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

Backscatter digital holography of microparticles

Open Access Open Access

Abstract

This work investigates a method for digital holographic imaging of microparticles. Traditional digital holographic techniques use a particle’s forward scattered light to form the hologram, whereas here we use the backscattered light. Images of a particle are then computationally reconstructed from the backscatter hologram, and several examples of such reconstructions are presented. A potential advantage of this technique is that the backscatter holograms may be more sensitive to particle-surface details.

©2013 Optical Society of America

Corrections

22 December 2016: A correction was made to the body text.

1. Introduction

In recent years, digital holography has been developed as a useful technique to image small particles ~10 μmin situ and potentially in near real-time. In particular, the characterization of small aerosol particles is possible with such techniques and has important potential applications, including climate change studies, imaging and tracking, and life sciences [110]. The objective of this article is to introduce a digital holographic microscopy (DHM) concept that uses a particle’s backscattered wave, rather than the forward, or side-scattered, wave that is typically used. Here, the backscattered wave interferes with a reference wave across a sensor-array, which results in an interference pattern that is the backscattered hologram. This hologram then yields an image of the particle once processed using the computational reconstruction routine described in [6]. This DHM arrangement is especially well suited for imaging optically opaque particle types [7].

2. Background

In our set up, a pellicle beam splitter is used to split a laser beam into two parts; one part undergoes transmission and illuminates the particle, while the other acts as the reference wave. The illumination wave backscatters from the particle and interferes with the reference before arriving at a two-dimensional optoelectronic sensor such as a digital camera. During this process both the reference and backscattered waves can be represented by their amplitudes as:

Aref(r)=exp(ikr)rAoref(r^),Asca(r)=exp(ikr)rAosca(r^),
respectively, where ris the position vector, k=2π/λ is the wave number and λ is the wavelength of the spherical wave [4, 5]. The light intensity at the detector when no particle is present is given by
Iref(r)=cεor2|Aoref(r^)|2
where c and εo are the speed of light and electric permittivity in vacuum respectively. Likewise, the intensity when the particle is present is given by
Iholo(r)=cεor2|Aoref(r^)+Aosca(r^)|2.
Subtracting the reference intensity Eq. (2) from the total intensity Eq. (3) gives the so-called contrast hologram as
Icon(r)=cεor2[{[Aoref(r^)]*Aosca(r^)+[Aosca(r^)]*Aoref(r^)}+|Aosca(r^)|2],
where the asterisk denotes the complex conjugate [2].

In Eq. (4), the first two terms are due to the interference between the unscattered and backscattered wave whereas the last term is due to the backscattered wave only. In principle, an image of the particle can be reconstructed from this pattern. In this work the particles block only a small fraction of the illuminating light, and consequently, the intensity of the reference wave dominates the intensity of the backscattered wave. Thus, the last term in Eq. (4) can be neglected, leaving

Icon(r)=cεor2[{[Aoref(r^)]*Aosca(r^)+[Aosca(r^)]*Aoref(r^)}],

By subtracting the reference intensity leading to Eq. (4), imperfections in the incident beam profile are greatly suppressed [5]. The key characteristic of Icon is its linear dependence on the amplitude of the backscattered wave, i.e., the wave phase is encoded in this detected intensity distribution. Equation (5) can then be used with the reconstruction algorithm described in [6] to render an image of the particle.

We note that in the conventional digital in-line holographic arrangement, i.e., that using the forward scattered light, the particle image is accompanied by its out-of-focus twin upon reconstruction [1113]. This twin image is also present in the backscattered arrangement and can potentially degrade its imaging capability. However, the twin can be largely suppressed using the method of subtracting the average value of the hologram intensity from Eq. (4) as described in [6] for a forward scattering arrangement.

3. Experiment

The experimental setup of our backscatter digital holographic (BDH) arrangement is shown in Fig. 1. The arrangement is initially established by constructing a Michelson interferometer so that proper alignment of the beamsplitter (BS) can be achieved. The purpose of this step is to align the two beams (reference and illumination) such that they interfere constructively at the detector when no particle is present.

 figure: Fig. 1

Fig. 1 Diagram of the experimental arrangement used to demonstrate BDH imaging of microparticles. Note that this arrangement shows the beam dump placed in front of M2 following the beam alignment step described below.

Download Full Size | PDF

The light source used in the experiment is a Q-Switched Nd:YLF laser, frequency doubled to 527 nm. At a distance of 5 cm from the laser is an iris to remove stray light from the laser head, and then at 9 cm from the iris we place an achromatic-doublet lens (L1) of focal length 4 cm. Next, at the focus of the lens we use a pin-hole (PH) of diameter 25μm,which produces the circular diffraction pattern. Using another iris, the central diffraction peak is selected and followed by another lens (L2) of focal length 40 cm, which collimates the central diffraction peak. The purpose of this step is to produce a clean transverse beam-profile. Next a pellicle BS of diameter 5.08 cm with a reflection-to-transmission ratio of 45%:55% is used to split the beam into the reference and illumination waves. This BS is 38 cm from L2 and is mounted on a translation stage to ease its alignment. Then, silver mirrors M1 and M2 are each placed at 50 cm from the BS. Here M1 is fixed and M2 is on a linear translation stage. We then adjust the position of M2 until an interference maximum is seen on the detector, which is a CCD camera (Finger Lakes Instrument, Model ML0081909) at a distance of 4 cm from the BS, see Fig. 1.

Following this alignment step, we remove mirror M2 and place a beam dump (BD) at 150 cm from the BS. We then place a glass window (W) on a xyz-translation stage, and another BD in front of M1. Later, this window will hold the particles under investigation. The intensity of the backscattered light from the (particle-free) window is recorded, which provides an estimate for the amount of light scattered by the bare glass. Next, we place a particle on the window, remove the BD from in front of M1, and record the backscattered hologram. By subtracting the particle-free and particle-loaded intensity patterns at the detector, we get the contrast backscattered hologram of Eq. (4).

4. Particle-image reconstruction

To render an image of the particle, we apply the reconstruction algorithm described in [2] and [6] to Icon using the commercial software Mathematica. This process is essentially an application of the Fresnel-Kirchhoff diffraction theory to the backscattered hologram. An image is generated by simulating the diffraction of light through the hologram as though it were a transmission diffraction. The result is a reconstruction of the particle’s backscattered wave amplitude in a plane parallel to the hologram. The absolute-value squared of the wave in this plane, which is called the reconstruction plane, yields a particle image. However, this particle image will be out-of-focus unless the (computational) distance between the hologram and reconstruction planes is selected properly. We determine this distance by incrementally adjusting it until an in-focus image is obtained.

Once an in-focus image is generated, it is then necessary to calibrate the length scales in the reconstructed images. This is done using an optical fiber (Newport Corporation, F-SMF-28) of diameter 245μmas a calibration “particle” with known size. The fiber is placed on the same glass window described in Sec. 3. A backscatter hologram is then recorded, from which the image is reconstructed. Then, the same window, containing the optical fiber, is imaged with a traditional optical microscope to verify that the image reconstruction procedure can adequately render particle-size and shape. Figure 2 presents this microscope/holographic image comparison for the same optical fiber.

 figure: Fig. 2

Fig. 2 Comparison between a traditional microscope image (a) of an optical-fiber and the corresponding image (b) of the same fiber reconstructed from its backscattered hologram. The optical arrangement used is that shown in Fig. 1. Note that white lines have been drawn in (b) to delineate the (low contrast) edges of the fiber.

Download Full Size | PDF

Next, we use the same optical arrangement to image a variety of microparticles from backscattered holograms. These particles, which include ragweed pollen spores (~10 μm dia.), borosilicate glass microspheres (10 μm dia.), and Aspergillus flavus (~5-25μm) spores are deposited on the glass window in Fig. 1. The holographic images of these particles are shown in Fig. 3. One can see from the reconstructed images in Fig. 3 that the resolution is insufficient to image individual ragweed and Aspergillus flavus spore particles, which occur here in many-particle clusters, but can resolve individual microspheres. There are many techniques that could be employed to improve this image resolution, and such is the intent of future work, see e.g., see [2, 7, 14].

 figure: Fig. 3

Fig. 3 These images show the backscatter contrast hologram (left column) measured for various particles along with the reconstructed particle-images (right column). The particles include ragweed pollen spores (a, b), borosilicate glass microspheres (c, d), and Aspergillus flavus spores (e, f).

Download Full Size | PDF

5. Comments

One motivation for this work relates to imaging opaque particles. In the traditional in-line holographic microscope arrangement, i.e., that using the forward-scattered light, an opaque particle yields only its silhouette after image reconstruction. Thus, the particle surface characteristics are obscured. Using backscattered light, however, it may be possible to discern surface textures, or structural features, that are sufficiently larger than the diffraction limit. In this spirit, recent work in [7] achieves backscattered-light holographic microscopy, but is unable to resolve such surface characteristics and the particles under investigation are immobilized at short distances from the (CMOS) sensor array using a slide or microfluidic channel. What our work shows is that the concept in [7] can be extended to investigate particles that cannot be confined to a microfluidic channel or microscope slide, such as would be useful for in situ studies of aerosol particles.

6. Conclusion

This work explores the feasibility of imaging microparticles with digital holography using the backscattered light rather than the more commonly used forward scattered light. We show successful imaging of an optical fiber, ragweed pollen spore-cluster, glass microspheres, and Aspergillus flavus spore-clusters. While the resolution of the imaging technique can be improved, our results demonstrate that backscattered light holographic imaging may be useful for studies of opaque micrometer-sized particles that must be examined in a contact-free manner, such as aerosols.

Acknowledgments

We are thankful to two reviewers for their helpful comments and suggestions. We also thank Ben Ardahl for assistance with electronics instrumentation. This work was supported by the U.S. Army Research Office undergraduate research apprenticeship program.

References and links

1. D. Gabor, “A new microscopic principle,” Nature 161(4098), 777–778 (1948). [CrossRef]   [PubMed]  

2. M. J. Berg and G. Videen, “Digital holographic imaging of aerosol particles in flight,” J. Quant. Spectrosc. Radiat. Transf. 112(11), 1776–1783 (2011). [CrossRef]  

3. M. I. Mishchenko, L. D. Travis, and A. A. Lacis, Scattering, Absorption, and Emission of Light by Small Particles (Cambridge, 2002).

4. V. Mico, C. Ferreira, Z. Zalevsky, and J. Garcia, “Basic principles and application of digital holography microscopy,” (Formatex, 2010) http://www.formatex.info/microscopy4/1411-1418.pdf

5. W. Xu, M. H. Jericho, I. A. Meinertzhagen, and H. J. Kreuzer, “Digital in-line holography of microspheres,” Appl. Opt. 41(25), 5367–5375 (2002). [CrossRef]   [PubMed]  

6. T. Kreis, Handbook of Holographic Interferometry (Wiley, 2005).

7. M. Lee, O. Yaglidere, and A. Ozcan, “Field-portable reflection and transmission microscopy based on lensless holography,” Biomed. Opt. Express 2(9), 2721–2730 (2011). [CrossRef]   [PubMed]  

8. V.R. Singh and G. Hegle,A. Asundi “Particle field imaging using digital in-lineholography,” Current Science 96(3), 391–397(2009)

9. S. Murata and N. Yasuda, “Potential of digital holography in particle measurement,” Opt. Laser Technol. 32(7-8), 567–574 (2000). [CrossRef]  

10. M. K. Kim, “Principles and techniques of digital holographic microscopy,” SPIE Rev. 1(1), 018005 (2010). [CrossRef]  

11. T. Latychevskaia and H. W. Fink, “Solution to the twin image problem in holography,” Phys. Rev. Lett. 98(23), 233901 (2007). [CrossRef]   [PubMed]  

12. I. Yamaguchi and T. Zhang, “Phase-shifting digital holography,” Opt. Lett. 22(16), 1268–1270 (1997). [CrossRef]   [PubMed]  

13. E. Cuche, P. Marquet, and C. Depeursinge, “Spatial filtering for zero-order and twin-image elimination in digital off-axis holography,” Appl. Opt. 39(23), 4070–4075 (2000). [CrossRef]   [PubMed]  

14. Z. Göröcs and A. Ozcan, “On-chip biomedical imaging,” IEEE Rev Biomed Eng 6, 29–46 (2013). [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 (3)

Fig. 1
Fig. 1 Diagram of the experimental arrangement used to demonstrate BDH imaging of microparticles. Note that this arrangement shows the beam dump placed in front of M2 following the beam alignment step described below.
Fig. 2
Fig. 2 Comparison between a traditional microscope image (a) of an optical-fiber and the corresponding image (b) of the same fiber reconstructed from its backscattered hologram. The optical arrangement used is that shown in Fig. 1. Note that white lines have been drawn in (b) to delineate the (low contrast) edges of the fiber.
Fig. 3
Fig. 3 These images show the backscatter contrast hologram (left column) measured for various particles along with the reconstructed particle-images (right column). The particles include ragweed pollen spores (a, b), borosilicate glass microspheres (c, d), and Aspergillus flavus spores (e, f).

Equations (5)

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

A ref ( r ) = exp ( i k r ) r A o ref ( r ^ ) , A sca ( r ) = exp ( i k r ) r A o sca ( r ^ ) ,
I ref ( r ) = c ε o r 2 | A o ref ( r ^ ) | 2
I holo ( r ) = c ε o r 2 | A o ref ( r ^ ) + A o sca ( r ^ ) | 2 .
I con ( r ) = c ε o r 2 [ { [ A o ref ( r ^ ) ] * A o sca ( r ^ ) + [ A o sca ( r ^ ) ] * A o ref ( r ^ ) } + | A o sca ( r ^ ) | 2 ] ,
I con ( r ) = c ε o r 2 [ { [ A o ref ( r ^ ) ] * A o sca ( r ^ ) + [ A o sca ( r ^ ) ] * A o ref ( r ^ ) } ] ,
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.