen and S.-H. Lim, “Optimal laser pulse shaping for interferometric multiplex coherent anti-Stokes Raman scattering microscopy,” J. Phys. Chem. B 112(12), 3653–3661 (2008). [CrossRef]   [PubMed]  

32. F. L. Galeener, R. A. Barrio, E. Martinez, and R. J. Elliott, ”Vibrational decoupling of rings in amorphous solids,” Phys. Rev. Lett. 53(25), 2429 (1984). [CrossRef]  

33. C.-S. Liao, M. N. Slipchenko, P. Wang, J. Li, S.-Y. Lee, R. A. Oglesbee, and J.-X. Cheng, ”Microsecond scale vibrational spectroscopic imaging by multiplex stimulated Raman scattering microscopy,” Light Sci. Appl. 4(3), e265 (2015). [CrossRef]   [PubMed]  

34. A. M. Weiner, “Femtosecond pulse shaping using spatial light modulators,” Rev. Sci. Instrum. 71(5), 1929–1960 (2000). [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 (6)

Fig. 1
Fig. 1 Schematic of the single-beam heterodyne FAST CARS. (a) Energy diagram of the FAST CARS scheme. (b) Spectral intensity of the excitation pulse with the probe and pump/Stokes parts marked with red and black lines, respectively. (c) Anti-Stokes signals I+(ω) (red) and I(ω) (blue) simulated corresponding to delays τ0 + δτ and τ0δτ, respectively, with τ0 ≈ 0. Inset shows the constructive (red) and destructive (blue) interferences between the FAST CARS signal and the FWM LO. (d) FAST CARS signal retrieved from the spectra in (c) using Eq. (1).
Fig. 2
Fig. 2 Single-beam heterodyne FAST CARS of TeCA using the forward detection: (a) 2D plot of the heterodyne FAST CARS spectra at different probe delays. (b) Heterodyne FAST CARS signal at 0 (black), 0.6 (red), and 1.2 ps (blue) delay. (c) FAST CARS signal of TeCA at zero probe delay from [29]. (d) Spontaneous Raman spectrum of TeCA. Spectra in (b) are normalized to the same factor.
Fig. 3
Fig. 3 FAST CARS spectra of fused silica using the forward detection at different probe delays: (a) 0 ps, (b) 0.2 ps, (c) 0.4 ps, (d) 0.6 ps, (e) 0.8 ps, and (f) 1 ps.
Fig. 4
Fig. 4 Input power and concentration dependence of the CCl4 FAST CARS signal measured in the forward detection. (a) Retrieved FAST CARS spectrum from CCl4 at zero probe delay. (b) Log-log plot of the intensity of the 459 cm−1 peak as a function of the laser power. (c) Intensity of the 459 cm−1 peak measured at different concentrations of CCl4 in acetone.
Fig. 5
Fig. 5 Epi-detection FAST CARS microscopy of Si and MoS2 microstructures. Optical images of the structured Si (a) and MoS2 flake (d) samples, with scale bars of 2μm. FAST CARS spectra of Si (b) and MoS2 (e) plotted in black lines compared to the background in red lines, with corresponding FAST CARS images in (c) and (f), respectively, using the Si peak at 520 cm−1 and the MoS2 peak at 385 cm−1. The two spectra in (b) were acquired at (6.4 μm, 6.4 μm) (black) and (4.0 μm, 4.0 μm) (red) corresponding to the image in (c), and the two spectra in (e) were acquired at (4.6 μm, 3.8 μm) (black) and (1.0 μm, 2.2 μm) (red) corresponding to (f).
Fig. 6
Fig. 6 Experimental setup: DC, chirp mirror pair for dispersion compensation; LPF, long-pass filter; G, grating; CM, concave mirror; M, mirror; SPF, short-pass filter; L, lens; Obj, microscope objective; S, sample on X–Y translation stage; spec, spectrometer. The excitation and FAST CARS beams are shown by red and green colors, respectively.

Equations (14)

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

I CARS = I + ( ω ) I ( ω ) I + ( ω ) + I ( ω ) ,
I CARS ( Input power ) 1.5 .
P ( ω ) [ E pr ( ω Ω ) χ ( Ω ) + E p s ( ω Ω ) χ ( Ω ) ] A ( Ω ) d Ω ,
A ( Ω ) = E * ( ω ) E ( Ω + ω ) d ω ,
χ ( Ω ) = χ N R ( Ω ) + χ R ( Ω ) = χ N R ( Ω ) + m A m Ω R m Ω + i Γ R m .
P ( ω ) E pr ( ω Ω ) e i ( ω Ω ) τ χ ( Ω ) A ( Ω ) d Ω + P FWM ps ( ω ) .
f ( ω ) E pr ( ω Ω ) e i ( ω Ω ) τ 0 χ ( Ω ) A ( Ω ) d Ω ,
P + ( ω ) = { Re [ f ( ω ) ] cos ( ω pr δ τ ) + Im [ f ( ω ) ] sin ( ω pr δ τ ) } + i { Im [ f ( ω ) ] cos ( ω pr δ τ ) Re [ f ( ω ) ] sin ( ω pr δ τ ) } + P FWM ps ( ω ) ,
P ( ω ) = { Re [ f ( ω ) ] cos ( ω pr δ τ ) Im [ f ( ω ) ] sin ( ω pr δ τ ) } + i { Im [ f ( ω ) ] cos ( ω pr δ τ ) + Re [ f ( ω ) ] sin ( ω pr δ τ ) } + P FWM ps ( ω ) .
I + ( ω ) I ( ω ) | P + ( ω ) | 2 | P ( ω ) | 2 4 P FWM ps ( ω ) Im [ f ( ω ) ] sin ( ω pr δ τ ) ,
I CARS = I + ( ω ) I ( ω ) I + ( ω ) + I ( ω ) Im [ f ( ω ) ] .
I CARS E pr ( ω Ω ) A ( Ω ) { Im [ χ ( Ω ) ] cos [ ( ω Ω ) τ 0 ] Re [ χ ( Ω ) ] sin [ ( ω Ω ) τ 0 ] } d Ω .
I + ( ω ) I ( ω ) 4 sin ( ω pr δ τ ) { Re [ P FWM ps ( ω ) ] Im [ f ( ω ) ] Im [ P FWM ps ( ω ) ] Re [ f ( ω ) ] } ,
I CARS E pr ( ω Ω ) A ( Ω ) { Im [ χ ( Ω ) ] cos [ ( ω Ω ) τ 0 + θ ( ω ) ] Re [ χ ( Ω ) ] sin [ ( ω Ω ) τ 0 + θ ( ω ) ] } d Ω .
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.