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

Determination of the full scattering matrix using coherent Fourier scatterometry

Open Access Open Access

Abstract

We demonstrate a method to obtain within an arbitrary numerical aperture (NA) the entire scattering matrix of a scatterer by using focused beam coherent Fourier scatterometry. The far-field intensities of all scattered angles within the NA of the optical system are obtained in one shot. The corresponding phases of the field are obtained by an interferometric configuration. This method enables the retrieval of the maximum available information about the scatterer from scattered far-field data contained in the given NA of the system.

© 2016 Optical Society of America

1. INTRODUCTION

In angular-resolved scatterometry, given a certain illumination (incident amplitude, phase, and polarization) and optical system [incident wavelength and numerical aperture (NA)], the set of all elements of the scattering matrix at all angles that can be detected by the system contains the maximum information about the scatter.

In this paper, we explore and demonstrate the limit of optical scatterometry by determining the polarization-resolved amplitude and phase of the scattered field using coherent illumination. The maximum information content not only is relevant for the basic understanding of diffraction problems in optics, but also is crucial for applications where pushing the technique to its limits is required. Important applications are metrology for chip fabrication in the semiconductor industry and surface inspection for particle contamination and defect detection in both semiconductor and plastic electronics.

For example, in optical lithography, along with the ability to print sub-wavelength structures, stringent quality control of the lithographic process is required. Generally, gratings are used in situ as the metrology target. Slight variations in the dose, exposure, or the environmental conditions, such as temperature and pressure, can lead to deviation of the grating shape from the desired one. Measurement of the shape parameters of the gratings can be done in principle with scanning electron microscopes and/or atomic force microscopes, but optical scatterometry is the de-facto preferred method because it is noninvasive and fast.

In optical scatterometry, the retrieval of the shape parameters of the grating is done by matching the experimental far-field scattering intensity distribution with the expected distribution that is calculated by rigorous computations using, for example, the rigorous coupled-wave analysis (RCWA) method [1,2]. To determine the ultimate resolution, one should gather the maximum amount of information about the scattering matrix. There are many variants of optical scatterometry techniques such as single-incidence angle reflectometry, 2-Θ scatterometry, spectroscopic ellipsometry, Fourier scatterometry, interferometric Fourier scatterometry, and, in recent years, coherent Fourier scatterometry (CFS) [313]. In particular, the latter can be made very fast since a focused coherent beam is used, and the scattering information at many angles is captured in one shot. Furthermore, when the period of the structure is such that diffracted orders overlap and the focused spot is scanned, the phase differences between these orders in CFS makes this technique more sensitive than incoherent Fourier scatterometry [14]. However, diffracted orders only overlap when the pitch is large enough for the given NA of the focusing lens and the illumination wavelength used. This implies that the advantage of CFS over incoherent scatterometry methods is limited to gratings of a certain minimum pitch. Therefore, we have recently proposed [15] an interferometric version of CFS, by which not only the amplitude but also the phase of the scattering matrix elements are determined, and hence higher sensitivity is achieved for arbitrary pitch. In interferometric CFS (ICFS), the polarization-resolved scattered field can be retrieved for not only scattering angles where orders overlap but also any scattering angle. ICFS utilizes a reference wave that interferes with the far field generated by CFS to capture phase information present in the zeroth order and overlapping orders if they exist. In the latter case, the difference in phase between the overlapping orders is retrieved by scanning, as is done for noninterferometric CFS. Furthermore, by measuring all possible orthogonal incident and scattered polarizations, one can recover the entire complex scattering matrix, i.e., all information that can be obtained within the given numerical aperture of the system. Note that we use the same lens to focus the incident wave into a spot and to project the scattered field onto the CCD.

In the present paper, we show experimental proof of the predictions presented in [15] by determining the full scattering matrix using ICFS. The content of the paper is as follows. In Section 2, we summarize the most important aspects of the theory. In Section 3, we describe experimental details, in particular setup and data acquisition. In Section 4, we discuss the comparison with simulations, and Section 5 contains the conclusions.

2. THEORY

Let us consider the scattering problem of a one-dimensional grating. We choose a coordinate system (x,y,z) as shown in Fig. 1 with the z axis coinciding with the optical axis of the focusing system. The permittivity is a periodic function of x and invariant along y. It is conventional to set z=0 at the top of the grating and to choose z positive in the direction of incidence (see Fig. 1). The incident field is denoted as Ei, and the reflected field from the grating is denoted as Er. If the grating thickness is d, then we can express the electric fields as Ei+Er when z<0.

 figure: Fig. 1.

Fig. 1. Schematic diagram of our approach to the problem of one shot scattering matrix determination for a large number of incident angles. (A) The xyz coordinate system is attached to the sample with the z=0 plane chosen at the interface between the grating and the incident medium, which is also the geometric focus plane of the objective. (B) The far-field maps of the complex amplitude of the field is projected on the CCD in the exit pupil of the objective where the coordinate system (ξη) is chosen such that ξ and η are parallel to the x and y direction, respectively. (C) The interface between the half-space z<0 and the grating and the interface between the grating and the substrate z>d are indicated by the dotted lines.

Download Full Size | PDF

With Rayleigh’s method [1618], the total field in the grating region can be computed, for example, with the RCWA [1,19,20]. The reflected field is then obtained by subtracting the incident field from the total field. In the half-space z<0, the reflected field is expanded into a sum of plane waves as follows:

Er=m[Emsrs+Emprp]exp[i(kmxrx+kmyrykmzrz)],z<0,
where kmr=(kmxr,kmyr,kmzr) is the reflected wave vector. s and p denote the s- and p-polarization states, respectively, i.e., the states for which the electric field is perpendicular and parallel, respectively, to the plane of incidence. kmr is related to incident wave vector ki by
kmxr=kxi+j2πΛ,kmyr=kyi.
kmzr=(ki)2(kmxr)2(kmyr)2,
where Λ is the grating pitch. It is helpful to express the grating reflection in a matrix form Emr=Rm(ki)Eji with
Rm=(rmsseiϕmssrmspeiϕmsprmpseiϕmpsrmppeiϕmpp),
where r˜msp=rmspeiϕmsp denotes the complex amplitude of the s-polarized component of the m-th reflected order due to an incident wave with wave vector ki and polarized parallel to the p direction. Note that when kyi=0, i.e., when the incident plane wave is in the plane perpendicular to the grating structure, the matrix Rm is diagonal because in that case s and p polarizations are uncoupled. The matrix Rm contains the complete set of reflection coefficients of the grating. In the inverse diffraction problem of grating shape reconstruction [4], it is of extreme importance to determine these coefficients as precisely as possible. The reason for this is that, because of the ill-posedness of the inverse diffraction problem [21], a fast and stable solution depends heavily on the accurate measurement of these coefficients. Any intensity-based techniques can at best reveal the phase difference between different elements.

In this work, by combining coherent Fourier scatterometry with temporal phase shifting interferometry, we show that it is possible to determine the complex reflection coefficients for many angles of incidence in one shot. The basic idea is to use a microscope objective (MO) to focus incident beam onto the scattering sample, which for our case is the one-dimensional grating (shown in regions A and C of Fig. 1). In region B of the same figure, the far-field coordinates (ξη) are defined by

ξ=kxrkiη=kyrki.
The maximum angle that can be detected is limited by the NA of the objective (ξ2+η2NA). We note that the definition [Eq. (5)] implies that ξ and η are parallel to the x and y direction, respectively.

In practice, because of the presence of the objective, there are some additional geometric transformations to consider. If the objective is perfectly isotropic, it does not create any additional phase difference between the s- and p-polarized components. In that case, the incident and reflected fields are related through Rm instead of Rm, where

Rm=fΩRmΩ1,
where Ω is the rotation of the electric field introduced by the objective, and f is a factor for energy conservation [15,22]. However, except for Rm, all the other factors in Eq. (6) are merely geometric and do not contain any information about the object. Rm maps the incident pupil (before the objective) to the outgoing pupil (after the objective) directly and so can be expressed in terms of the (ξη) system. Thus, from now on, we express the incident and outgoing fields in the (ξη) system as well, which, respectively, denote the field right before being incident on the objective and the field right after passing through the objective.

In the experimental results that we are going to present, the incident state of polarization is known and the incident field is measured by a wavefront sensor, so that the complex amplitudes in every point of the entrance pupil are known. The incident field is then decomposed on the (ξ, η) basis. The amplitudes of the matrix elements at a particular input/output polarization combination are directly measured by imaging the exit pupil onto a CCD camera, while the phase is obtained interferometrically by combining the field at the exit pupil with a reference beam.

3. EXPERIMENTAL REALIZATION

A. Setup

We designed and built a coherent scatterometer, along with an interferometer functionality, based on temporal phase shifting interferometry. The design is basically the one of a coherent Fourier scatterometer (see Ref. [14]) where a reference mirror with a piezoelectric transducer is added to the open port of the beam splitter (BS). This setup allows us to determine the scattering matrix of any object of interest. In this paper, we consider a periodic silicon-on-silicon square grating described by the parameters listed in Table 1. Those parameters, which have been used as input values for the RCWA simulations, have been measured with atomic force microscopy (grating height and period) and scanning electron microscopy (grating period). A schematic overview is shown in Fig. 2. The light from a He–Ne laser (S) beam (λ=633nm) is coupled to a single-mode fiber (SMF). The light coming out from the fiber is collimated (L1=20cm), and the desired input polarization direction is set with a Glan–Taylor polarizer (POLin in the figure). To define the different states of polarization, we will use the notation provided by Section 2. A BS separates the beam for the sample and reference mirror arm. On the sample arm, the polarized light is focused on the grating with a MO (Leitz Wetzlar, 20×, infinity corrected) of NA=0.4. In the reference arm, the beam is reflected by a λ/20 flat aluminum mirror controlled by a piezo translation stage (PZT). The reflected light from the grating in the exit pupil of the MO as seen from the sample is optically conjugated with the detector plane by two lenses (L2=40cm and L3=25cm). Before the beam is recorded by the CCD camera, another polarizer (POLout) allows the selection of output polarization. In our scheme, output polarization (ξ,η) means that the polarizer is parallel to the input polarization (ξ,η).

Tables Icon

Table 1. Physical Dimensions of the Grating Under Investigation

 figure: Fig. 2.

Fig. 2. Schematic overview of the experimental setup. S, He–Ne laser; SMF, single-mode fiber; L1, collimating lens; L2 and L3, telescopic lenses; BS, nonpolarizing beam splitter; POLin and POLout: polarizers; MO, microscope objective; PZT, piezoelectric translation stage; CCD, data acquisition camera.

Download Full Size | PDF

B. Data Acquisition

As previously mentioned, the setup (and therefore its working principle) can be employed to quantify the scattering properties of a variety of different structures. For the etched silicon grating mentioned above and for a wavelength of 633 nm and a NA=0.4 of the MO, only the zeroth order is captured by the CCD camera. Since there are no overlapping orders, the spot does not have to be scanned to retrieve the phase difference between them [14]. We obtain data for four input/output polarization combinations, namely, ξξ, ξη, ηη, and ηξ. For each of them, intensity frames for several axial positions of the reference mirror are obtained by applying specific voltages to the PZT. The phase of the object is then reconstructed from the intensity data by means of a five-step phase shifting algorithm [23]. In fact, since only the zeroth order is present, only one complex amplitude is required to be calculated, thus the aforementioned algorithm is sufficient. When higher orders are also present, we will need more phase steps.

In the case of phase retrieval by a temporal phase shifting algorithm, there is always an uncertainty in the piezo movement that results in an error in the intended phase change of the reference beam. This error can be minimized either by using phase retrieval algorithms that are less sensitive to the error in the piezo movement and/or by choosing the correct frame corresponding to the intended phase shift (in our case, it is π/2). We devised a correlation-based technique to minimize the error. The method is explained in Fig. 3. The phase retrieval algorithm is implemented with five intensity patterns recorded for the corresponding π/2 phase shifted reference arm of the interferometer. The correlation coefficient between the images is computed for displacement versus voltage, which indeed gives the information about the phase shift between the images. Five images are then chosen for the phase retrieval. A single image recorded by the detector is an interference pattern for a defined input and output polarization

Iμν=Iμνref+Iμνobj+2IμνrefIμνobjcos(ϕr),
where Iref and Iobj indicate the beam intensities of the reference arm and the object arm, respectively, μ=ξ,η, and ν=ξ,η. If we shift the interference pattern by the five phase step values Φr=0,π/2,π,3π/2,2π, then we have a collection of five different measured intensities {Ir} with r=1,,5. Acquiring five frames in identical input and output polarization conditions and then rearranging the intensities give the phase in the pupil [24]
ϕμ,ν=arctan[2(I2I4)2I3I5I1].
To extract the five intensity maps from the experimental data, we analyze the correlation ρj of the captured experimental images j with respect to the first measured picture j=1
ρj=ξ,η[(Ij(ξ,η)Ij)(I1(ξ,η)I1)][ξ,η(Ij(ξ,η)Ij)2][ξ,η(I1(ξ,η)I1)]2,
where Ij(ξ,η) indicates the j-th image at the pixel position (ξ,η), while Ij is the average over all the pixels of the j-th image.

 figure: Fig. 3.

Fig. 3. Correlation function ρ as a function of the piezo translation stage displacement. The maxima and minimums of the correlation curve are examined to monitor the movement of the piezo. Only four experimental images are shown for simplicity.

Download Full Size | PDF

Because we are recording images of an interference pattern, we expect the function ρj to have a cosine-like behavior (as confirmed by the data trend in Fig. 3). The five different phase steps Φr we need to consider correspond to the maxima and minima of ρj, along with the points of ρj=0 when Φr=π/2,3π/2 over one period. These points are highlighted with red marks in Fig. 3. In this way, by selecting the right images from the experimental data and substituting them into Eq. (8), we can retrieve the phase information we need. The obtained phase data is unwrapped using a quality guided path algorithm [25], specifically the one in Ref. [26]. Experimental data were treated with a smoothing filter to minimize the noise influence using a Gaussian function [27] in the windowed Fourier transform [28].

4. COMPARISON BETWEEN MEASUREMENT AND SIMULATION

To validate the measurements, rigorously simulated data have been obtained with the RCWA method [2,29]. As input for the simulations, we used the intensity and phase of the field as measured in the plane before the objective (MO in Fig. 2). In Figs. 4 and 5, the experimental and simulated intensities and phases of the scattered far field are shown for three different sets of polarization between the incident and the scattered light, namely ηη, ηξ, and ξξ. In the case of a nonbirefringent grating, we assume ηξ and ξη to be identical. Since in our example the scattered far field consists only of the zeroth order, we immediately obtain the complex elements of the scattering matrix.

 figure: Fig. 4.

Fig. 4. Measured (left) and simulated (right) intensities of the far field scattered by a grating illuminated by a focused field for different combinations of input and output polarizations. The incident wavelength is 633 nm, and NA=0.4. The grating parameters are given in Table 1.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. Phase of the scattered far field retrieved from measurements (left) and simulations (right) for different combinations of input and output polarizations. The incident wavelength is 633 nm, and NA=0.4. The grating parameters are given in Table 1.

Download Full Size | PDF

The measured far-field intensities and phases show a good match between experiments and simulations. The differences are attributed to extra insertion losses and noise that have not been taken into account in the simulations. In Fig. 5, the deviation between experimental and simulated phase for the ηξ case for normal incidence can be attributed to the low intensity levels in the far fields used to reconstruct the phase map. Finally, we add the ability to obtain the phase maps to define the value of the technique for phase sensitive scatterometry. With phase and amplitude knowledge of the scatterer, all field components (in our case, reflected) from the object within the NA of the optical system, except a constant phase, are obtained.

5. CONCLUSIONS

In this paper, we demonstrate a robust, fast, and reliable way to obtain the entire scattering matrix of a periodic object by measuring the phase of the scattered field from a periodic object at all angles within the numerical aperture of the system using ICFS. With ICFS, one is able to extend the present method of optical scatterometry to its maximum potential. The method presented here provides information on the scattering matrix resolved over polarization for all scattering angles, which can be extended to higher diffraction orders if they exist. Although the results presented above are for small NA (0.4) and long wavelength (633 nm), they can be scaled to higher NA and shorter wavelengths. This approach is not limited to periodic objects and thus can be applied when scattered light from an arbitrary scatterer is used to retrieve information about it. Because the complex scattering matrix provides all possible information, we believe that this method can also be used to set the limits of optical scatterometry in different applications such as object parameter retrieval, detection of (sub-wavelength) particle contamination, defect detection, and surface characterization. Furthermore, depending on the features that are to be retrieved, one can determine the measured data that contain the most sensitive part of the information for this feature and discard the other data. In this way, the speed of scatterometry can be increased.

Funding

NanoNextNL.

Acknowledgment

The Ph.D. work of L.C. is supported by NanoNextNL, a micro and nanotechnology consortium of the government of the Netherlands and 130 partners.

REFERENCES

1. M. G. Moharam and T. K. Gaylord, “Rigorous coupled-wave analysis of planar-grating diffraction,” J. Opt. Soc. Am. 71, 811–818 (1981). [CrossRef]  

2. L. Li, “Use of Fourier series in the analysis of discontinuous periodic structures,” J. Opt. Soc. Am. A 13, 1870–1876 (1996). [CrossRef]  

3. S. Roy, A. C. Assafrão, S. F. Pereira, and H. P. Urbach, “Coherent Fourier scatterometry for detection of nanometer-sized particles on a planar substrate surface,” Opt. Express 22, 13250–13262 (2014). [CrossRef]  

4. V. F. Paz, S. Peterhansel, K. Frenner, and W. Osten, “Solving the inverse grating problem by white light interference Fourier scatterometry,” Light Sci. Appl. 1, e36 (2012).

5. H.-T. Huang, W. Kong, and F. L. Terry Jr., “Normal-incidence spectroscopic ellipsometry for critical dimension monitoring,” Appl. Phys. Lett. 78, 3983 (2001). [CrossRef]  

6. Q. Zhan and J. Leger, “High resolution imaging ellipsometer,” Appl. Opt. 41, 4443–4450 (2002). [CrossRef]  

7. P. Boher, J. Petit, L. Teroux, J. Foucher, Y. Desieres, J. Hazard, and P. Chaton, “Optical Fourier transform scatterometry for LER and LWR metrology,” Proc. SPIE 5752, 192–203 (2005). [CrossRef]  

8. B. B. M. Wurm and F. Pilarski, “A new flexible scatterometer for critical dimensional metrology,” Rev. Sci. Instrum. 81, 023701 (2010). [CrossRef]  

9. O. El Gawhary, N. Kumar, S. F. Pereira, W. M. J. Coene, and H. P. Urbach, “Performance analysis of coherent optical scatterometry,” Appl. Phys. B 105, 775–781 (2011).

10. D. Kim, M. Jin, H. Lee, S. Kim, and R. Magnusson, “Snapshot conical diffraction phase image measurement in angle-resolved microellipsometry,” in Imaging and Applied Optics (Optical Society of America, 2013), paper. CTu3C.5.

11. A. Faridian, V. F. Paz, K. Frenner, G. Pedrini, A. D. Boef, and W. Osten, “Phase-sensitive structured illumination to detect nanosized asymmetries in silicon trenches,” J. Micro/Nanolithogr. MEMS MOEMS 14, 021104 (2015).

12. M. H. Madsen and P.-E. Hansen, “Imaging scatterometry for flexible measurements of patterned areas,” Opt. Express 24, 1109–1117 (2016). [CrossRef]  

13. R. M. Silver, B. M. Barnes, R. Attota, J. Jun, M. Stocker, E. Marx, and H. J. Patrick, “Scatterfield microscopy for extending the limits of image-based optical metrology,” Appl. Opt. 46, 4248–4257 (2007). [CrossRef]  

14. N. Kumar, O. El Gawhary, S. Roy, S. E. Pereira, and H. P. Urbach, “Phase retrieval between overlapping orders in coherent Fourier scatterometry using scanning,” J. Eur. Opt. Soc. 8, 13048 (2013). [CrossRef]  

15. S. Roy, N. Kumar, S. F. Pereira, and H. P. Urbach, “Interferometric coherent Fourier scatterometry: a method for obtaining high sensitivity in the optical inverse-grating problem,” J. Opt. 15, 075707 (2013).

16. L. Rayleigh, “On the dynamical theory of gratings,” Proc. R. Soc. London A 79, 399–416 (1907).

17. P. M. van den Berg, “Reflection by a grating: Rayleigh methods,” J. Opt. Soc. Am. 71, 1224–1229 (1981). [CrossRef]  

18. A. V. Tishchenko, “Numerical demonstration of the validity of the Rayleigh hypothesis,” Opt. Express 17, 17102–17117 (2009). [CrossRef]  

19. K. Knop, “Rigorous diffraction theory for transmission phase gratings with deep rectangular grooves,” J. Opt. Soc. Am. 68, 1206–1210 (1978). [CrossRef]  

20. J. Chandezon, G. Raoult, and D. Maystre, “A new theoretical method for diffraction gratings and its numerical application,” J. Opt. 11, 235–241 (1980).

21. M. N. Vesperinas, Scattering and Diffraction in Physical Optics, 2nd ed. (World Scientific, 2005), Chap. 9.9, pp. 316–319.

22. P. Török and T. Wilson, “Rigorous theory for axial resolution in confocal microscopes,” Opt. Commun. 137, 127–135 (1997).

23. D. Malacara, M. Servin, and Z. Malacara, Interferogram Analysis for Optical Testing, 2nd ed. (CRC Press, 2005), Chap. 7.

24. N. Kumar, “Coherent Fourier scatterometry,” Ph.D. thesis (TU Delft, 2014).

25. M. A. Herraez, D. R. Burton, M. J. Lalor, and M. A. Gdeisat, “Fast two-dimensional phase-unwrapping algorithm based on sorting by reliability following a noncontinuous path,” Appl. Opt. 41, 7437–7444 (2002).

26. M. Gdeisat and F. Lilley, “Two-dimensional phase unwrapping problem,” https://www.ljmu.ac.uk/~/media/files/ljmu/about-us/faculties-and-schools/tae/geri/two_dimensional_phase_unwrapping_finalpdf.pdf?la=en.

27. S. Qian, Introduction to Time-Frequency and Wavelet Transform, 2nd ed. (Prentice-Hall, 2002).

28. Q. Kemao, “Two-dimensional windowed Fourier transform for fringe pattern analysis: principles, applications and implementation,” Opt. Lasers Eng. 45, 304–317 (2007).

29. M. van Kraaij and J. Maubauch, “A more efficient rigorous coupled-wave analysis algorithm,” in Progress in Industrial Mathematics at ECMI 2004, Vol. 8 of Mathematics in Industry, (Springer, 2006), pp. 164–168.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. Schematic diagram of our approach to the problem of one shot scattering matrix determination for a large number of incident angles. (A) The x y z coordinate system is attached to the sample with the z = 0 plane chosen at the interface between the grating and the incident medium, which is also the geometric focus plane of the objective. (B) The far-field maps of the complex amplitude of the field is projected on the CCD in the exit pupil of the objective where the coordinate system ( ξ η ) is chosen such that ξ and η are parallel to the x and y direction, respectively. (C) The interface between the half-space z < 0 and the grating and the interface between the grating and the substrate z > d are indicated by the dotted lines.
Fig. 2.
Fig. 2. Schematic overview of the experimental setup. S, He–Ne laser; SMF, single-mode fiber; L1, collimating lens; L2 and L3, telescopic lenses; BS, nonpolarizing beam splitter; POL in and POL out : polarizers; MO, microscope objective; PZT, piezoelectric translation stage; CCD, data acquisition camera.
Fig. 3.
Fig. 3. Correlation function ρ as a function of the piezo translation stage displacement. The maxima and minimums of the correlation curve are examined to monitor the movement of the piezo. Only four experimental images are shown for simplicity.
Fig. 4.
Fig. 4. Measured (left) and simulated (right) intensities of the far field scattered by a grating illuminated by a focused field for different combinations of input and output polarizations. The incident wavelength is 633 nm, and NA = 0.4 . The grating parameters are given in Table 1.
Fig. 5.
Fig. 5. Phase of the scattered far field retrieved from measurements (left) and simulations (right) for different combinations of input and output polarizations. The incident wavelength is 633 nm, and NA = 0.4 . The grating parameters are given in Table 1.

Tables (1)

Tables Icon

Table 1. Physical Dimensions of the Grating Under Investigation

Equations (9)

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

E r = m [ E m s r s + E m p r p ] exp [ i ( k m x r x + k m y r y k m z r z ) ] , z < 0 ,
k m x r = k x i + j 2 π Λ , k m y r = k y i .
k m z r = ( k i ) 2 ( k m x r ) 2 ( k m y r ) 2 ,
R m = ( r m s s e i ϕ m s s r m s p e i ϕ m s p r m p s e i ϕ m p s r m p p e i ϕ m p p ) ,
ξ = k x r k i η = k y r k i .
R m = f Ω R m Ω 1 ,
I μ ν = I μ ν ref + I μ ν obj + 2 I μ ν ref I μ ν obj cos ( ϕ r ) ,
ϕ μ , ν = arctan [ 2 ( I 2 I 4 ) 2 I 3 I 5 I 1 ] .
ρ j = ξ , η [ ( I j ( ξ , η ) I j ) ( I 1 ( ξ , η ) I 1 ) ] [ ξ , η ( I j ( ξ , η ) I j ) 2 ] [ ξ , η ( I 1 ( ξ , η ) I 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.