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

Double-channel angular-multiplexing polarization holography with common-path and off-axis configuration

Open Access Open Access

Abstract

We propose a double-channel angular-multiplexing polarization holographic imaging system with common-path and off-axis configurations. In the system, its input plane is spatially divided into three windows: an object window and two reference windows, and two orthogonal linear polarizers are attached, respectively, on the two reference windows; a two-dimensional cross grating is inserted between the input and output planes of the system. Thus the object beam passing through the object window and the two orthogonal polarized reference beams passing through the two reference windows can overlap each other at the output plane of the system and form a double-channel angular-multiplexing polarization hologram (DC-AM-PH). Using this system, the complex amplitude distributions of two orthogonal polarized components from an object can be recorded and reconstructed by one single-shot DC-AM-PH at the same time. Theoretical analysis and experimental results demonstrated that the system can be used to measure the Jones matrix parameters of polarization-sensitive or birefringent materials.

© 2017 Optical Society of America

1. Introduction

Digital holography (DH), due to its superior ability in measuring and imaging the complex amplitude distribution (including both the amplitude and phase information) of an object wavefront, has drawn much attention in recent years and has been widely applied in many significant fields such as quantitative phase microscopy, contour measurement and nondestructive tests [1–8]. According to the relative orientation between the object and the reference beams in recording geometry, DHs are often divided into two categories: on-axis (or in-line) and off-axis (or off-line). It is well known that the on-axis configuration can reduce the spatial bandwidth of the recorded holograms and thus make full use of the resolving power of the image sensor; but the reconstructed images often suffer from the disturbances of the zero-order and twin images, which need to be removed by using phase shifting operation or other time-consuming elimination algorithms [9–18]. In off-axis DHs, the zero-order and twin images are well separated from the required image in spatial frequency domain and so they can be removed from a single hologram simply by a digital spatial filtering, but at the cost of a high bandwidth demand on the imaging sensor. As the spatial resolution and the pixel number of image sensors have been significantly increased in recent years (for example, the pixel size of a current commercial image sensor can be close to 1 μm and the pixel number can reach more than 20 megapixels), off-axis digital holography is becoming a powerful tool in dynamic phase measuring and imaging applications [19–25].

DHs with common-path and off-axis configuration denote a class of holographic recording architectures, in which both the imaging and the reference beams for off-axis interferometric recording propagate nearly through the same optical path and the same imaging lenses [26–33]. Compared with the non-common path DHs such as those based on Mach-Zehnder or Michelson interferometers, common-path configurations are more stable, relatively insensitive to mechanical vibrations and air fluctuations, and require fewer optical elements. One-arm lateral shearing interferometers (LSIs) [33] are the most widely used common-path configurations. In conventional LSI-based common-path DHs, the holograms are mainly formed by interference between two beams sheared from the same object beam passing through the same optical path and imaging system. Because both the two beams involved in the interference are disturbed by the object, the reconstructed image will be disrupted by some additional undesired terms (such as the duplicated image), along with the zero-order and twin images. To overcome this drawback, some modified LSI systems used for common-path off-axis DHs have been developed in recent years [34–41]. For example, Mico et al. [34] proposed a common-path off-axis DH system based on spatially-multiplexed LSI, in which the input plane of the system is spatially divided into two parts in side-by-side configuration: the object region under test and a black region for the reference; at the same time, a one-dimensional grating is inserted between the input and output planes of the system to realize the overlap and off-axis interference of the object and reference beams at the output plane. Seo et al. [35] also proposed a similar modified LSI for the same purpose nearly at the same time, in which an optical window glass was used as the shearing element. Because the common-path and off-axis configurations, based on the modified LSI, have the advantages of simple structure and high anti-disturbing ability and provide a low-cost way to convert a regular microscope into a holographic one, they have drawn much attention most recently and have been applied in quantitative phase imaging and surface inspection [36–41].

In this paper, we propose a common-path off-axis polarization holographic (COPH) imaging system based on a modified double-channel LSI configuration. In the proposed system, the input plane of the imaging system is spatially divided into three windows: an object window (or object window) and two reference windows; two orthogonal linear polarizers are placed, respectively, on the two reference windows; in the path space between the input and output planes of the imaging system is inserted a specially designed two-dimensional (2-D) cross grating. Thus, the object beam passing through the object window and the two orthogonal polarized reference beams passing through the two reference windows can overlap each other at the output plane of the system and form a double-channel angular-multiplexing polarization hologram (DC-AM-PH). Using this system, the complex amplitude distributions (including both the amplitude and phase information) of the two orthogonal polarized components of the object image can be recorded and retrieved by one single-shot DC-AM-PH at the same time.

2. Principle of the system

Figure 1 shows the thematic of our proposed COPH system. It is essentially an imaging system mainly composed of an objective lens L1 and a tube lens L2. The main improvements are: a special limiting screen P1 with three windows is placed at the input plane and a 2D cross grating CG is inserted between L1 and L2. The object S to be tested is put in the central window (named object window). Two linear polarizer with orthogonal polarization directions, PR1 and PR2, are respectively attached behind the other two windows (named reference windows). Thus, after passing through this limiting screen P1, the input beam (its polarization state can be adjusted by a half wave plate WP, as required) is divided into three parts: the object beam passing through the object window and the two orthogonal polarized reference beams passing through the two reference windows. These three parallel transmission beams co-pass through the objective lens L1, the cross grating CG and the tube lens L2 and are finally imaged on the output plane P2. Due to the diffraction effect of the cross grating CG inserted between L1 and L2, the complex amplitude distribution of the output field at the output plane P2 will be composed of some replica images separated from each other, each corresponding to one of the diffraction orders from the grating CG. If the separation distance of the windows, the grating parameters and other system parameters are set to satisfy some conditions, one image of the object beam will be superimposed with the images of the two reference beams at the center part of the output plane as shown in Fig. 1, and form a DC-AM-PH recording the complex amplitude information of two orthogonal polarized components of the object image, which can be reconstructed by conventional spatial filtering algorithm used in angular-multiplexing off-axis holography [42]. In the following paragraphs, we will give first its theoretical analysis and then give its verification by some experiments in next section.

 figure: Fig. 1

Fig. 1 Schematic of the proposed COPH system

Download Full Size | PDF

We know there are two approaches to describe polarized beams and polarization-sensitive materials: using Stokes vector and Mueller matrix or using Jones vector and Jones matrix [43]. The Stokes formalism is usually used for intensity superposition problems, while the Jones calculus is often applicable for complex amplitude superposition problems. Like most of holographic polarization imaging methods [44–50], we also use Jones formalism to describe the measured parameters in the following analysis.

Supposing the input beam is first linearly polarized along x axis (horizontal) and the transmission directions of the linear polarizers PR1 and PR2 are set, respectively, along + 45° and −45°, the Jones vector of the field passing through the limiting screen P1 can be expressed as

u1(x,y)=rect(xa,ya)[Ex(x,y)Ey(x,y)]+R1rect(xba,ya)[11]+R2rect(xa,yba)[11],
where, rect(x/a,y/a) is a square window function with the width of a, R1 and R2 are the amplitudes of the beams respectively passing through the two reference windows, and b is the central distance of the reference windows away from the object window along x or y axis, respectively; while Ex(x,y) and Ey(x,y) are the complex amplitudes of two Jones vector components of the object field passing through the object window, which also can be respectively expressed by Jones matrix parameters of the object as Ex=AxJxx(x,y) and Ey=AxJyx(x,y) (where Ax corresponds to the amplitude of the input horizontally polarized beam). As indicated above, the complex amplitude distribution of the output field at the image plane P2 will be composed of some replica images separated from each other, due to the diffraction effect of the cross grating CG. Considering only the replicas corresponding to the zero order and the two + 1 orders respectively along x and y directions, as shown at the image plane P2 of Fig. 1, the output complex amplitude distribution can be written as
u2(x,y)=D0u1(xM,yM)+D1xu1(x+dgM,yM)exp[j2πξgx]+D1yu1(xM,y+dgM)exp[j2πξgy0],
in which, M=f2/f1 is the amplification of the image system, f1 and f2 are, respectively, the focal lengths of L1 and L2; D0, D1x and D1y are three constants related to the amplitude of the input beam and the diffraction efficiencies of the cross grating; the shift distance dg and the tilt frequency parameter ξg (corresponding to the carrier frequency of the reference beams) of the replicas depend on the configuration parameters of the grating and the imaging system, which can be determined by the following formulas [51]:
dg=λf2T,and ξg=Δf2T,
where λ is the wavelength of the input beam, T is the period of the cross grating, and Δ is the distance of the cross grating away from the front focal plane of the tube lens as shown in Fig. 1. It can be seen from Eq. (3) that the carrier frequency or the off-axis angle (sinθ=λξg) of the reference beams can be simply aligned by changing the distance parameter Δ.

Substituting Eq. (1) into Eq. (2), we can find that, if the central distance b between the object window and the reference windows is set to satisfy the condition of

b=dgM=λf1T,
the images of the beams from the two reference windows with orthogonal linear polarization states will exactly overlap with the image of the object window in the center region highlighted by the red dotted square shown in the image plane P2 of Fig. 1. The total complex amplitude in this region can be expressed as
uc(x,y)=rect(xMa,yMa)[D0Jxx(xM,yM)+D1xR1exp[j2πξgx]+D1yR2exp[j2πξgy]D0Jyx(xM,yM)+D1xR1exp[j2πξgx]D1yR2exp[j2πξgy]].
Its intensity distribution just forms a DC-AM-PH simultaneously recording the two orthogonal linear polarization components of the object beam from the object window, which can be written as
H(x,y)=|D0Jxx(xM,yM)+D1xR1exp[j2πξgx]+D1yR2exp[j2πξgy]|2+|D0Jyx(xM,yM)+D1xR1exp[j2πξgx]D1yR2exp[j2πξgy]|2,=Y0(x,y)+Y1x(x,y)+Y1y(x,y)+Y1x*(x,y)+Y1y*(x,y)
in which,
Y0(x,y)=|D0Jxx(xM,yM)|2+|D0Jyx(xM,yM)|2+2|D1xR1|2+2|D1yR2|2,
and

Y1x(x,y)=D0D1xR1[Jxx(xM,yM)+Jyx(xM,yM)]exp[j2πξgx]Y1y(x,y)=D0D1yR2[Jxx(xM,yM)Jyx(xM,yM)]exp[j2πξgy].

Figure 2 gives a typical example of the spatial Fourier transform of the DC-AM-PH described by Eq. (6). It can be seen from Fig. 2 that simultaneously recording of the two polarization components is performed while sharing the dynamic range of an off-axis hologram by multiplexing two off-axis interferograms into a single hologram. This kind of angular multiplexing technique for digital holograms has already been shown to be feasible and has been used for many applications such as holographically recording ultrafast events and increasing super-resolution capabilities of an imaging system [52–55]. So the terms Y1x and Y1y in Eq. (8) can be directly extracted from one DC-AM-PH using a spatial filtering algorithm for reconstructing angular multiplexing holograms. Then the Jones parameters Jxx and Jyx under test can be determined by the following formulas

Jxx(xM,yM)=12(Y1x(x,y)Y1x0(x,y)+Y1y(x,y)Y1y0(x,y))Jyx(xM,yM)=12(Y1x(x,y)Y1x0(x,y)Y1y(x,y)Y1y0(x,y)).
Here Y1x0 and Y1y0 correspond to the reconstructed values from a background DC-AM-PH, which can be obtained by performing a measurement before the object is placed in the object window.

 figure: Fig. 2

Fig. 2 A typical example of the spatial spectrum of the DC-AM-PH described by Eq. (6).

Download Full Size | PDF

In the above analysis, if we rotate the polarization direction of the input linearly polarized beam to y axis (vertical direction), the other two Jones matrix parameters Jxy and Jyy of the object can be measured using a formula similar to Eq. (9). Using all the measured Jones parameters Jxx, Jyx, Jxy and Jyy, we can further image and quantitatively analyze the optical anisotropic properties of the object such as birefringent properties [45–47].

3. Experimental results

For demonstrating the feasibility of our proposed method, we established an experimental setup based on the schematic shown in Fig. 1 and measured the Jones matrix parameters of a liquid crystal display (LCD) panel using the setup. In the experiments, a He-Ne laser, with the wavelength of 632.8 nm, is used as the light source. The 2-D cross grating adopted in the experiments is a cross Ronchi grating of equal widths of opaque and transparent rulings with a period of 8 μm, which is photoetched on a chromium-coated optical glass substrate. The tested object is a twisted-nematic LCD panel, with pixel size of 18 × 18 μm, and pixel number of 1024 × 768, without input and output polarizers, which was placed at the object window. Two film linear polarizers, with orthogonal polarization directions are, respectively, attached behind the two reference windows. Because the object to be imaged is of a relative large size, the pupil diameter of the objective lens and the tube lens used in the experiments are taken as 50 mm, with focal lengths of, respectively, 180 mm and 150mm. To further increase the effective size of the object window under the given field of view (FOV) of the objective lens, the object window designed in the experimental setup was shifted to the lower left of the input plane, as shown in Fig. 3(a), in which the dotted circle indicates the range of the FOV of the objective lens. In this configuration, the size of the object window is about a quarter of the original FOV of the objective lens.

 figure: Fig. 3

Fig. 3 (a) Photo of the designed input widows attached with the test object. (b) Gray level picture displayed on the measured LCD; scale bar: 1.8 mm. (c) An example of the DC-AM-PHs recorded in experiments; scale bar: 645um.

Download Full Size | PDF

Because the polarization modulation properties of individual LCD pixels are a function of an applied bias (or voltage) and the bias is controlled, point by point, by the gray levels (usually ranging from 0 to 255) of a gray-scale picture displayed on the LCD panel, the applied bias is often replaced by the corresponding gray level in characterizing the modulation properties of the LCD. In the experiments, we design a picture with a continuous variation of gray levels from 0 to 255, along one dimension, as shown in Fig. 3(b) (only part of the picture is shown). Thus the information for calculating the Jones matrix parameters of the LCD pixels versus different gray levels (corresponding to different applied bias) can be recorded by a single shot DC-AM-PH through displaying the designed picture on the measured LCD panel. The DC-AM-PHs are recorded by a CCD image sensor, with pixel size of 6.45 × 6.45 μm, and pixel number of 1392 × 1040. Figure 3(c) shows an example of the DC-AM-PHs recorded in our experiments. The region outlined with the red dashed rectangle in Fig. 3(c) corresponds to the parts marked with the dashed rectangles in Figs. 3(a) and 3(b). The procedures for extracting the four Jones matrix parameters from the recorded DC-AM-PHs are as follows: (1) transform the DC-AM-PH, recorded when the polarization direction of the input beam is set to x axis (horizontal direction), into the spatial frequency domain using a fast Fourier transform; (2) crop the two terms Y1x and Y1y, respectively, from the spatial spectrum of the DC-AM-PH as shown in Fig. 2, and inversely transform them back to the spatial domain; (3) calculate the Jones parameters Jxx and Jyx, based on Eq. (9); (4) repeat the previous two steps to extract the other two Jones parameters Jxy and Jyy from the DC-AM-PH recorded after we rotate the polarization direction of the input beam to y axis (vertical direction). The terms Y1x0 and Y1y0 used in Eq. (9) can be obtained from the corresponding background DC-AM-PHs recorded before the object is inserted into the object window.

Figure 4 gives an example of the measured Jones parameter distributions of the tested LCD panel in the part marked by the dashed rectangle in Fig. 3. Figures 4(a)-4(d) are, respectively, the amplitude distributions of the Jones parameters Jxx, Jxy, Jyx and Jyy, while Figs. 4(e)-4(h) correspond to their phase distributions. In order to quantitatively evaluate the measured values,

 figure: Fig. 4

Fig. 4 Example of the measurement results. (a)-(d) retrieved amplitude distributions of the four Jones matrix parameters Jxx, Jxy, Jyx and Jyy of the measured LCD; (e)-(h) their respective phase distributions. Scale bar: 1.6 mm.

Download Full Size | PDF

In order to quantitatively evaluate the measured values, Fig. 5 further gives the curves of the relation between the Jones parameters and the applied gray levels, fitted according to the experimental data shown in Fig. 4. Figure 5(a) gives the curves of the amplitudes of the Jones parameters versus the gray levels, while Fig. 5(b) shows the curves of the corresponding phases versus the gray levels. It can be seen that the amplitudes of the Jones matrix parameters of the tested LCD panel have the relationships of |Jxx||Jyy| and |Jxy||Jyx|, which is consistent with the theoretical prediction for such a twisted-nematic LCD panel [56,57].

 figure: Fig. 5

Fig. 5 Fitted curves of the Jones parameters versus the gray levels displayed on the LCD. (a) and (b) are, respectively, the amplitude and the phase values versus gray level.

Download Full Size | PDF

For testing the correctness and the usability of the measured Jones parameters, we further calculated the complex amplitude transmittance of the tested LCD panel with both input and output linear polarizers, based on the measured data and the following formula:

t=cos(θ2)(Jxxcos(θ1)+Jxysin(θ1))+sin(θ2)(Jyxcos(θ1)+Jyysin(θ1)),
in which, θ1 and θ2 are, respectively, the azimuth angles of the input polarizer axis and the output polarizer axis. At the same time, we also experimentally measured the amplitude transmittance of the tested LCD panel when the same gray-scale picture shown in Fig. 3(b) was displayed on it, under the corresponding polarizer configurations. Figures 6(a) and 6(b) give two examples of the results, in which Fig. 6(a) is the amplitude transmittance |t| in the marked part of the LCD panel when the input and output polarizers are set, respectively, to θ1=45° and θ1=90°, while Fig. 6(b) is the result when the polarizer axes are taken as θ1=45° and θ1=135°. It can be seen that the predicted results, according to the measured Jones parameters, basically agree with the results directly measured in real experiments.

 figure: Fig. 6

Fig. 6 Comparison between the calculated amplitude transmittance and the measured results. (a) Amplitude transmittance |t| of the LCD panel when the input and output polarizers are set, respectively, to θ1=45° and θ1=90°; (b) the result when the polarizer axes are taken as θ1=45° and θ1=135°.

Download Full Size | PDF

4. Conclusion

Our theoretical analysis and experimental results demonstrated that the proposed COPH system, established by implementing only two small modifications to a conventional lens imaging system: a specific three-window screen put on the input plane and a 2D cross grating inserted between the objective lens and tube lens, can realize single-shot holographic measurement of two orthogonally polarized components of the object beam, and so can be used to measure the Jones matrix parameters of polarization-selective or birefringent materials. As an example, we measured the four Jones parameters of a twisted-nematic LCD panel using this system. Based on the measured Jones parameters, we predicted the amplitude transmittance of the tested LCD panel after adding both input and output linear polarizers. The predicted results basically agree with the results directly observed in real experiments. Compared with existing methods for realizing holographic polarization measurement [44–50], the proposed COPH system may has some advantages because of its common-path interferometric architecture and a minimum of optical elements in the experimental setup. For example, the phase measurement accuracy can be improved since both the object and the reference beams for off-axis interferometric recording nearly propagate through the same optical path and the same imaging elements and so the system is insensitive to environmental perturbations due to mechanical or thermal changes. And a low coherent light source can be used as the illumination to eliminate the influence of the coherent noise on the measurements. This method may provide a low-cost way to convert a regular polarization microscope into a holographic polarization one with only minimal modifications and could be applied for quantitative measurements of birefringent and optical anisotropic specimens often required in biological and material science [44–50,58]. But what needs to be indicated is that, the size of the tested object will be limited because of the spatial multiplexing at the input plane.

Funding

National Natural Science Foundation of China (NSFC) (11474186).

References and links

1. W. Osten, A. Faridian, P. Gao, K. Körner, D. Naik, G. Pedrini, A. K. Singh, M. Takeda, and M. Wilke, “Recent advances in digital holography [invited],” Appl. Opt. 53(27), G44–G63 (2014). [CrossRef]   [PubMed]  

2. A. Anand and B. Javidi, “Digital holographic microscopy for automated 3D cell identification: an overview,” Chin. Opt. Lett. 12(6), 060012 (2014). [CrossRef]  

3. P. Memmolo, L. Miccio, M. Paturzo, G. Di Caprio, G. Coppola, P. A. Netti, and P. Ferraro, “Recent advances in holographic 3D particle tracking,” Adv. Opt. Photonics 7(4), 713–755 (2015). [CrossRef]  

4. X. Chen, J. Zhao, J. Wang, J. Di, B. Wu, and J. Liu, “Measurement and reconstruction of three-dimensional configurations of specimen with tiny scattering based on digital holographic tomography,” Appl. Opt. 53(18), 4044–4048 (2014). [CrossRef]   [PubMed]  

5. M. Aakhte, V. Abbasian, E. A. Akhlaghi, A. R. Moradi, A. Anand, and B. Javidi, “Microsphere-assisted super-resolved Mirau digital holographic microscopy for cell identification,” Appl. Opt. 56(9), D8–D13 (2017). [CrossRef]   [PubMed]  

6. N. Liu, Y. Zhang, and J. Xie, “Large object investigation by digital holography with effective spectrum multiplexing under single-exposure approach,” Appl. Phys. Lett. 105(15), 151901 (2014). [CrossRef]  

7. F. Merola, P. Memmolo, L. Miccio, R. Savoia, M. Mugnano, A. Fontana, G. D’Ippolito, A. Sardo, A. Iolascon, A. Gambale, and P. Ferraro, “Tomographic flow cytometry by digital holography,” Light Sci. Appl. 6(4), e16241 (2017). [CrossRef]  

8. M. J. Berg, N. R. Subedi, and P. A. Anderson, “Measuring extinction with digital holography: nonspherical particles and experimental validation,” Opt. Lett. 42(5), 1011–1014 (2017). [CrossRef]   [PubMed]  

9. X. F. Xu, L. Z. Cai, Y. R. Wang, X. F. Meng, W. J. Sun, H. Zhang, X. C. Cheng, G. Y. Dong, and X. X. Shen, “Simple direct extraction of unknown phase shift and wavefront reconstruction in generalized phase-shifting interferometry: algorithm and experiments,” Opt. Lett. 33(8), 776–778 (2008). [CrossRef]   [PubMed]  

10. J. Deng, H. Wang, D. Zhang, L. Zhong, J. Fan, and X. Lu, “Phase shift extraction algorithm based on Euclidean matrix norm,” Opt. Lett. 38(9), 1506–1508 (2013). [CrossRef]   [PubMed]  

11. M. Shan, B. Hao, Z. Zhong, M. Diao, and Y. Zhang, “Parallel two-step spatial carrier phase-shifting common-path interferometer with a Ronchi grating outside the Fourier plane,” Opt. Express 21(2), 2126–2132 (2013). [CrossRef]   [PubMed]  

12. C. Meneses-Fabian, R. Kantun-Montiel, G. P. Lemus-Alonso, and U. Rivera-Ortega, “Double aperture common-path phase-shifting interferometry by translating a ruling at the input plane,” Opt. Lett. 38(11), 1850–1852 (2013). [CrossRef]   [PubMed]  

13. C. S. Guo, X. J. Zhang, and B. Sha, “Non-iterative blind phase-shifting algorithm for two-step phase-shifting interferometry based on an analytical formula,” Opt. Commun. 315, 275–279 (2014). [CrossRef]  

14. C. S. Guo, B. Sha, Y. Y. Xie, and X. J. Zhang, “Zero difference algorithm for phase shift extraction in blind phase-shifting holography,” Opt. Lett. 39(4), 813–816 (2014). [CrossRef]   [PubMed]  

15. E. Stoykova, H. Kang, and J. Park, “Twin-image problem in digital holography-a survey,” Chin. Opt. Lett. 12(6), 060013 (2014). [CrossRef]  

16. T. Man, Y. Wan, and D. Wang, “Phase shift steps extraction and phase shift error correction in partially coherent illuminated phase-shifting digital holography,” Appl. Opt. 54(7), 1839–1843 (2015). [CrossRef]  

17. T. D. Yang, H. J. Kim, K. J. Lee, B. M. Kim, and Y. Choi, “Single-shot and phase-shifting digital holographic microscopy using a 2-D grating,” Opt. Express 24(9), 9480–9488 (2016). [CrossRef]   [PubMed]  

18. C. Tian and S. Liu, “Demodulation of two-shot fringe patterns with random phase shifts by use of orthogonal polynomials and global optimization,” Opt. Express 24(4), 3202–3215 (2016). [CrossRef]   [PubMed]  

19. P. Girshovitz and N. T. Shaked, “Compact and portable low-coherence interferometer with off-axis geometry for quantitative phase microscopy and nanoscopy,” Opt. Express 21(5), 5701–5714 (2013). [CrossRef]   [PubMed]  

20. H. Gabai, M. Baranes-Zeevi, M. Zilberman, and N. T. Shaked, “Continuous wide-field characterization of drug release from skin substitute using off-axis interferometry,” Opt. Lett. 38(16), 3017–3020 (2013). [CrossRef]   [PubMed]  

21. B. Sha, X. Liu, X. L. Ge, and C. S. Guo, “Fast reconstruction of off-axis digital holograms based on digital spatial multiplexing,” Opt. Express 22(19), 23066–23072 (2014). [CrossRef]   [PubMed]  

22. I. Frenklach, P. Girshovitz, and N. T. Shaked, “Off-axis interferometric phase microscopy with tripled imaging area,” Opt. Lett. 39(6), 1525–1528 (2014). [CrossRef]   [PubMed]  

23. C. S. Guo, B. Y. Wang, B. Sha, Y. J. Lu, and M. Y. Xu, “Phase derivative method for reconstruction of slightly off-axis digital holograms,” Opt. Express 22(25), 30553–30558 (2014). [CrossRef]   [PubMed]  

24. S. Karepov, N. T. Shaked, and T. Ellenbogen, “Off-axis interferometer with adjustable fringe contrast based on polarization encoding,” Opt. Lett. 40(10), 2273–2276 (2015). [CrossRef]   [PubMed]  

25. A. Nativ and N. T. Shaked, “Compact interferometric module for full-field interferometric phase microscopy with low spatial coherence illumination,” Opt. Lett. 42(8), 1492–1495 (2017). [CrossRef]   [PubMed]  

26. V. Mico, Z. Zalevsky, and J. García, “Superresolution optical system by common-path interferometry,” Opt. Express 14(12), 5168–5177 (2006). [CrossRef]   [PubMed]  

27. D. Fu, S. Oh, W. Choi, T. Yamauchi, A. Dorn, Z. Yaqoob, R. R. Dasari, and M. S. Feld, “Quantitative DIC microscopy using an off-axis self-interference approach,” Opt. Lett. 35(14), 2370–2372 (2010). [CrossRef]   [PubMed]  

28. Y. Kim, H. Shim, K. Kim, H. Park, J. H. Heo, J. Yoon, C. Choi, S. Jang, and Y. Park, “Common-path diffraction optical tomography for investigation of three-dimensional structures and dynamics of biological cells,” Opt. Express 22(9), 10398–10407 (2014). [CrossRef]   [PubMed]  

29. B. Bhaduri, C. Edwards, H. Pham, R. Zhou, T. H. Nguyen, L. L. Goddard, and G. Popescu, “Diffraction phase microscopy: principles and applications in materials and life sciences,” Adv. Opt. Photonics 6(1), 57–119 (2014). [CrossRef]  

30. X. T. Zhang and C. S. Guo, “Common-path on-axis Fresnel holography based on a pinhole array plate,” Appl. Opt. 54(1), A32–A38 (2015). [CrossRef]   [PubMed]  

31. T. H. Nguyen, C. Edwards, L. L. Goddard, and G. Popescu, “Quantitative phase imaging of weakly scattering objects using partially coherent illumination,” Opt. Express 24(11), 11683–11693 (2016). [CrossRef]   [PubMed]  

32. C. Zheng, R. Zhou, C. Kuang, G. Zhao, Z. Yaqoob, and P. T. C. So, “Digital micromirror device-based common-path quantitative phase imaging,” Opt. Lett. 42(7), 1448–1451 (2017). [CrossRef]   [PubMed]  

33. C. Falldorf, M. Agour, and R. B. Bergmann, “Digital holography and quantitative phase contrast imaging using computational shear interferometry,” Opt. Eng. 54(2), 024110 (2015). [CrossRef]  

34. V. Mico, C. Ferreira, Z. Zalevsky, and J. García, “Spatially-multiplexed interferometric microscopy (SMIM): converting a standard microscope into a holographic one,” Opt. Express 22(12), 14929–14943 (2014). [CrossRef]   [PubMed]  

35. K. B. Seo, B. M. Kim, and E. S. Kim, “Digital holographic microscopy based on a modified lateral shearing interferometer for three-dimensional visual inspection of nanoscale defects on transparent objects,” Nanoscale Res. Lett. 9(1), 471 (2014). [CrossRef]   [PubMed]  

36. B. M. Kim and E. S. Kim, “Visual inspection of 3-D surface and refractive-index profiles of microscopic lenses using a single-arm off-axis holographic interferometer,” Opt. Express 24(10), 10326–10344 (2016). [CrossRef]   [PubMed]  

37. J. A. Picazo-Bueno, Z. Zalevsky, J. García, C. Ferreira, and V. Micó, “Spatially multiplexed interferometric microscopy with partially coherent illumination,” J. Biomed. Opt. 21(10), 106007 (2016). [CrossRef]   [PubMed]  

38. J. Di, Y. Li, M. Xie, J. Zhang, C. Ma, T. Xi, E. Li, and J. Zhao, “Dual-wavelength common-path digital holographic microscopy for quantitative phase imaging based on lateral shearing interferometry,” Appl. Opt. 55(26), 7287–7293 (2016). [CrossRef]   [PubMed]  

39. J. Á. Picazo-Bueno, Z. Zalevsky, J. García, and V. Micó, “Superresolved spatially multiplexed interferometric microscopy,” Opt. Lett. 42(5), 927–930 (2017). [CrossRef]   [PubMed]  

40. B. M. Kim, S. J. Park, and E. S. Kim, “Single-shot digital holographic microscopy with a modified lateral-shearing interferometer based on computational telecentricity,” Opt. Express 25(6), 6151–6168 (2017). [CrossRef]   [PubMed]  

41. S. Rawat, S. Komatsu, A. Markman, A. Anand, and B. Javidi, “Compact and field-portable 3D printed shearing digital holographic microscope for automated cell identification,” Appl. Opt. 56(9), D127–D133 (2017). [CrossRef]   [PubMed]  

42. T. Colomb, F. Dürr, E. Cuche, P. Marquet, H. G. Limberger, R.-P. Salathé, and C. Depeursinge, “Polarization microscopy by use of digital holography: application to optical-fiber birefringence measurements,” Appl. Opt. 44(21), 4461–4469 (2005). [CrossRef]   [PubMed]  

43. D. H. Goldstein, Polarized Light, 3rd ed. (CRC Press, 2010).

44. X. Liu, B. Y. Wang, and C. S. Guo, “One-step Jones matrix polarization holography for extraction of spatially resolved Jones matrix of polarization-sensitive materials,” Opt. Lett. 39(21), 6170–6173 (2014). [CrossRef]   [PubMed]  

45. J. Park, H. Yu, J. H. Park, and Y. Park, “LCD panel characterization by measuring full Jones matrix of individual pixels using polarization-sensitive digital holographic microscopy,” Opt. Express 22(20), 24304–24311 (2014). [CrossRef]   [PubMed]  

46. T. D. Yang, K. Park, Y. G. Kang, K. J. Lee, B. M. Kim, and Y. Choi, “Single-shot digital holographic microscopy for quantifying a spatially-resolved Jones matrix of biological specimens,” Opt. Express 24(25), 29302–29311 (2016). [CrossRef]   [PubMed]  

47. X. Liu, Y. Yang, L. Han, and C. S. Guo, “Fiber-based lensless polarization holography for measuring Jones matrix parameters of polarization-sensitive materials,” Opt. Express 25(7), 7288–7299 (2017). [CrossRef]   [PubMed]  

48. Z. Wang, L. J. Millet, M. U. Gillette, and G. Popescu, “Jones phase microscopy of transparent and anisotropic samples,” Opt. Lett. 33(11), 1270–1272 (2008). [CrossRef]   [PubMed]  

49. Y. Kim, J. Jeong, J. Jang, M. W. Kim, and Y. Park, “Polarization holographic microscopy for extracting spatio-temporally resolved Jones matrix,” Opt. Express 20(9), 9948–9955 (2012). [CrossRef]   [PubMed]  

50. Q. Y. Yue, Z. J. Cheng, L. Han, Y. Yang, and C. S. Guo, “One-shot time-resolved holographic polarization microscopy for imaging laser-induced ultrafast phenomena,” Opt. Express 25(13), 14182–14191 (2017). [CrossRef]   [PubMed]  

51. C. Meneses-Fabian and G. Rodriguez-Zurita, “Carrier fringes in the two-aperture common-path interferometer,” Opt. Lett. 36(5), 642–644 (2011). [CrossRef]   [PubMed]  

52. P. Girshovitz and N. T. Shaked, “Doubling the field of view in off-axis low coherence interferometric imaging,” Light Sci. Appl. 3(3), e151 (2014). [CrossRef]  

53. X. Wang, H. Zhai, and G. Mu, “Pulsed digital holography system recording ultrafast process of the femtosecond order,” Opt. Lett. 31(11), 1636–1638 (2006). [CrossRef]   [PubMed]  

54. M. Paturzo, P. Memmolo, A. Tulino, A. Finizio, and P. Ferraro, “Investigation of angular multiplexing and de-multiplexing of digital holograms recorded in microscope configuration,” Opt. Express 17(11), 8709–8718 (2009). [CrossRef]   [PubMed]  

55. C. Yuan, G. Situ, G. Pedrini, J. Ma, and W. Osten, “Resolution improvement in digital holography by angular and polarization multiplexing,” Appl. Opt. 50(7), B6–B11 (2011). [CrossRef]   [PubMed]  

56. B. Ma, B. Yao, S. Yan, F. Peng, J. Min, M. Lei, and T. Ye, “Simulation and optimization of spatial light modulation of twisted-nematic liquid crystal display,” Chin. Opt. Lett. 8(10), 960–963 (2010). [CrossRef]  

57. D. Amaya, D. Actis, G. Rumi, and A. Lencina, “Least squares method for liquid crystal display characterization,” Appl. Opt. 56(5), 1438–1446 (2017). [CrossRef]  

58. M. Koike-Tani, T. Tani, S. B. Mehta, A. Verma, and R. Oldenbourg, “Polarized light microscopy in reproductive and developmental biology,” Mol. Reprod. Dev. 82(7-8), 548–562 (2015). [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 (6)

Fig. 1
Fig. 1 Schematic of the proposed COPH system
Fig. 2
Fig. 2 A typical example of the spatial spectrum of the DC-AM-PH described by Eq. (6).
Fig. 3
Fig. 3 (a) Photo of the designed input widows attached with the test object. (b) Gray level picture displayed on the measured LCD; scale bar: 1.8 mm. (c) An example of the DC-AM-PHs recorded in experiments; scale bar: 645um.
Fig. 4
Fig. 4 Example of the measurement results. (a)-(d) retrieved amplitude distributions of the four Jones matrix parameters Jxx, Jxy, Jyx and Jyy of the measured LCD; (e)-(h) their respective phase distributions. Scale bar: 1.6 mm.
Fig. 5
Fig. 5 Fitted curves of the Jones parameters versus the gray levels displayed on the LCD. (a) and (b) are, respectively, the amplitude and the phase values versus gray level.
Fig. 6
Fig. 6 Comparison between the calculated amplitude transmittance and the measured results. (a) Amplitude transmittance | t | of the LCD panel when the input and output polarizers are set, respectively, to θ 1 =45° and θ 1 =90°; (b) the result when the polarizer axes are taken as θ 1 =45° and θ 1 =135°.

Equations (10)

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

u 1 (x,y)=rect( x a , y a )[ E x (x,y) E y (x,y) ]+ R 1 rect( xb a , y a )[ 1 1 ]+ R 2 rect( x a , yb a )[ 1 1 ],
u 2 (x,y)= D 0 u 1 ( x M , y M )+ D 1x u 1 ( x+ d g M , y M )exp[j2π ξ g x] + D 1y u 1 ( x M , y+ d g M )exp[j2π ξ g y 0 ],
d g = λ f 2 T ,and  ξ g = Δ f 2 T ,
b= d g M = λ f 1 T ,
u c (x,y)=rect( x Ma , y Ma )[ D 0 J xx ( x M , y M )+ D 1x R 1 exp[j2π ξ g x]+ D 1y R 2 exp[j2π ξ g y] D 0 J yx ( x M , y M )+ D 1x R 1 exp[j2π ξ g x] D 1y R 2 exp[j2π ξ g y] ].
H(x,y)= | D 0 J xx ( x M , y M )+ D 1x R 1 exp[j2π ξ g x]+ D 1y R 2 exp[j2π ξ g y] | 2 + | D 0 J yx ( x M , y M )+ D 1x R 1 exp[j2π ξ g x] D 1y R 2 exp[j2π ξ g y] | 2 , = Y 0 (x,y)+ Y 1x (x,y)+ Y 1y (x,y)+ Y 1x * (x,y)+ Y 1y * (x,y)
Y 0 (x,y)= | D 0 J xx ( x M , y M ) | 2 + | D 0 J yx ( x M , y M ) | 2 +2 | D 1x R 1 | 2 +2 | D 1y R 2 | 2 ,
Y 1x (x,y)= D 0 D 1x R 1 [ J xx ( x M , y M )+ J yx ( x M , y M )]exp[j2π ξ g x] Y 1y (x,y)= D 0 D 1y R 2 [ J xx ( x M , y M ) J yx ( x M , y M )]exp[j2π ξ g y].
J xx ( x M , y M )= 1 2 ( Y 1x (x,y) Y 1x0 (x,y) + Y 1y (x,y) Y 1y0 (x,y) ) J yx ( x M , y M )= 1 2 ( Y 1x (x,y) Y 1x0 (x,y) Y 1y (x,y) Y 1y0 (x,y) ).
t=cos( θ 2 )( J xx cos( θ 1 )+ J xy sin( θ 1 ))+sin( θ 2 )( J yx cos( θ 1 )+ J yy sin( θ 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.