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

A theoretical study of diffraction limit breaking via coherent control of the relative phase in coherent anti-Stokes Raman scattering microscopy

Open Access Open Access

Abstract

We present a theoretical investigation of how coherent control of the relative phase in coherent anti-Stokes Raman scattering (CARS) microscopy can break the diffraction limit. In quantum theory, it is found that the relative phase of the pump and Stokes pulses can be used to periodically tune the intensity of the anti-Stokes signal. Thus, by controlling the relative phase around the center of the pump and Stokes pulses, the anti-Stokes signal can be tuned to zero in this region. In turn, the useful spot-generating anti-Stokes signal is substantially suppressed to a much smaller dimension, and scanning of the spots renders CARS images with sub-diffraction resolutions. Such super-resolutions can greatly enhance the advantage of using CARS microscopy in many potential applications.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Far-field optical microscopy is a basic and useful technique for sample analysis in bioscience. In addition to classical transmission and reflection microscopy, fluorescence microscopy and nonlinear microscopy techniques have developed rapidly in the recent past, such as two-photon-excited luminescence (TPEL) [1], two-photon-excited-fluorescence (TPEF) [2], second harmonic generation (SHG) [3,4], third-harmonic generation (THG) [5–7], stimulated Raman scattering (SRS) [8–10] and coherent anti-Stokes Raman scattering (CARS) [11–15]. With proper manipulation of the illumination field or sample properties, it is possible to extract spatial features that are much smaller than the diffraction limit, leading to super-resolution optical microscopy, which has become a hot research topic in recent years. In fluorescence microscopy, several super-resolution techniques have been successfully demonstrated in experiments, such as photoactivated localization microscopy (PLAM) [16], stochastic optical reconstruction microscopy (STORM) [17], and stimulated emission depletion (STED) microscopy [18,19], which rely on the ability to control the emissive properties of fluorophores with optical radiation. However, significant disadvantages of fluorescence include: photobleaching, toxicity, and the effects of fluorophores on labeled molecules in live samples.

For nonlinear microscopy that does not include exogenous labeling agents of chemical species, several super-resolution methods have been proposed theoretically, such as super-resolution CARS microscopy [20–28] and super-resolution SRS microscopy [29,30]. These methods require new and powerful light sources, the optical systems are quite complicated to build, and the high-energy photons used in their methods can damage biological tissue. In this study, with the intention of avoiding these defects, we propose a scheme to suppress the CARS signal by controlling the relative phase between the pump light and the Stokes light to realize super-resolution CARS microscopy.

2. Principle

CARS microscopy is a third-order nonlinear optical imaging method in which two laser pulses at different frequencies, denoted as pump light (ωp) and Stokes light (ωs), are focused onto a sample. Raman resonance occurs whenωpωsmatches a Raman-active molecular vibrational level. CARS system detects the outputted anti-Stokes signal (ωas) at a new frequency2ωpωswhich is enhanced by the Raman resonance, and should satisfy the frequency-matchingΔω=2ωpωsωas=0. In addition, the pump, Stokes, and anti-Stokes lights should also satisfy the wave-vector-matchingΔK=2KpKsKas=0, whereKp, Ks, Kasare the wave vectors of the pump light, the Stokes light, and the anti-Stokes signal, respectively [13,31–33]. CARS is a parametric phenomenon in which the input and output photons exchange energy, while the quantum state of the molecule is left unchanged. As a vibrationally enhanced four-wave mixing process, the anti-Stokes signal in CARS is generated at a new frequency that differs to the input beams, which is accompanied by a non-resonant signal caused by electronic contributions to the four-wave mixing process. So, for these photons, frequency-matching means energy conservation, and wave-vector-matching means momentum conservation. Moreover, for CARS microscopy, the wave-vector-matching conditions are relaxed because of the large cone of wave vectors and the short interaction length [11].

The phases of pump light, Stokes light and anti-Stokes signal are ωpt+Kp·r+ϕp, ωst+Ks·r+ϕs, and ωast+Kas·r+ϕas, where ϕp, ϕs, ϕas are the initial phases of the pump light, the Stokes light, and the anti-Stokes signal, respectively (t is the time coordinate, and r is the spatial vector). The relative phase Δϕbetween the pump light, Stokes light and anti-Stokes signal is

Δϕ=2(ωpt+Kp·r+ϕp)(ωst+Ks·r+ϕs)(ωast+Kas·r+ϕas)=Δωt+ΔK·r+2ϕpϕsϕas=2ϕpϕs,
where we ignore the initial phases (ϕas=0) of the anti-Stokes light. Because in CARS, the phase of the anti-Stokes field is only important when it is detected in an interferometric fashion. If the anti-Stokes field is applied as a local oscillator field, its phase can be tuned, and the transition between the ground state and the first vibrationally excited state can be controlled. This was explained elegantly by S. Rahav et al. in [34], where the CARS process with a phase-controlled local oscillator field was presented as an interference between two quantum pathways in the material. In the absence of the local oscillator field, the initial phase of anti-Stokes field is a constant value, which cannot be controlled by the pump and Stokes light (in order to simplify the physical model, the initial phase of the anti-Stokes light can be taken as zero). The relative phase of a circular level system is very important as it can impact some physical quantities of the system [35], such as the nonlinear susceptibility. Coincidentally, there are experiment and quantum theory studies about that THz wave intensity generated from a two-color laser-induced gas plasma can also be coherent controlled by the relative phase [36–39], which is the same as four-wave mixing phenomenon. Then by the same theory we can think that the relative phase of the CARS system can also influence the intensity of the CARS signal, and the relationship between the intensityIasand the relative phaseΔφis [40]

Iassin2Δϕ.

This formula demonstrates that the intensity of the anti-Stokes light Iascan be coherently controlled by the relative phase between the pump light and Stokes light Δϕ, which satisfies the sine curve relationship. It is noteworthy that the anti-Stokes light Iascan even be zero for the condition Δϕ=Nπ.

3. Analysis and discussion

Next, we present an experimental method to regulate and control the relative phase. A pair of wedge crystals are put along the light path, as shown in Fig. 1(a). When the pump light and Stokes light pass through the pair of crystals, the direction of the two-colored lights remains the same. Due to the angle of the wedge, we can also control the length of the vertical direction of the crystals to change the optical path length. However, the relative phase of the two-colored lights changes when the optical path length changes. Thus, we can control the length of the vertical direction of the crystals to alter the relative phase. Thenc, with the correct angle of the wedge and the appropriate dispersion we can obtain the relative phase Δϕ

Δϕ=2ϕpϕs=2π(2npΔlλpnsΔlλs)=2ωpnpωsnscΔl=tanθ2ωpnpωsnscΔh,
where θ is the angle of the wedge, c is the velocity of light in a vacuum, Δh and Δl are, respectively, the length changes in the vertical and horizontal directions of the crystals, λp and λs are, respectively, the wavelengths of the pump and Stokes lights in a vacuum, and np, ns are the refractive indexes of the pump and Stokes lights in the crystal, respectively. If we assume that Eq. (3) is equal to 2π, the rate of the regulation ρcan be obtained as

 figure: Fig. 1

Fig. 1 The sketch of the pair of wedge crystals (a) and the phase plate (b).

Download Full Size | PDF

ρ=nsΔlλs=λp/λs2np/nsλp/λs.

The rate of regulation is defined as ρ=nsΔl/λs, as shown in Eq. (4), which reflects the tenability of the relative phase Δϕ by a length changeΔl. The quantity ρ is larger, and this method is easier to be realized in experiments. If the rate of the regulation ρis too small, the length changeΔl maybe below the nanometer order of magnitude, which is hard to realize in experiments by stepping motor. As shown in Fig. 2(a), we have plotted a 3-dimensional image of Eq. (4) where the three quantities ρ, np/ns and λp/λsare the coordinate variables. From Fig. 2(a) it is found that the regulation rate ρis infinite on the 2-dimensional hyperplane 2np/nsλp/λs=0. In order to illustrate this more clearly, the top panel of Fig. 2(a) is replotted in Fig. 2(b), where the hyperplane instead appears as a line. In the CARS system, the wavelength of the Stokes light is longer than the pump light. In other words, we obtain the inequality condition λp/λs<1. For the normal dispersion condition, which has the relationship n(λ)=a+b/λ2+c/λ4, we also obtain another inequality condition, np/ns>1, implying we only need to study the quantities in the rectangular region, (i.e. the black dashed box shown in the top left corner of Fig. 2(b), where Fig. 2(c) is a 3-dimensional image of this region). The results show that if we want to obtain larger values of ρ, the quantities np/ns and λp/λs must respectively be smaller and larger. Furthermore, we obtained more intuitive curves of the variables ρand λp/λs with different values of np/ns, as shown in Fig. 2(d), where the parameter np/nswas chose as 1, 1.1, 1.2, 1.3, 1.4, shown respectively as the colors red, yellow, green, blue, and pink dotted.

 figure: Fig. 2

Fig. 2 2-dimensional and 3-dimensional images of the three physical quantities ρ, np/ns andλp/λs.

Download Full Size | PDF

For many kinds of media, the small difference between np and nscan be ignored because the frequency difference 2ωpωs in the CARS system is not zero. However, the optical delay Δτ of the pump and Stokes lights is written as

Δτ=(npns)Δlc=tanθ(npns)Δhc.

If we compare Eqs. (3) and (5), it is found that we can control the relative phase Δϕ without altering the optical delayΔτ, even in the air condition. This means that the intensity of the anti-Stokes light also depends on the light path length in air. This dependence may explain why it is not easy to obtain anti-Stokes signals in experiment.

Thus, in order to break the diffraction limit, we must first demonstrate an appropriate theoretical phase plate. The proposed phase plate has a phase shift of π/2 in the central area, as shown in Fig. 1(b). Do and Di are the outer and inner diameters of the phase plate, respectively. Through proper adjustment to the pair of wedge crystals, the relative phase in the outer area can be Nπ, and the inner area can be (N+0.5)π. Therefore, the intensity of the anti-Stokes light in the outer area is zero because of Eq. (2), and the effective area needed to generate an anti-Stokes signal can be reduced to just the inner area. An illustration of this physical mechanism is shown in Fig. 3, where Fig. 3(a) is the normal CARS system, and Fig. 3(b) is the phase plate system that requires a narrower area to generate an anti-Stokes signal. Figures 3(a) and 3(b) are just sketches, where the actual values of the x-axis and y-axis don't make sense. Hence, the units in these graphs are missing.

 figure: Fig. 3

Fig. 3 Sketches of the normal CARS system (a) and the super-resolution CARS system (b).

Download Full Size | PDF

Finally, in order to verify that our scheme is suitable for realizing super-resolution CARS microscopy, we considered a numerically generated test sample and simulated super-resolution images, as shown in Fig. 4. The sample consists of a dendritic line structure similar to structures found in, for example, brain tissue, as shown in Fig. 4(a), where the size is 1000 × 1000nm, and the line thickness is 6 nm. CARS is a third-order nonlinear optical phenomenon, and through Eq. (2) we can obtain the following equation

IasEas2ρ[sin2(Δϕ)+b][Ep2(r)Es(r)]2,
where ρis the density of the Raman-active molecule, b is the signal offset, Eas, Ep, Es are the electric field, where r denotes the radial component of the cylindrical coordinate system and reflects the characteristic of the light spot, of the anti-Stokes, pump, and Stokes lights, respectively. Here the Gaussian function is used to simplify the physical model and describe the spot shape, and the amplitude is assumed to be 1. It is noteworthy that they are not Gaussian beams, in which specific phase is needed and depends on spatial coordinates. Thus, the electric fieldsEpandEshave the expressions Ep(r)exp(r2/rp2)andEs(r)exp(r2/rs2), where rp and rs are the spot widths. Then, Eq. (6) can be expressed as
Iasρ[sin2(Δϕ)+b]exp(r2reff2),
reff=(4rp2+2rs2)12,
where reff is the effective width of the CARS course. However, areas that are larger than the effective width reffcan also generate a CARS signal, and so we assumed the radius of the area, which can effectively generate a CARS signal, to be ro=reff/0.8=160nm(i.e.rp=rs=313.53nm). In other words, no CARS signal is generated in an area whose radius is larger than ro. The outer diameter of the phase plate Do is larger than 2ro=320nm.

 figure: Fig. 4

Fig. 4 (a) Test sample simulating a fine structure (e.g. brain tissue). By simulating the super-resolution CARS imaging course of the test sample, the CARS images are obtained for phase plate inner diameters Di of 320nm (b), 160nm (c), 40nm (d) and 20nm (e), where the signal offset b is zero. The remaining images are for signal offsets b of 1 (f), 0.8 (g), 0.4 (h) and 0.1 (i), where the inner diameterDiis 20nm.

Download Full Size | PDF

The sub-micrometric spot widths are obtained by focusing the beams with a high numerical aperture (and consequently low working distance) objective. If the phase plate is placed before the objective, we should consider the amplitude and phase modification induced by the objective [41,42]. Therefore, the phase plate may be some nano-structure films after the objective. However, this is a theoretical study, and then we simplify the physical model to better reflect the physical connotation without concerning about the objective. With these arguments, we simulated CARS super-resolution images with Matlab, as shown in Figs. 4(b)-4(i), where the colors from blue to red correspond to the intensity of the anti-Stokes light from weak to strong. The Figs. 4(b)-4(e) correspond to phase plate inner diameters Diof 320 nm, 160 nm, 40 nm and 20 nm, where the signal offset b was zero. When comparing these imaging results, it is found that images with smaller inner diameters display more detail. However, it is more difficult to fabricate smaller inner diameter phase plates.

Next, we studied the influence of the signal offset b on the simulated images, as shown in Figs. 4(f)-4(i), which correspond to signal offsets of b = 1, 0.8, 0.4 and 0.1, where the inner diameter Di was chosen as 20 nm. Signal offset b is the noise signal offset, which depends on precision and stability of the experiment system. From these results, it is found that with a small noise signal offset we can still obtain super-resolution images.

4. Conclusions

In summary, we presented a theoretical study on the suppression of CARS by the relative phase between pump light and Stokes light and its application to super-resolution imaging. Here the relative phase of the pump and Stokes lights was defined asΔϕ=2ϕpϕs. The relationship between the anti-Stokes signal and the relative phase was derived asIassin2Δϕ. Subsequently, we studied the possibility of regulating and controlling the relative phase by a pair of wedge crystals. In this approach, we proposed a special phase plate to reduce the effective region of the generation of the CARS signal and simulated super-resolution images. We note however, that in practical terms, this kind of phase plate may not be fabricated very easily, though it may be realized via metamaterials or the other nano-optics technology.

Funding

National Natural Science Foundation of China (NSFC) (11775147); Science and Technology Program of Shenzhen (STPS) (JCYJ20170302153912966, JCYJ20160608173121055); Scientific Research Start-up Project for Newly Introduced Teacher of Shenzhen University (2017015).

Acknowledgments

Thanks for the help of professor Junle Qu and professor Jun Song.

References

1. G. T. Boyd, Z. H. Yu, and Y. R. Shen, “Photoinduced luminescence from the noble metals and its enhancement on roughened surfaces,” Phys. Rev. B Condens. Matter 33(12), 7923–7936 (1986). [CrossRef]   [PubMed]  

2. W. Kaiser and C. G. B. Garrett, “Two-photon excitation in CaF2:Eu2+,” Phys. Rev. Lett. 7(6), 229–231 (1961). [CrossRef]  

3. P. A. Franken, A. E. Hill, C. W. Peters, and G. Weinreich, “Generation of optical harmonics,” Phys. Rev. Lett. 7(4), 118–119 (1961). [CrossRef]  

4. R. Hellwarth and P. Christensen, “Nonlinear optical microscopic examination of structure in polycrystalline ZnSe,” Opt. Commun. 12(3), 318–322 (1974). [CrossRef]  

5. R. W. Terhune, P. D. Maker, and C. M. Savage, “Optical harmonic generation in calcite,” Phys. Rev. Lett. 8(10), 404–406 (1962). [CrossRef]  

6. P. D. Maker and R. W. Terhune, “Study of optical effects due to an induced polarization third order in the electric field strength,” Phys. Rev. 137(3A), A801–A818 (1965). [CrossRef]  

7. M. Müller, J. Squier, K. R. Wilson, and G. J. Brakenhoff, “3D microscopy of transparent objects using third-harmonic generation,” J. Microsc. 191(3), 266–274 (1998). [CrossRef]   [PubMed]  

8. E. Ploetz, S. Laimgruber, S. Berner, W. Zinth, and P. Gilch, “Femtosecond stimulated Raman microscopy,” Appl. Phys. B 87(3), 389–393 (2007). [CrossRef]  

9. C. W. Freudiger, W. Min, B. G. Saar, S. Lu, G. R. Holtom, C. He, J. C. Tsai, J. X. Kang, and X. S. Xie, “Label-free biomedical imaging with high sensitivity by stimulated Raman scattering microscopy,” Science 322(5909), 1857–1861 (2008). [CrossRef]   [PubMed]  

10. B. G. Saar, C. W. Freudiger, J. Reichman, C. M. Stanley, G. R. Holtom, and X. S. Xie, “Video-rate molecular imaging in vivo with stimulated Raman scattering,” Science 330(6009), 1368–1370 (2010). [CrossRef]   [PubMed]  

11. A. Zumbusch, G. R. Holtom, and X. S. Xie, “Three-dimensional vibrational imaging by coherent anti-Stokes Raman scattering,” Phys. Rev. Lett. 82(20), 4142–4145 (1999). [CrossRef]  

12. 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]   [PubMed]  

13. 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]   [PubMed]  

14. 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]  

15. C. Cleff, A. Gasecka, P. Ferrand, H. Rigneault, S. Brasselet, and J. Duboisset, “Direct imaging of molecular symmetry by coherent anti-stokes Raman scattering,” Nat. Commun. 7(1), 11562 (2016). [CrossRef]   [PubMed]  

16. K. M. Hajek, B. Littleton, D. Turk, T. J. McIntyre, and H. Rubinsztein-Dunlop, “A method for achieving super-resolved widefield CARS microscopy,” Opt. Express 18(18), 19263–19272 (2010). [CrossRef]   [PubMed]  

17. E. Betzig, G. H. Patterson, R. Sougrat, O. W. Lindwasser, S. Olenych, J. S. Bonifacino, M. W. Davidson, J. Lippincott-Schwartz, and H. F. Hess, “Imaging intracellular fluorescent proteins at nanometer resolution,” Science 313(5793), 1642–1645 (2006). [CrossRef]   [PubMed]  

18. S. W. Hell and J. Wichmann, “Breaking the diffraction resolution limit by stimulated emission: stimulated-emission-depletion fluorescence microscopy,” Opt. Lett. 19(11), 780–782 (1994). [CrossRef]   [PubMed]  

19. M. J. Rust, M. Bates, and X. Zhuang, “Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (STORM),” Nat. Methods 3(10), 793–795 (2006). [CrossRef]   [PubMed]  

20. W. P. Beeker, P. Gross, C. J. Lee, C. Cleff, H. L. Offerhaus, C. Fallnich, J. L. Herek, and K. J. Boller, “A route to sub-diffraction-limited CARS Microscopy,” Opt. Express 17(25), 22632–22638 (2009). [CrossRef]   [PubMed]  

21. W. P. Beeker, C. J. Lee, K. J. Boller, P. Groß, C. Cleff, C. Fallnich, H. L. Offerhaus, and J. L. Herek, “Spatially dependent Rabi oscillations: an approach to sub-diffraction-limited coherent anti-Stokes Raman-scattering microscopy,” Phys. Rev. A 81(1), 012507 (2010). [CrossRef]  

22. K. M. Hajek, B. Littleton, D. Turk, T. J. McIntyre, and H. Rubinsztein-Dunlop, “A method for achieving super-resolved widefield CARS microscopy,” Opt. Express 18(18), 19263–19272 (2010). [CrossRef]   [PubMed]  

23. V. Raghunathan and E. O. Potma, “Multiplicative and subtractive focal volume engineering in coherent Raman microscopy,” J. Opt. Soc. Am. A 27(11), 2365–2374 (2010). [CrossRef]   [PubMed]  

24. W. P. Beeker, C. J. Lee, K. J. Boller, P. Groß, C. Cleff, C. Fallnich, H. L. Offerhaus, and J. L. Herek, “A theoretical investigation of super-resolution CARS imaging via coherent and incoherent saturation of transitions,” J. Raman Spectrosc. 42(10), 1854–1858 (2011). [CrossRef]  

25. W. Liu and H. Niu, “Diffraction barrier breakthrough in coherent anti-Stokes Raman scattering microscopy by additional probe-beam-induced phonon depletion,” Phys. Rev. A 83(2), 023830 (2011). [CrossRef]  

26. C. Cleff, P. Gross, C. Fallnich, H. L. Offerhaus, J. L. Herek, K. Kruse, W. P. Beeker, C. J. Lee, and K. J. Boller, “Ground-state depletion for subdiffraction-limited spatial resolution in coherent anti-Stokes Raman scattering microscopy,” Phys. Rev. A 86(2), 023825 (2012). [CrossRef]  

27. C. Cleff, P. Gross, C. Fallnich, H. L. Offerhaus, J. L. Herek, K. Kruse, W. P. Beeker, C. J. Lee, and K. J. Boller, “Stimulated-emission pumping enabling sub-diffraction-limited spatial resolution in coherent anti-Stokes Raman scattering microscopy,” Phys. Rev. A 87(3), 033830 (2013). [CrossRef]  

28. D. Wang, S. Liu, Y. Chen, J. Song, W. Liu, M. Xiong, G. Wang, X. Peng, and J. Qu, “Breaking the diffraction barrier using coherent anti-Stokes Raman scattering difference microscopy,” Opt. Express 25(9), 10276–10286 (2017). [CrossRef]   [PubMed]  

29. L. Gong and H. Wang, “Breaking the diffraction limit by saturation in stimulated-Raman-scattering microscopy: A theoretical study,” Phys. Rev. A 90(1), 013818 (2014). [CrossRef]  

30. L. Gong and H. Wang, “Suppression of stimulated Raman scattering by an electromagnetically-induced-transparency–like scheme and its application for super-resolution microscopy,” Phys. Rev. A 92(2), 023828 (2015). [CrossRef]  

31. J.-X. Cheng and X. S. Xie, Coherent Raman Scattering Microscopy, 1st ed. (CRC Press, 2012).

32. M. Muller and J. Squier, C. A. De Lange and G. J. Brakenhoff, “CARS microscopy with folded BoxCARS phasematching,” J. Microsc. 197(Pt 2), 150–158 (2000). [CrossRef]   [PubMed]  

33. B. N. Toleutaev, T. Tahara, and H. Hamaguchi, “Broadband (1000 cm−1) multiplex CARS spectroscopy: application to polarization sensitive and time-resolved measurements,” Appl. Phys. B 59(4), 369–375 (1994). [CrossRef]  

34. S. Rahav and S. Mukamel, “Stimulated coherent anti-Stokes Raman spectroscopy (CARS) resonances originate from double-slit interference of two-photon Stokes pathways,” Proc. Natl. Acad. Sci. U.S.A. 107(11), 4825–4829 (2010). [CrossRef]   [PubMed]  

35. D. Wang and J. Liu, “Polarization property of the THz wave generated from a two-color laser-induced gas plasma,” Phys. Rev. A 86(2), 023833 (2012). [CrossRef]  

36. J. Dai, N. Karpowicz, and X.-C. Zhang, “Coherent polarization control of terahertz waves generated from two-color laser-induced gas plasma,” Phys. Rev. Lett. 103(2), 023001 (2009). [CrossRef]   [PubMed]  

37. H. Wen and A. M. Lindenberg, “Coherent terahertz polarization control through manipulation of electron trajectories,” Phys. Rev. Lett. 103(2), 023902 (2009). [CrossRef]   [PubMed]  

38. N. Karpowicz and X.-C. Zhang, “Coherent terahertz echo of tunnel ionization in gases,” Phys. Rev. Lett. 102(9), 093001 (2009). [CrossRef]   [PubMed]  

39. Z. Chu, J. Liu, K. Wang, and J. Yao, “Four-wave mixing model solutions for polarization control of terahertz pulse generated by a two-color laser field in air,” Chin. Opt. Lett. 8(7), 697–700 (2010). [CrossRef]  

40. C. Dorrer, N. Belabas, J.-P. Likforman, and M. Joffre, “Spectral resolution and sampling issues in Fourier-transform spectral interferometry,” J. Opt. Soc. Am. B 17(10), 1795–1802 (2000). [CrossRef]  

41. X. Gao, F. Gan, and W. Xu, “Superresolution by three-zone pure phase plate with 0, π, 0 phase variation,” Opt. Laser Technol. 39(5), 1074–1080 (2007). [CrossRef]  

42. R. Beams, J. W. Woodcock, J. W. Gilman, and S. J. Stranick, “Phase mask-based multimodal superresolution microscopy,” Photonics 4(3), 39 (2017). [CrossRef]   [PubMed]  

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 (4)

Fig. 1
Fig. 1 The sketch of the pair of wedge crystals (a) and the phase plate (b).
Fig. 2
Fig. 2 2-dimensional and 3-dimensional images of the three physical quantities ρ, n p / n s and λ p / λ s .
Fig. 3
Fig. 3 Sketches of the normal CARS system (a) and the super-resolution CARS system (b).
Fig. 4
Fig. 4 (a) Test sample simulating a fine structure (e.g. brain tissue). By simulating the super-resolution CARS imaging course of the test sample, the CARS images are obtained for phase plate inner diameters D i of 320nm (b), 160nm (c), 40nm (d) and 20nm (e), where the signal offset b is zero. The remaining images are for signal offsets b of 1 (f), 0.8 (g), 0.4 (h) and 0.1 (i), where the inner diameter D i is 20nm.

Equations (8)

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

Δϕ=2( ω p t+ K p · r + ϕ p )( ω s t+ K s · r + ϕ s )( ω as t+ K as · r + ϕ as ) =Δωt+Δ K · r +2 ϕ p ϕ s ϕ as =2 ϕ p ϕ s ,
I as sin 2 Δϕ.
Δϕ=2 ϕ p ϕ s =2π( 2 n p Δl λ p n s Δl λ s )= 2 ω p n p ω s n s c Δl=tanθ 2 ω p n p ω s n s c Δh,
ρ= n s Δl λ s = λ p / λ s 2 n p / n s λ p / λ s .
Δτ=( n p n s ) Δl c =tanθ( n p n s ) Δh c .
I as E as 2 ρ[ sin 2 ( Δϕ )+b ] [ E p 2 ( r ) E s ( r ) ] 2 ,
I as ρ[ sin 2 ( Δϕ )+b ]exp( r 2 r eff 2 ),
r eff = ( 4 r p 2 + 2 r s 2 ) 1 2 ,
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.