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

Breaking the diffraction barrier using coherent anti-Stokes Raman scattering difference microscopy

Open Access Open Access

Abstract

We propose a method to improve the resolution of coherent anti-Stokes Raman scattering microscopy (CARS), and present a theoretical model. The proposed method, coherent anti-Stokes Raman scattering difference microscopy (CARS-D), is based on the intensity difference between two differently acquired images. One being the conventional CARS image, and the other obtained when the sample is illuminated by a doughnut shaped spot. The final super-resolution CARS-D image is constructed by intensity subtraction of these two images. However, there is a subtractive factor between them, and the theoretical model sets this factor to obtain the best imaging effect.

© 2017 Optical Society of America

1. Introduction

Far-field optical microscopy has rapidly developed recently, 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]. However, the diffraction barrier in the imaging system limits the lateral resolution of conventional optical microscopes to approximately half the illumination wavelength. Thus, breaking the diffraction barrier is an intense research topic. During the past two decades several methods of breaking the diffraction limit have been proposed, including stochastic optical reconstruction microscopy (STORM) [16], photoactivated localization microscopy (PALM) [17], stimulated emission depletion microscopy (STED) [18], structured illumination microscopy (SIM) [19,20], digital image processing [21,22], and total internal reflection microscopy based on a nanostructured substrate [23]. STORM, PALM, STED, and SIM are the most widely applied, and have all been demonstrated to achieve super-resolution in bioscience [24–27]. However, significant disadvantages of these super-resolution methods include: photobleaching, toxicity, and the effects of fluorophores on labeled molecules in live samples. Thus, nondestructive detection cannot be achieved for these methods, but nondestructive detection is important for in vivo experiments in life sciences.

CARS is one of the most important microscopy methods, and can realize nondestructive detection. However, it is difficult to improve spatial resolution for CARS systems, and CARS super-resolution has been an active research topic [28–35]. Fluorescence emission difference (FED) microscopy has been recently proposed for breaking the diffraction barrier, which is based on the intensity difference between two differently acquired images [36]. This paper employs a similar concept and presents a theoretical investigation of how this kind of difference method in CARS microscopy can break the diffraction limit. We have called the method coherent anti-Stokes Raman scattering difference (CARS-D) microscopy. We build a theoretical CARS-D model and analyze related parameters settings to obtain the best super-resolution effect. Comparing with existing methods, the present CARS-D method can achieve non-invasive super-resolution imaging in situ, in real-time and without labelling.

2. Principle

CARS microscopy is a third order nonlinear optical imaging method in which two laser pulses at different frequencies, pump light (ωp) and Stokes light (ωs), are focused onto a sample (the probe light is also ωp). Raman resonance occurs when ωp-ωs matches a Raman active molecular vibrational level. CARS detects the output anti-Stokes signal (ωas) at a new frequency, 2ωp-ωs, which is enhanced by the Raman resonance. The anti-Stokes signal intensity can be expressed as

IasEas2[ρEp2(r)Es(r)]2,
where ρ is the density of the Raman active molecule, and Eas(r), Ep(r), Es(r) are the electric fields, 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.

Figure 1 illustrates the CARS-D method. Two different scanned images are processed: one being the conventional CARS image acquired when the sample is illuminated by a Gaussian shaped spot, and the other is obtained with the sample is illuminated by a doughnut shaped spot generated by modulating the illumination beam with a vortex 0–2π phase plate. The final super-resolution CARS-D image is constructed by intensity subtraction of these two images. The green lines and arrow in Fig. 1 represent the conventional CARS course, where Figs. 1(a) and 1(b) show the electric field distribution of Gaussian shaped Stokes (Es1) and pump (Ep) lights, respectively, and Fig. 1(d) shows the Gaussian shaped effective excitation area of anti-Stokes (Ias1) light. The yellow lines and arrow represent the proposed CARS-D course, where Fig. 1(c) shows the electric field distribution of doughnut shaped Stokes (Es2), and Fig. 1(e) shows the doughnut shaped effective excitation area of anti-Stokes (Ias2) light. The red lines and arrow represent the difference of the two courses, and Fig. 1(f) shows the effective excitation area of the CARS-D signal is smaller and sharper, although there is some negative area. Point by point scanning provides a higher resolution image by the excitation spot difference as shown in Fig. 1(f). In Section 3 we build a physical model to further analyze the impact of the relevant parameters on imaging results.

 figure: Fig. 1

Fig. 1 Illustration of CARS-D method.

Download Full Size | PDF

3. Theoretical model

CARS is a third order nonlinear optical phenomenon, and based on Eq. (1) with some reasonable assumption we can obtain the following analytical model.

The Gaussian shaped pump (Ep) and Stokes (Es1) lights can be expressed as

Ep(r)=Ep0exp(r2rp2),
and
Es1(r)=Es10exp(r2rs12),
where Ep0 and Es10 are the amplitudes, and rp and rs1 are the spot radii of the electric field of pump and Stokes lights, respectively. To simplify the model to more directly reflect the physical mechanism, the doughnut shaped Stokes (Es2) can be expressed as [18,37]
Es2(r)=Es20rexp(r2rs22),
where Es20 and rs2 are the amplitude and radius, respectively, of the electric field of the doughnut shaped Stokes light.

From Eqs. (1)–(3) the conventional and unconventional CARS courses with effective excitation areas of anti-Stokes (Ias1 and Ias2) can be expressed as

Ias1[ρEp02exp(2r2rp2)Es1exp(r2rs12)]2,
and

Ias2[ρEp02exp(2r2rp2)Es2rexp(r2rs22)]2.

The difference of Eqs. (5) and (6) is the CARS-D course,

ΔIas=Ias1kMIas2,
where kM is the subtractive factor. To simplify the model, we assume the radius of the pump light, rp = 0.5λ (λ is the wavelength of the pump light), and the relationships between the radii rp, rs1, and rs2 are assumed to be rs1 = k1rp and rs2 = k2rp (k1 and k2 are ratio coefficients). The amplitudes Ep0 and Es10 are determined by the normalization condition and generally taken as 1, while the amplitude Es20 is replaced by M. Then, from Eqs. (5)–(7),

ΔIasρ2exp(16r2λ2){exp[8r2(k1λ)2]kMM2r2exp[8r2(k2λ)2]}.

We need to determine the factor M. There is some negative area using the CARS-D method, and the desired signal may be covered by the negative signal when the negative value is too large. Therefore, we define M as the specific value between the positive and negative signal when the subtractive factor kM = 1 and ρ = 1 everywhere over the area. That is to say M equals the ratio between the volume integrals of Ias1 and Ias2. The volume integrals of Ias1 and Ias2 are, respectively,

V1=0rdr02πdθexp(16r2λ2)exp[8r2(k1λ)2]=πa1,
a1=16λ2+8(k1λ)2,
and
V2=0rdr02πdθexp(16r2λ2)(r2)exp[8r2(k2λ)2]=πa22,
a2=16λ2+8(k2λ)2,
and, hence,

M=V1V2=a22a1.

Note that all results given in Eqs. (9)-(13) have been transformed to no-dimensional form. The integration of the formula (8) will be larger than zero when there exists noise in the process. Instead, if the volume integral is smaller than zero, it means that some useful information in these images is lost. Therefore, the optimal integral should be equal to zero where ρ = kM = 1 to achieve a perfect super-resolution image without noise or information loss. In order to obtain nice images around the optimal point, we need to adjust the factor kM around 1 in the following studies, and define M (shown in Eq. (13) for facilitate analysis.

4. Numerical results

In this section, we present numerical results for images based on conventional CARS and the present CARS-D method by Matlab simulation according to above-mentioned analytical equations (see Eqs. (1)-(13)). To verify our numerical model we have carried out imaging experiments based on a commercial CARS imaging system. The experimental samples are polystyrene beads with the 110 nm diameter. Figure 2(a) is the TEM (Transmission Electron Microscope) image of the polystyrene beads with the size 6.25 μm × 6.25μm. Based on the CARS imaging system, Fig. 2(b) gives the image of one polystyrene bead marked with the green dotted circle in Fig. 2(a), where the brighter color corresponds to the stronger CARS signal (Ias). The wavelengths of pump and Stokes lights are 655 nm and 820 nm, respectively, while the Raman shift of the polystyrene bead is 1600 cm−1. Figure 2(c) is the intensity distribution of Ias along the green dotted line in the Fig. 2(b), where the position along the line is the abscissa. We can obtain the value of the spatial resolution as 484 nm for the current CARS system.

 figure: Fig. 2

Fig. 2 (a) TEM image of the polystyrene beads with the size 6.25 μm×6.25 μm; (b) Experimental CARS image of one polystyrene bead marked with the green circle in Fig. 2 (a); (c) Intensity distribution of Ias along the green line in Fig. 2 (b); (d) Theoretical CARS image of Fig. 2 (a); (e) Intensity distribution of Ias along the green line in Fig. 2 (d); (f) Local enlarged curve marked with the green circle in Fig. 2 (e).

Download Full Size | PDF

Then we apply our numerical model to image the same sample given in Fig. 2(a), and all parameters in the simulation results from above experimental data. The diameter of the pump light spot is 655 nm, the coefficient k1 is 820/655, and kM is 1. The corresponding calculated CARS image is shown in Fig. 2(d). The intensity distribution of Ias along the line in Fig. 2(d) is given in Fig. 2(e), and a local enlarged curve marked with the green circle in Fig. 2 (e) can be found in Fig. 2(f). From the numerical result, we can learn the theoretical value of the spatial resolution is 480nm, which agrees well with the previous experimental value (see Fig. 2(c)).

Furthermore, we also obtain the theoretical images based on the doughnut CARS image (Fig. 3(a)) and the CARS-D image (Fig. 3(b)), respectively, with the coefficient k2 = 820/655. Figure 3(c) is the intensity distribution of Ias along the green dotted line in Fig. 3(b), and Fig. 3(d) is the local enlarged curve marked with the green circle in Fig. 3 (c). We can find that CARS-D method has better spatial resolution (338 nm) than conventional CARS method.

 figure: Fig. 3

Fig. 3 (a) Doughnut CARS image of the polystyrene beads; (b) CARS-D image of the polystyrene beads; (c) Intensity distribution of Ias along the green line in Fig. 3(b); (d) Local enlarged curve marked with the green circle in Fig. 3 (e).

Download Full Size | PDF

Next, we further accurately estimate the performance of the proposed CARS-D method and optimize its performance to achieve super-resolution images. For the particular case k1 = k2 = kM = 1, we obtained the effective excitation areas of conventional CARS signal (Ias1), doughnut CARS signal (Ias2), and CARS-D signal (ΔIas), as shown in Fig. 4(a), corresponding to the red, blue, and green dotted lines, respectively. To estimate the attainable spatial resolution, we performed the simulation test on a spoke-like sample [36], as shown in Fig. 4(b) with size 18λ×18λ/π. In the process of our theory analysis, results are closely related to several key parameters: the wavelength of the pump light λ, and the radii of the Stokes light spot determined by the factors k1 and k2. Therefore, both the size of the test sample and the pixels of the sample image are set to be related to λ. Then, in the following analysis the value of wavelength λ will be not especially emphasized anymore.

 figure: Fig. 4

Fig. 4 (a) Illustration of the effective excitation areas of conventional CARS signal (red dotted line), doughnut CARS signal (blue dotted line) and CARS-D signal (green dotted line). (b) A spoke-like test sample.

Download Full Size | PDF

The spoke-like sample is widely used because the resolution of an imaging technique can be assessed very simply by measuring the radius of a circle delimiting the ‘unresolved’ central area and the ‘resolved’ peripheral area in the corresponding image. Figure 5 shows the imaging results using an excitation spot to scan the sample point by point. Figures 5(a), (b), and (c) correspond to conventional CARS, doughnut CARS, and CARS-D images, respectively. From the Fig. 5(a), one also sees that the patterns beyond the black circle are the same as those shown in Fig. 4(b), but the patterns have such obvious fuzzy characteristic that the spoke-like lines cannot be distinguished from Fig. 5(a) obtained by the conventional CARS method due to the effect of diffraction limit. Similar patterns can also be seen in Fig. 3(b) with the only difference of the irradiation light (i.e., a doughnut shaped spot). Figure 5(c) shows the CARS-D signal (ΔIas) with a constraint that a point is set to zero when the difference is negative. Comparing Figs. 5(a) and 5(c), while features inside the region indicated by the black circle cannot be clearly discerned in the conventional CARS image, some features in the same region of the CARS-D image can be distinguished, which indicates enhanced resolution.

 figure: Fig. 5

Fig. 5 (a) Conventional CARS image. (b) Doughnut CARS image. (c) CARS-D image.

Download Full Size | PDF

Figures 6 and 7 (a)–(f) show the effect of kM from 0.8 to 1.3, step 0.1, for CARS-D images, respectively. The images are colored following the parula colormap. Figure 5 shows that when kM = 1, the features in the middle region can be distinguished clearly. However, as kM decreases, the features in the middle region cannot be as well discerned, due to increased noise in the middle region with smaller kM. When kM>1, features in the middle region are also eliminated, and this effect increases with increasing kM. To display this problem more intuitively, Fig. 7 shows the same images colored following the lines colormap. There are two regions in Figs. 7(a)–5(f), indicated by the inner and outer black circles. Within the inner black circle, where there is full noise, features cannot be completely discerned. Outside the outer black circle, features are discernible with little or no noise. Tuning kM shows that features between the inner and outer black circles are most clearly distinguished for kM = 1. Combined with Eqs. (8)–(13), Fig. 7 shows that to obtain the best CARS-D resolution, the volume integral of Eq. (8) must be zero for ρ = 1. For kM>1, the volume integral of Eq. (8) is negative, which means that some of the information will be lost in the process of the difference. However for kM<1, the volume integral of Eq. (8) is positive, which means that there exists a certain amount of noise between inner and outer black circles in the CARS-D method. Therefore, only for kM = 1, we can expect that neither the noise or information loss can be found, i.e., a perfect super-resolution image can be achieved.

 figure: Fig. 6

Fig. 6 CARS-D images with different subtractive factors using the parula color-map.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 CARS-D images with different subtractive factors using the lines color-map.

Download Full Size | PDF

Given the black circles in Fig. 7, we can define inner and outer spatial resolutions (SRin and SRout, respectively) to estimate CARS-D imaging effects. Each spatial resolution (SR) can be calculated by the diameter, D, of the black circle, i.e., SR = πD/36 (there are 36 lines in the spoke-like sample as shown in Fig. 4(b)). Then we may analyze the impact of the factors k1 and k2 on CARS-D spatial resolution. Figure 8 shows SRin and SRout as functions of k1 and k2, keeping one factor constant while varying the other. The spatial resolution reduces as the factor reduces, and tends toward a constant when the factor is large enough. The results suggested as long as the radius of the pump light spot is a determined value, the radii of the Gaussian shaped and doughnut shaped Stokes spots have little influence on the spatial resolution, unless the radii are negligible values. However k2 has a greater effect on SR than k1. Maximum SRout≈0.35λ and SRin≈0.065λ. However, it is unreasonable to take the spatial resolution of CARS-D as 0.065λ, because that not all the features between the inner and outer black circles can be distinguished. Therefore, we can only regard the spatial resolution SR of CARS-D as between 0.35λ and 0.065λ, and probably closer to 0.065λ.

 figure: Fig. 8

Fig. 8 Spatial resolution as function of factors k1 and k2: (a) inner spatial resolution, SRin, as a function of k1 with k2 = kM = 1; (b) SRin as a function of k2 with k1 = kM = 1; (c) outer spatial resolution, SRout, as a function of k1 with k2 = kM = 1; (d) SRout as a function of k2 with k1 = kM = 1.

Download Full Size | PDF

As above all simulations are based on the spoke-like test sample as shown in Fig. 4(b), we will further evaluate the reliability of the CARS-D method by using an irregular sample. To illustrate that the CARS-D method provides superior imaging than conventional CARS based on irregular sample, we used the Golgi complex image as an irregular test sample, as shown in Fig. 9. Figure 9(a) shows a visual image of the Golgi complex, and the size is 9λ × 9λ. Note that the size is not the real size of the Golgi complex in bioscience, but an irregular test sample to highlight details. Figures 9(b), (c) and (d) show conventional CARS doughnut CARS and CARS-D images, respectively, which are numerical simulations for the special case with k1 = k2 = kM = 1. There is considerable resolution improvement when this kind of irregular sample was imaged by CARS-D method. Many features disclosed by CARS-D imaging are completely hidden under conventional CARS and doughnut CARS observation. Therefore, the present CARS-D method can effectively be used for super-resolution imaging not only based on the special regular samples (e.g., spoke-like samples) but also on any irregular samples.

 figure: Fig. 9

Fig. 9 (a) Golgi complex visual image (as an irregular test sample), (b) conventional CARS image, (c) doughnut CARS image, and (d) CARS-D image.

Download Full Size | PDF

5. Conclusion

We proposed a CARS-D method to improve the resolution of CARS microscopy. On the basis of intensity differences between two differently acquired images, CARS-D can achieve spatial resolution between 0.35λ and 0.065λ, which is beyond the conventional diffraction barrier that conventional CARS microscopy cannot break. In the proposed CARS-D theoretical model, the best resolution is produced when the volume integral of Eq. (8) is zero for ρ = 1. Based on numerical simulation, we verify the CARS-D can be effectively applied for super-resolution microscopy.

Funding

National Natural Science Foundation of China (NSFC) (61235012, 61620106016, 61525503, 61378091, 61405123); National Basic Research Program of China (Program 973) (2015CB352005); Guangdong Natural Science Foundation Innovation Team (2014A030312008); Hong Kong, Macao and Taiwan cooperation innovation platform & major projects of international cooperation in Colleges and Universities in Guangdong Province (2015KGJHZ002); and Shenzhen Basic Research Project (JCYJ20160328144746940, JCYJ20160308093035903, JCYJ20150930104948169).

References and links

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. A. Pliss, X. Peng, L. Liu, A. Kuzmin, Y. Wang, J. Qu, Y. Li, and P. N. Prasad, “Single cell assay for molecular diagnostics and medicine: monitoring intracellular concentrations of macromolecules by two-photon fluorescence lifetime imaging,” Theranostics 5(9), 919–930 (2015). [CrossRef]   [PubMed]  

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. Y. Shao, H. Liu, W. Qin, J. Qu, X. Peng, H. Niu, and B. Z. Gao, “Addressable, large-field second harmonic generation microscopy based on 2D acousto-optical deflector and spatial light modulator,” Appl. Phys. B 108(4), 713–716 (2012). [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. J. Song, J. H. Xian, H. B. Niu, and J. L. Qu, “Significantly enhanced third harmonic generation using individual Au nanorods coated with gain materials,” IEEE Photonics J. 7(4), 4500909 (2015). [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. 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]  

9. Y. F. Yuan, Y. N. Lin, B. B. Gu, N. Panwar, S. C. Tjin, J. Song, J. L. Qu, and K. T. Yong, “Optical trapping-assisted SERS platform for chemical and biosensing applications: Design perspectives,” Coord. Chem. Rev. 339, 138–152 (2017). [CrossRef]  

10. Y. F. Yuan, N. Panwar, S. H. K. Yap, Q. Wu, S. W. Zeng, J. H. Xu, S. C. Tjin, J. Song, J. L. Qu, and K. T. Yong, “SERS-based ultrasensitive sensing platform: An insight into design and practical applications,” Coord. Chem. Rev. 337, 1–33 (2017). [CrossRef]  

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. S. L. Liu, W. Liu, D. N. Chen, J. L. Qu, and H. B. Niu, “Research on coherent anti-Stokes Raman scattering microscopy,” ACTA Phys. 65, 064204 (2016).

14. S. W. Zhang, D. N. Chen, S. L. Liu, W. Liu, and H. B. Niu, “Nanometer resolution coherent anti-Stokes Raman scattering microscopic imaging,” ACTA Phys. 64, 223301 (2015).

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, 11562 (2016). [CrossRef]   [PubMed]  

16. 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]  

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. G. L. Gustafsson, “Surpassing the lateral resolution limit by a factor of two using structured illumination microscopy,” J. Microsc. 198(2), 82–87 (2000). [CrossRef]   [PubMed]  

20. R. Heintzmann and C. Cremer, “Laterally modulated excitation microscopy: Improvement of resolution by using a diffraction grating,” Proc. SPIE 3568, 185–196 (1999). [CrossRef]  

21. W. A. Carrington, R. M. Lynch, E. D. W. Moore, G. Isenberg, K. E. Fogarty, and F. S. Fay, “Superresolution three-dimensional images of fluorescence in cells with minimal light exposure,” Science 268(5216), 1483–1487 (1995). [CrossRef]   [PubMed]  

22. H. Kano, H. T. M. Voort, M. Schrader, G. M. P. Kempen, and S. W. Hell, “Avalanche photodiode detection with object scanning and image restoration provides 2-4 fold resolution increase in two-photon fluorescence microscopy,” Bioimaging 4(3), 187–197 (1996). [CrossRef]  

23. A. Sentenac, K. Belkebir, H. Giovannini, and P. C. Chaumet, “High-resolution total-internal-reflection fluorescence microscopy using periodically nanostructured glass slides,” J. Opt. Soc. Am. A 26(12), 2550–2557 (2009). [CrossRef]   [PubMed]  

24. Z. Yang, A. Sharma, J. Qi, X. Peng, D. Y. Lee, R. Hu, D. Lin, J. Qu, and J. S. Kim, “Super-resolution fluorescent materials: an insight into design and bioimaging applications,” Chem. Soc. Rev. 45(17), 4651–4667 (2016). [CrossRef]   [PubMed]  

25. V. Westphal and S. W. Hell, “Nanoscale resolution in the focal plane of an optical microscope,” Phys. Rev. Lett. 94(14), 143903 (2005). [CrossRef]   [PubMed]  

26. K. Xu, H. P. Babcock, and X. Zhuang, “Dual-objective STORM reveals three-dimensional filament organization in the actin cytoskeleton,” Nat. Methods 9(2), 185–188 (2012). [CrossRef]   [PubMed]  

27. A. Brodehl, P. N. Hedde, M. Dieding, A. Fatima, V. Walhorn, S. Gayda, T. Šarić, B. Klauke, J. Gummert, D. Anselmetti, M. Heilemann, G. U. Nienhaus, and H. Milting, “Dual color photoactivation localization microscopy of cardiomyopathy-associated desmin mutants,” J. Biol. Chem. 287(19), 16047–16057 (2012). [CrossRef]   [PubMed]  

28. 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]  

29. 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]  

30. 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]  

31. 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]  

32. 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]  

33. 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), 4795–4804 (2011). [CrossRef]  

34. 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), 5946–5951 (2012). [CrossRef]  

35. 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]  

36. C. Kuang, S. Li, W. Liu, X. Hao, Z. Gu, Y. Wang, J. Ge, H. Li, and X. Liu, “Breaking the diffraction barrier using fluorescence emission difference microscopy,” Sci. Rep. 3, 1441 (2013). [CrossRef]   [PubMed]  

37. Y. Wang, C. Kuang, Z. Gu, Y. Xu, S. Li, X. Hao, and X. Liu, “Time-gated stimulated emission depletion nanoscopy,” Opt. Eng. 52(9), 093107 (2013). [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 (9)

Fig. 1
Fig. 1 Illustration of CARS-D method.
Fig. 2
Fig. 2 (a) TEM image of the polystyrene beads with the size 6.25 μm × 6.25 μm; (b) Experimental CARS image of one polystyrene bead marked with the green circle in Fig. 2 (a); (c) Intensity distribution of Ias along the green line in Fig. 2 (b); (d) Theoretical CARS image of Fig. 2 (a); (e) Intensity distribution of Ias along the green line in Fig. 2 (d); (f) Local enlarged curve marked with the green circle in Fig. 2 (e).
Fig. 3
Fig. 3 (a) Doughnut CARS image of the polystyrene beads; (b) CARS-D image of the polystyrene beads; (c) Intensity distribution of Ias along the green line in Fig. 3(b); (d) Local enlarged curve marked with the green circle in Fig. 3 (e).
Fig. 4
Fig. 4 (a) Illustration of the effective excitation areas of conventional CARS signal (red dotted line), doughnut CARS signal (blue dotted line) and CARS-D signal (green dotted line). (b) A spoke-like test sample.
Fig. 5
Fig. 5 (a) Conventional CARS image. (b) Doughnut CARS image. (c) CARS-D image.
Fig. 6
Fig. 6 CARS-D images with different subtractive factors using the parula color-map.
Fig. 7
Fig. 7 CARS-D images with different subtractive factors using the lines color-map.
Fig. 8
Fig. 8 Spatial resolution as function of factors k1 and k2: (a) inner spatial resolution, SRin, as a function of k1 with k2 = kM = 1; (b) SRin as a function of k2 with k1 = kM = 1; (c) outer spatial resolution, SRout, as a function of k1 with k2 = kM = 1; (d) SRout as a function of k2 with k1 = kM = 1.
Fig. 9
Fig. 9 (a) Golgi complex visual image (as an irregular test sample), (b) conventional CARS image, (c) doughnut CARS image, and (d) CARS-D image.

Equations (13)

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

I a s E a s 2 [ ρ E p 2 ( r ) E s ( r ) ] 2 ,
E p ( r ) = E p 0 e x p ( r 2 r p 2 ) ,
E s 1 ( r ) = E s 10 e x p ( r 2 r s 1 2 ) ,
E s 2 ( r ) = E s 20 r e x p ( r 2 r s 2 2 ) ,
I a s 1 [ ρ E p 0 2 exp ( 2 r 2 r p 2 ) E s 1 exp ( r 2 r s 1 2 ) ] 2 ,
I a s 2 [ ρ E p 0 2 exp ( 2 r 2 r p 2 ) E s 2 r exp ( r 2 r s 2 2 ) ] 2 .
Δ I a s = I a s 1 k M I a s 2 ,
Δ I a s ρ 2 exp ( 16 r 2 λ 2 ) { exp [ 8 r 2 ( k 1 λ ) 2 ] k M M 2 r 2 exp [ 8 r 2 ( k 2 λ ) 2 ] } .
V 1 = 0 r d r 0 2 π d θ exp ( 16 r 2 λ 2 ) exp [ 8 r 2 ( k 1 λ ) 2 ] = π a 1 ,
a 1 = 16 λ 2 + 8 ( k 1 λ ) 2 ,
V 2 = 0 r d r 0 2 π d θ exp ( 16 r 2 λ 2 ) ( r 2 ) exp [ 8 r 2 ( k 2 λ ) 2 ] = π a 2 2 ,
a 2 = 16 λ 2 + 8 ( k 2 λ ) 2 ,
M = V 1 V 2 = a 2 2 a 1 .
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.