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

Laser Doppler holographic microscopy in transmission: application to fish embryo imaging

Open Access Open Access

Abstract

We have extended Laser Doppler holographic microscopy to transmission geometry. The technique is validated with living fish embryos imaged by a modified upright bio-microcope. By varying the frequency of the holographic reference beam, and the combination of frames used to calculate the hologram, multimodal imaging has been performed. Doppler images of the blood vessels for different Doppler shifts, images where the flow direction is coded in RGB colors or movies showing blood cells individual motion have been obtained as well. The ability to select the Fourier space zone that is used to calculate the signal, makes the method quantitative.

© 2014 Optical Society of America

1. Introduction

Blood flow imaging techniques are widely used in biomedical studies, since they can assess physiological processes or can be used for early detection of disease [13]. However, many blood flow studies require, for imaging purposes, the use of a contrast agent, making the blood flow characterization invasive [4, 5]. Scanning Doppler imaging techniques can be considered to alleviate this issue, but due to the scanning step, acquisition of an image is a time consuming process [6]. An overview of the main techniques has been proposed by J. D. Briers [7]. Among these, one can outline two ways of monitoring the blood flow: performing measurements either in spatial domain (speckle analysis) or in temporal domain (Doppler analysis). Blood flow monitoring has been demonstrated by Laser Speckle Contrast Analysis/Imaging (LASCA/LSCI) [8]. Here, spatial statistics of the dynamic speckle are used to obtain blood flow images [9, 10]. Improvement of the acquired constrast image have been achieved through exposure time optimization [11], or intensity fluctuation analysis [12] resulting in high quality perfusion images. However, these techniques are only limited to perfusion monitoring and do not allow to assess other quantities such as amplitude or phase contrasts.

Originally proposed by Gabor [13] as an optimization of electron microscopy, classical holography aims at recording, on a photo-sensitive material, the interference between a reference field and a diffracted object field. However, within the proposed common-path interferometric configuration, holographic imaging suffer from the so-called twin-image noise [14]. In 1962, Leith and Upatnieks propose to introduce an off-axis reference field, which lead to a separation of the object and its twin image contribution in spatial frequencies domain [15]. Development and democratization of high resolution CCD/CMOS sensors paid a major role in holographic imaging spreading out, making possible both digital recording and processing of holographic images [16]. Therefore, due to its intrinsic properties, digital holographic imaging is used in a wide variety of studies in fluid mechanics [1721], biomedical imaging and microscopy [2226], or mechanical inspection [2729].

Further development led to the introduction of heterodyne digital holography [30]. Here, the reference field is dynamically phase shifted with respect to the object field. Therefore, the recorded hologram is time modulated, thus enabling phase-shifted interferometric measurements [31]. Moreover, this data acquisition scheme as been demonstrated to be shot-noise limited [32, 33]. Temporal modulation feature of digital heterodyne holography can also be used to investigate dynamic phenomena, and being considered as a laser Doppler imaging technique [34]. The ability of heterodyne digital holography to perform Doppler imaging has been demonstrated in various domains such as microfluidics [35], vibration motion characterization [3638], in vivo vasculature assessment, without contrast agent [3941] and motion of biological objects [4244].

In this paper, we combine laser Doppler holography [34] and transmission microscopy to analyze blood flow in fish embryos. We have adapted a laser Doppler holographic setup to a standard bio-microscope by carrying the two beams of the holographic interferometer (illumination of the object and reference), whose frequency offset is controlled, by optical fibers. Multimodal acquisition and analysis of the data is made by acting on the frequency offset of the two beams, and on the location of the Fourier space filtered zone. With the same set of data, we have imaged, with amplitude contrast, the whole fish embryo, or only the moving blood vessels. In that last case, we have obtained images where the flow direction is coded in RGB color. For large vessels, the dependance of the holographic signal with the frequency offset yields the Doppler frequency shift, and thus the flow velocity. For small vessels, individual Red Blood Cells (RBC) can be imaged, and movies showing the RBC motion are obtained.

2. Materials

2.1. Heterodyne holographic microscopy setup

Our off-axis digital holography set-up is a Mach-Zehnder interferometer in which the recombining cube beam splitter is angularly tilted. To fit our set-up into regular upright microscope, the interferometer is split into an injection (a) and a recombination (b) part, which are connected together with optical fibers (Fig. 1).

 figure: Fig. 1

Fig. 1 Heterodyne digital holographic microscopy experimental arrangement (a) Injection part of the interferometer. HWP: half wave plate; PBS: polarizing beam splitter; AOM1, AOM2: acousto optic modulators (Bragg cells). (b) Classical upright microscope used for the off-axis recombinaison of reference and object beams. BS: cube beam splitter; CCD: CCD camera.

Download Full Size | PDF

The injection part (Fig. 1(a)) is mounted on a separate breadboard, and is used to build both object illumination (field EI) and reference (or local oscillator (LO) ELO) beams. Laser light is provided by a 785 nm, 80 mW Sanyo® DL7140-201S continuous mono-mode laser diode (angular frequency ωI). Laser beam is split by a polarizing beam splitter (PBS), which in combination with the first half wave plate (HWP) allows to adjust the optical power in both reference and object arms of the interferometer. To enable phase shifting interferometry and frequency scanning of the holographic detection frequency, we control the frequency ωLO of the reference beam by using the heterodyne holography method [30]. The reference arm is thus frequency shifted by using two acousto-optic modulators AOM1 and AOM2 (80 MHz AA electronics®), operating at ω1 and ω2 respectively, so that ωLO = ωI + Δω with Δω = ω2ω1. Reference and object fields are finally injected into two mono-mode optical fibers (Thorlabs® P1-780A-FC-2 fiber patch).

Both reference (ELO) and object illumination (EI) fibers are brought to an upright microscope (Olympus® CX41) working in transmission configuration (Fig. 1(b)). The object is imaged by an Olympus DPlan microscope objective (MO: NA= 0.25, G=10, corrected at 160 mm). The illumination field (EI) is scattered by the studied living object (the zebrafish embryo) yielding the object signal field (E). Signal (E) and reference field (ELO) are combined together by a cube beam splitter (BS) that is angularly tilted thus introducing an off-axis angle in the interferometric arrangement. Interferences (i.e. E + ejΔωtELO) between both fields are recorded on a 1360×1024 pixel (6.45 μm square pitch) 12-bits CCD camera (Jenoptik ProgRes MF®) operating at a framerate of ωS/(2π) ≤ 10 Hz. Recorded data are cropped to 1024 × 1024 and FFT (Fast Fourier Transform) are made on a 1024 × 1024 calculation grid. The CCD to MO distance is such that the effective MO enlargement factor is G′ = 10.4 (instead of G = 10). In order to perfectly control the phase and frequency of the ELO reference beam, the CCD camera is triggered at frequency ωCCD, and the ω1, ω2 and ωCCD signals are generated by digital synthesizers driven by a common 10 MHz clock.

Zebrafish (Danio rerio, wild type AB line) were maintained according to standard protocols [45]. Larvae were mounted on a standard glass slide under a 22 mm 1.5 round coverslip with a 0.5 mm thick caoutchouc ring as a spacer and sealant. They were embeded in a drop of cooling 1.5% low melting point agarose at 37°C and oriented before gelation. The chamber was filled with 100 mg/l tricaine in filtered tank water and imaging was performed at room temperature. Blood vessels nomenclature is according to [46].

3. Laser Doppler holography mode

Due to motions within the sample (mainly the RBC motion within the embryo vasculature), the incident field EI (of wave vector kI) that is scattered to wave vector kS (yielding ES) undergoes a Doppler shift

ωSωI=q.v
where v is the velocity of the scatterer and q = kSkI the scattering vector (see Fig. 2(a)) and q.v the inner product of vectors q and v. As detection is made for all wave vectors kS within the MO collection cone, the frequency spectrum of the moving scatterers signal is broadened. One then gets two components: a sharp narrow peak from the photons that are not scattered (ballistic photons) or scattered by objects that do not move, and a broad pedestal from the photons that are scattered by moving objects and that are collected by MO. The width of the pedestal is proportional to the MO numerical aperture NA since |q| ∼ NA|kI|. With NA = 0.25, the Doppler broadening we observe here is much smaller (10 to 100 Hz) than the one (100 to 1000 Hz) observed in previous in vivo Doppler holographic experiments made in reflection configuration [41, 4749], where |q| ∼ 2|kI|. Since our camera operates at ωS/(2π) = 10 Hz, the bandwidth of our holographic detection (±5 Hz around ωLO/(2π)) in not enough to acquire the moving scatterer Doppler signal. As in previous works, the frequency ωLO = ωI + Δω (with Δω = ω1ω2) is swept in order to acquire the whole Doppler signal and to analyze its spectrum.

 figure: Fig. 2

Fig. 2 Detection of zebrafish embryo blood flow in transmission geometry. (a) Illustration of the spectral broadening of light encountering moving scattering objects (here, red blood cells). (b) Principles of the spectral measurement by heterodyne digital holography.

Download Full Size | PDF

4. Experimental results

The holographic information is acquired by sweeping the frequency shift Δω/(2π) from 0 up to ∼ 100 Hz by step of ωCCD/2π (i.e. Δω = CCD with m integer) except in section 4.1 where Δω = ωCCD/4. On the other hand, the movies Media 1 and Media 2 have been recorded with Δω = 0. To optimize sensitivity and selectivity, the illumination beam versus local oscillator ratio is adjusted in order to have |ELO|2 > |E|2, with |E|2 as large as possible. Exposure time and frame rate are Texp = 50 ms and ωCCD/(2π) = 1/TCCD = 10 Hz. For each frequency shift Δω/(2π), sequences of N = 32 (or N = 96) CCD camera frames (i.e. I0, I1...IN−1) are recorded.

4.1. Spatial filtering and reconstruction

The reconstruction procedure is similar to the one used in previous works [50, 51]. To illustrate reconstruction, let us consider 4 phase detection of the ballistic peak signal (no Doppler shift). We have thus adjusted ω1 and ω2 to have Δω = ωCCD/4, and calculated the demodulated hologram by:

H(x,y)=[I0(x,y)I2(x,y)]+j[I1(x,y)I3(x,y)]
Figure 3(a) shows the holographic signal |H(x, y)|2 displayed in log scale. To select the +1 holographic grating order by the Cuche et al. method [52], we have performed the holographic reconstruction of the MO pupil by the Schnars et al. method [16] yielding H1(kx, ky) with:
H1(kx,ky)=FFT[H(x,y)ej|k|(x2+y2)/2d]
where FFT is the Fast Fourier transform operator, and ej|k|(x2+y2)/2d the phase factor that makes the pupil image sharp (see Fig. 3(b)). For a plane wave reference, the parameter d is equal to the MO pupil to CCD camera distance. Optimal spatial filtering is then made by cropping the MO pupil disk on its sharp circular edge, and by moving the cropped zone in the center of the Fourier space (see Fig. 3(c)) yielding H2(kx, ky). The reconstructed hologram H3(x, y, z) is then calculated by the angular spectrum method that involves 2 FFT from FFT−1[H2(kx, ky)]. We have thus:
H3(x,y,z)=FFT1[H2(kx,ky)ej(kx2+ky2)/2z]
where ej(kx2+ky2)/2z is the phase factor that describe the propagation of the field from the CCD plane z = 0 to the location z of the zebrafish image (via MO). In most cases, the embryo is in focus on the camera, and z = 0. Here, the selection of the +1 grating order is made by the crop, the compensation of the MO phase curvature by the ej|k|(x2+y2)/2d phase factor, and the compensation of the off axis tilt by the translation of the cropped zone.

 figure: Fig. 3

Fig. 3 Spatial filtering and reconstruction principle. Holograms H(x, y) (a), H1(kx, ky) (b), H2(kx, ky) (c) and H3(x, y, z = 0) (d): see Media 1. The display is made in arbitrary Log scale for the average intensity 〈|HX|2〉. This average is calculated made by making the reconstruction with I0..I3, with I1..I4... and with I28..I31, and by averaging the resulting |HX|2. Images (a) and (d) show a side view of a 6 days old zebrafish embryo. Ventral side in on the right and anterior is up.

Download Full Size | PDF

We have calculated the so called instantaneous intensity signal |H3|2 by using data of successive sequences of 4 frames, i.e. by calculating |H3|2 with I0..I3, then with I1..I4... and to the end with I28..I31). The movie of the instantaneous intensity signal |H3|2 ( Media 1) shows that the time varying components of |H3(t)|2 are extremely low. We have then averaged |H3(t)|2 over the sequence of n frames, getting 〈|H3|2〉 that is displayed on Fig. 3(d).

The images show a side view of the zebrafish embryo. Due to the use of coherent light illumination, the image exhibits some speckle noise but anatomical details such as the notochord, somite boundaries and pigment cells are distinguished. In the corresponding movie Media 1, motion of blood can be barely seen in the caudal artery and vein but none is observe in the smaller capillaries.

4.2. Holographic Doppler images of the moving scatterers

To select the moving scatterer signal that corresponds to the Doppler pedestal, we have swept the detection frequency shift Δω by step ωCCD, and calculated 2 phase holograms from the recorded data. We have thus:

Δω=ωLOωI=mωCCDH(x,y)=I0(x,y)I1(x,y)
where m is integer. Note that H(x, y) is defined by Eq. 2 for 4 phase holograms, and by Eq. 5 for 2 phase holograms. We have kept the same notation (i.e. H(x, y)), because the reconstruction equations Eq. 3 and Eq. 4 are the same in both cases.

We have calculated the 2 phase holographic detection efficiency (i.e. η for the field and |η|2 for the intensity) as a function of the frequency ωS of the scattered field. This calculation is similar to the one done in previous works [31, 32]. We get:

η(x)=12Texpk=01(1)kt=kTCCDTexp/2kTCCD+Texp/2ej2πxtdt=12sinc(πxTexp)[1ej2πxTCCD]
where TCCD = 2π/ωCCD is the frame time interval, and x = (ωSωLO)/2π the LO versus signal frequency offset. The efficiency η2 is plotted on Fig. 4. As seen, the 2 phase detection selects scattered field components with x close to zero (i.e. with ωSωLO). Moreover, since |η|2 is null for x = −20, −10, 0, 10, 20... Hz, the holographic detection efficiency |η|2 remains null at the illumination frequency ωI, whatever the frequency shift Δω = CCD is. The ballistic peak (whose frequency is ωI), which is much bigger than the Doppler pedestal, is thus not seen during the sweep.

 figure: Fig. 4

Fig. 4 Two phase detection efficiency |η|2 as a function of the LO versus signal frequency offset: x. Calculation is made by Eq. 6 with TCCD = 100 ms and Texp = 50 ms.

Download Full Size | PDF

Figure 5 shows the Doppler average reconstructed holograms 〈|H3(x, y)|2〉 obtained with frequency shifts Δω/(2π) varying from 0 Hz (a) to 80 Hz (i) by step of 10 Hz. The recorded hologram H(x, y), and the reconstructed intensity hologram |H3(x, y)|2 are calculated from successive sequence of 2 frames (In and In+1) within the sequence of N = 32 frames (with N = 0 to 30), the displayed signal 〈|H3(x, y)|2〉 corresponding to the average of |H3(x, y)|2 over n. Since the ballistic peak is not seen, the contrast of the Fig. 5 images is reversed with respect to Fig. 3 (d), and vessels where blood flows are therefore seen in white on a black background. In image (a) caudal vein (A) and caudal artery (B,C) are seen as well as intersegmental vessels (D, E) and dorsal longitudinal anastomotic vessels. An abnormal shunt between caudal artery and vein in seen near arrow (C).

 figure: Fig. 5

Fig. 5 Reconstructed hologram H3(x, y, z = 0) made for (ωLOωI)/(2π) = 0 Hz (a): see also Media 2, 10 Hz (b) ... to 80 Hz (i). Images (a) to (i) are displayed with the same Log scale for the average intensity 〈|H3|2〉. The sample is the same as the one in Fig. 3.

Download Full Size | PDF

The reconstructed images (a) to (i) are displayed with the same Log scale for the average intensity 〈|H3|2〉. One thus clearly see the decrease of the holographic signal with the frequency shift (ωLOωI)/(2π) that is expected. Figure 6 analyzes this decreases for the zones of the reconstructed image marked by arrows (A) to (F) in Fig. 5(a). All curves are calculated by applying on the holographic signal 〈|H3|2〉 a gaussian low pass filter (of 10 pixels width for curves A,B,C and F curves, and 4 pixels width for curves D and E), and by summing the filtered holographic signal over 8 × 8 pixels. On curve (A), which corresponds to the flow in caudal vein, the holographic signal decreases regularly with the frequency offset Δω. The same occurs in (B) and (C), but the decrease is slower showing faster velocities for the blood cells in the caudal artery.

 figure: Fig. 6

Fig. 6 Dependance of the Doppler holographic signal 〈|H3(x, y)|2〉 with the frequency off-set (ωLOωI)/(2π) for different location A to F of the reconstructed image of Fig. 3(a). Curves are drawn with linear (a) and logarithmic (b) scales.

Download Full Size | PDF

In intersegmental vessels (curves D and E), the signal 〈|H3|2〉 is much lower than the one of larger artery and vein. The corresponding frequency spectrum (Fig. 6(d) and (e)) exhibit a decrease of the signal at Δω/(2π) = 40 and 80 Hz offset followed by an increase at 60 and 100 Hz. This might be related to the instrumental response of our holographic detection [31]. A fine analysis shows that the signal fluctuates a lot within each sequence of N frames, and from one sequence to the next, since the flowing blood cells can be visualized individually on the instantaneous intensity signal |H3(t)|2 reconstructed with successive sequence of 2 frames. Movie Media 2 shows |H3(t)|2 for the sequence of N = 32 frames recorded at 0 Hz (i.e. without frequency offset) that has been used to display 〈|H3|2〉 on Fig. 5(a). Movie Media 3 shows |H3(t)|2 for a sequence of N = 96 frames that image the embryo shown in Fig. 8(h) and (i). The RBCs individual motion in intersegmental vessels is clearly seen on Media 2 and Media 3. These two movies are very similar to those seen in reference [53]. They can be used to measure the RBC velocity by Particule Image Velocimetry (PIV).

 figure: Fig. 7

Fig. 7 Fourier space reconstructed hologram H2(kx, ky) made without (a) and with (b,c) selection of the scattered wave vector kS. In (b) the selected zone is oriented toward uφ=0. In (c), three zones oriented along uφ with φ = 0 (blue), 2π/3 (green) and 4π/3 (red) are selected. Display is made in arbitrary log scale for 〈|H2|2〉. Frequency shift is (ωLOωI) = 0.

Download Full Size | PDF

4.3. RGB holographic Doppler images yielding direction of motion

The holographic experiment made in transmission configuration gives here a Doppler signal corresponding to single scattering events like in Dynamic Light Scattering (DLS) experiments. The Doppler shift is thus simply q.v where q = kSkI. The illumination (or incident) wave vector kI is fixed by the geometry of the experimental setup. In transmission geometry, kI is parallel to uz (where ux, uy and uz are the unit vectors in the x, y and z directions). On the other hand, kS = (kx, ky, kz) with |kS| = 2π/λ is measured by our holographic experiment. In the reconstruction procedure, the cropped and translated zone seen on Fig. 3(c) and on Fig 7(a) is a map of the holographic signal as a function of kx and ky.

It is then possible to select kS by spatial filtering in the Fourier space, and to reconstruct images corresponding to given intervals of kS. To illustrate this idea, we have selected three kS zones covering half of the Fourier space. These zones are oriented toward uφ=0, uφ=2π/3 and uφ=4π/3 with uφ = cosφ ux + sinφ uy. With the holographic signal H2(kx, ky) of each zone, we have reconstructed a Red, Green or Blue colored image 〈|H3(x, y, z)|2〉, and we have combined these three images to get RGB images sensitive to the direction of kS (and thus to the direction of q and v). Figure 7 (b) shows the |H2(kx, ky)|2 signal, filtered by the φ = 0 spatial filter, which corresponds to the blue image. Figure 7 (c) shows the Fourier space RGB image of the |H2(kx, ky)|2 signal obtained by combining the three colored images corresponding to the φ = 0, φ = 2π/3 and φ = 4π/3 filters.

One must notice here, that the wave vector of ballistic light remains parallel to uz yielding a bright peak near the center of the Fourier space (kx, ky ≃ 0). Although most of the ballistic light is removed in 2 phase holograms recorded without frequency offset, this peak still remains visible in that case as seen on Fig. 7(a). To fully eliminate the ballistic light, the center of the Fourier space (i.e. kx, ky ≃ 0) has thus been removed from the selected zones as shown on Figure 7(b) and (c).

Figure 8(a) to (g) shows the RGB colored reconstructed image of the zebrafish sample obtained with the holographic data of Fig. 5(a) to (g). Without frequency offset (Fig. 8(a) with Δω = 0), the RGB colored reconstructed image is very similar to the black and white one (Fig. 5(a)), but with lower resolution, since half of the Fourier space information is used in the reconstruction. The images of Fig. 8 are obtained from a sequence of 32 images (i.e. with an acquisition time of 3.2 s). The reconstruction is made on GPU in less than one second for the 32 images. Since the detection shift Δω is null, the detection efficiency does not depend on the sign of the Doppler shift of the signal ωSωI. Detection is not sensitive to the sign of q.v, nor to the v direction (even if q direction is selected in the Fourier space). The colors in the reconstructed image of Fig. 8(a) are thus the same whatever the motion direction is. With increasing frequency shifts Δω, signal decreases but colors change, depending on the flow direction. On figure 8(d), with Δω = 30 Hz, caudal artery appears in cyan, while caudal vein, where blood flows in opposite direction, is seen in orange. The flow direction corresponds here to the Fig. 8 color wheel. This wheel corresponds to the colored image of the Fourier space filtered signal |H2(kx, ky)|2 of Fig. 7(d) rotated by 180°.

 figure: Fig. 8

Fig. 8 Colored reconstructed hologram H3(x, y, z = 0) made for (ωLOωI)/(2π) = 0 Hz (a), 10 Hz (b) ... to 60 Hz (g), and for −20 Hz (h) and + 20 Hz (i). Images (a) to (g) are made with same zebrafish sample and same viewpoint as Fig. 3 and Fig. 5. Images (h) to (i) correspond to another fish embryo. Images (a) to (g) (and (h) and (i)) are displayed with the same color RGB Log scale for the average intensity 〈|H3|2〉. Media 3 corresponds to images (h) and (i).

Download Full Size | PDF

To verify that the RGB colored images give a realistic information on the direction of the velocity v, we have recorded, with another embryo, holograms for both sign of the frequency offset Δω. Figure 8(h) and (i) shows the reconstructed RGB images obtained for Δω/(2π) = ±20 Hz. Here again, the colors of the reconstructed image depend on the flow direction, but the color coding is reversed for negative frequency offsets: Δω < 0. The correspondence of the motion direction with the color wheel is only valid for positive shifts Δω > 0.

5. Conclusion

In this paper, we have shown that an upright commercial bio-microscope can be transformed into a powerful holographic setup, by coupling the microscope to an heterodyne holographic [30] breadboard with optical fibers. By controlling the frequency offset Δω of the reference beam, and by using different combinations of frames to calculate the hologram H, we have imaged a living zebrafish embryo under different modalities. With Δω = 0 and H = (I0I2) + j(I1I3), we have calculated the amplitude and the phase of the transmitted light. Although the phase can be qualitatively explored, we have not displayed phase images because the phase variation is too fast (the thickness of the embryo is hundreds of microns).

With Δω = CCD (where m ≠ 0 is integer) and H = (I0I1), we have selected the signal whose Doppler shift ωSωI is close to Δω according to Eq. 4 and Fig. 4) and imaged the blood flow. With Δω = 0 and H = (I0I1), we have visualized individual moving bloods cells in intersegmental vessels, getting movies that show their instantaneous motion.

Since the scattering is moderate like in DLS (Dynamic Light Scattering) experiment, the Doppler shift is directly related to the scatterers (i.e. blood cells) velocity v by ωSωI = q.v with q = kSkI. Since kIuz is known, and since kS can be selected in the Fourier space, the scatterers velocity v can be measured quantitatively by a proper choice of both the frequency offset Δω and the Fourier space selected zone. To illustrate one of the numerous possibilities of Doppler holography in transmission, we have imaged the blood cell motion direction in RGB colors.

In the present paper, we have not used the possibilities for numerical refocusing to image different z layers of our sample. This could be useful, since all blood vessels are not in the same z plane. For instance, in Fig. 5 and Fig. 8(a)–8(g), the caudal artery is roughly on focus, while the caudal vein is out of focus by about 100 μm. Refocusing could thus be used to reconstruct the embryo vascular system in 3D.

We intend to further improve the method by using a faster camera (to better analyze the motion of individual blood cells) and a higher numerical aperture objective (for better optical sectioning).

Acknowledgments

We acknowledge OSEO-ISI Datadiag grant and ANR Blanc Simi 10 (n° 11 BS10 015 02) grant for funding, and Rémy Vialla for technical support.

References and links

1. E. Friedman, S. Krupsky, A. Lane, S. Oak, E. Friedman, K. Egan, and E. Gragoudas, “Ocular blood flow velocity in age-related macular degeneration,” Ophthalmology 102, 640–646 (1995). [CrossRef]   [PubMed]  

2. M. P. Pase, N. A. Grima, C. K. Stough, A. Scholey, and A. Pipingas, “Cardiovascular disease risk and cerebral blood flow velocity,” Stroke 43, 2803–2805 (2012). [CrossRef]   [PubMed]  

3. J. Kur, E. A. Newman, and T. Chan-Ling, “Cellular and physiological mechanisms underlying blood flow regulation in the retina and choroid in health and disease,” Prog. Retin. Eye Res. 31, 377–406 (2012). [CrossRef]   [PubMed]  

4. O. Sakurada, C. Kennedy, J. Jehle, J. Brown, G. L. Carbin, and L. Sokoloff, “Measurement of local cerebral blood flow with iodo [14c] antipyrine,” Am. J. Physiol.-Heart C. 234, H59–H66 (1978).

5. I. Kanno, H. Iida, S. Miura, M. Murakami, K. Takahashi, H. Sasaki, A. Inugami, F. Shishido, and K. Uemura, “A system for cerebral blood flow measurement using an h215o autoradiographic method and positron emission tomography,” J Cerebr. Blood F. Met. 7, 143–153 (1987). [CrossRef]  

6. Y. Yeh and H. Cummins, “Localized fluid flow measurements with an he-ne laser spectrometer,” Appl. Phys. Lett. 4, 176–178 (1964). [CrossRef]  

7. J. D. Briers, “Laser doppler, speckle and related techniques for blood perfusion mapping and imaging,” Physiol. Meas. 22, R35–R66 (2001). [CrossRef]  

8. J. D. Briers and S. Webster, “Laser speckle contrast analysis (lasca): a nonscanning, full-field technique for monitoring capillary blood flow,” J. Biomed. Opt. 1, 174–179 (1996). [CrossRef]   [PubMed]  

9. A. K. Dunn, “Laser speckle contrast imaging of cerebral blood flow,” Ann. Biomed. Eng. 40, 367–377 (2012). [CrossRef]  

10. D. Briers, D. D. Duncan, E. Hirst, S. J. Kirkpatrick, M. Larsson, W. Steenbergen, T. Stromberg, and O. B. Thompson, “Laser speckle contrast imaging: theoretical and practical limitations,” J. Biomed. Opt. 18, 066018 (2013). [CrossRef]   [PubMed]  

11. S. Yuan, A. Devor, D. A. Boas, and A. K. Dunn, “Determination of optimal exposure time for imaging of blood flow changes with laser speckle contrast imaging,” Appl. Opt. 44, 1823–1830 (2005). [CrossRef]   [PubMed]  

12. Y. Zeng, M. Wang, G. Feng, X. Liang, and G. Yang, “Laser speckle imaging based on intensity fluctuation modulation,” Opt. Lett. 38, 1313–1315 (2013). [CrossRef]   [PubMed]  

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

14. D. Gabor, “Microscopy by reconstructed wave-fronts,” Proc. R. Soc. A 197, 454–487 (1949). [CrossRef]  

15. E. N. Leith and J. Upatnieks, “Reconstructed wavefronts and communication theory,” J. Opt. Soc. Am. A 52, 1123–1128 (1962). [CrossRef]  

16. U. Schnars and W. Jüptner, “Direct recording of holograms by a ccd target and numerical reconstruction,” Appl. Opt. 33, 179–181 (1994). [CrossRef]   [PubMed]  

17. A. Lozano, J. Kostas, and J. Soria, “Use of holography in particle image velocimetry measurements of a swirling flow,” Exp. Fluids 27, 251–261 (1999). [CrossRef]  

18. Y. Pu and H. Meng, “Four-dimensional dynamic flow measurement by holographic particle image velocimetry,” Appl. Opt. 44, 7697–7708 (2005). [CrossRef]   [PubMed]  

19. J.-M. Desse, P. Picart, and P. Tankam, “Digital three-color holographic interferometry for flow analysis,” Opt. Express 16, 5471–5480 (2008). [CrossRef]   [PubMed]  

20. N. Verrier, S. Coëtmellec, M. Brunel, and D. Lebrun, “Digital in-line holography in thick optical systems: application to visualization in pipes,” Appl. Opt. 47, 4147–4157 (2008). [CrossRef]   [PubMed]  

21. N. Verrier, C. Remacha, M. Brunel, D. Lebrun, and S. Coëtmellec, “Micropipe flow visualization using digital in-line holographic microscopy,” Opt. Express 18, 7807–7819 (2010). [CrossRef]   [PubMed]  

22. F. Charrière, N. Pavillon, T. Colomb, C. Depeursinge, T. J. Heger, E. A. Mitchell, P. Marquet, and B. Rappaz, “Living specimen tomography by digital holographic microscopy: morphometry of testate amoeba,” Opt. Express 14, 7005–7013 (2006). [CrossRef]   [PubMed]  

23. W. Xu, M. Jericho, I. Meinertzhagen, and H. Kreuzer, “Digital in-line holography for biological applications,” Proc. Natl. Acad. Sci. USA 98, 11301–11305 (2001). [CrossRef]   [PubMed]  

24. B. Kemper and G. von Bally, “Digital holographic microscopy for live cell applications and technical inspection,” Appl. Opt. 47, A52–A61 (2008). [CrossRef]   [PubMed]  

25. M. K. Kim, “Principles and techniques of digital holographic microscopy,” SPIE Reviews 1, 018005 (2010).

26. K. Lee, K. Kim, J. Jung, J. Heo, S. Cho, S. Lee, G. Chang, Y. Jo, H. Park, and Y. Park, “Quantitative phase imaging techniques for the study of cell pathophysiology: from principles to applications,” Sensors 13, 4170–4191 (2013). [CrossRef]   [PubMed]  

27. P. Picart, J. Leval, D. Mounier, and S. Gougeon, “Some opportunities for vibration analysis with time averaging in digital fresnel holography,” Appl. Opt. 44, 337–343 (2005). [CrossRef]   [PubMed]  

28. J. Leval, P. Picart, J. P. Boileau, and J. C. Pascal, “Full-field vibrometry with digital fresnel holography,” Appl. Opt. 44, 5763–5772 (2005). [CrossRef]   [PubMed]  

29. A. AsundiV. R. Singh, et al., “Time-averaged in-line digital holographic interferometry for vibration analysis,” Appl. Opt. 45, 2391–2395 (2006). [CrossRef]   [PubMed]  

30. F. Le Clerc, L. Collot, and M. Gross, “Numerical heterodyne holography with two-dimensional photodetector arrays,” Opt. Lett. 25, 716–718 (2000). [CrossRef]  

31. M. Atlan, M. Gross, and E. Absil, “Accurate phase-shifting digital interferometry,” Opt. Lett. 32, 1456–1458 (2007). [CrossRef]   [PubMed]  

32. F. Verpillat, F. Joud, M. Atlan, and M. Gross, “Digital holography at shot noise level,” J. DisplayTechnol. 6, 455–464 (2010).

33. M. Lesaffre, N. Verrier, and M. Gross, “Noise and signal scaling factors in digital holography in weak illumination: relationship with shot noise,” Appl. Opt. 52, A81–A91 (2013). [CrossRef]   [PubMed]  

34. M. Atlan and M. Gross, “Laser doppler imaging, revisited,” Rev. of Sci. Instrum. 77, 116103 (2006). [CrossRef]  

35. M. Atlan, M. Gross, and J. Leng, “Laser doppler imaging of microflow,” J. Europ. Opt. Soc. Rapid pub. 1, 06025 (2006). [CrossRef]  

36. F. Verpillat, F. Joud, M. Atlan, and M. Gross, “Imaging velocities of a vibrating object by stroboscopic sideband holography,” Opt. Express 20, 22860–22871 (2012). [CrossRef]   [PubMed]  

37. N. Verrier and M. Atlan, “Absolute measurement of small-amplitude vibrations by time-averaged heterodyne holography with a dual local oscillator,” Opt. Lett. 38, 739–741 (2013). [CrossRef]   [PubMed]  

38. N. Verrier, M. Gross, and M. Atlan, “Phase-resolved heterodyne holographic vibrometry with a strobe local oscillator,” Opt. Lett. 38, 377–379 (2013). [CrossRef]   [PubMed]  

39. M. Atlan, M. Gross, B. C. Forget, T. Vitalis, A. Rancillac, and A. K. Dunn, “Frequency-domain wide-field laser doppler in vivo imaging,” Opt. Lett. 31, 2762–2764 (2006). [CrossRef]   [PubMed]  

40. M. Atlan, B. C. Forget, A. C. Boccara, T. Vitalis, A. Rancillac, A. K. Dunn, and M. Gross, “Cortical blood flow assessment with frequency-domain laser doppler microscopy,” J. Biomed. Opt. 12, 024019 (2007). [CrossRef]   [PubMed]  

41. M. Simonutti, M. Paques, J.-A. Sahel, M. Gross, B. Samson, C. Magnain, and M. Atlan, “Holographic laser doppler ophthalmoscopy,” Opt. Lett. 35, 1941–1943 (2010). [CrossRef]   [PubMed]  

42. C. Fang-Yen, S. Oh, Y. Park, W. Choi, S. Song, H. S. Seung, R. R. Dasari, and M. S. Feld, “Imaging voltage-dependent cell motions with heterodyne mach-zehnder phase microscopy,” Opt. Lett. 32, 1572–1574 (2007). [CrossRef]   [PubMed]  

43. H. Iwai, T. Yamauchi, M. Miwa, and Y. Yamashita, “Doppler-spectrally encoded imaging of translational objects,” Opt. Commmun. 319, 159–169 (2014). [CrossRef]  

44. S. Joseph, J.-M. Gineste, M. Whelan, and D. Newport, “A heterodyne mach-zehnder interferometer employing static and dynamic phase demodulation techniques for live-cell imaging,” Proc. SPIE 7554, 75540P (2010). [CrossRef]  

45. M. Westerfield, The Zebrafish Book: a Guide for the Laboratory use of Zebrafish (Danio rerio) (Institute of Neuroscience. University of Oregon, 1995).

46. S. Isogai, M. Horiguchi, and B. M. Weinstein, “The vascular anatomy of the developing zebrafish: an atlas of embryonic and early larval development,” Dev. Biol. 230, 278–301 (2001). [CrossRef]   [PubMed]  

47. M. Atlan, M. Gross, B. C. Forget, T. Vitalis, A. Rancillac, and A. K. Dunn, “Frequency-domain wide-field laser doppler in vivo imaging,” Opt. Lett. 31, 2762–2764 (2006). [CrossRef]   [PubMed]  

48. M. Atlan, B. C. Forget, A. C. Boccara, T. Vitalis, A. Rancillac, A. K. Dunn, and M. Gross, “Cortical blood flow assessment with frequency-domain laser doppler microscopy,” J. Biomed. Opt. 12, 024019 (2007). [CrossRef]   [PubMed]  

49. M. Atlan, M. Gross, T. Vitalis, A. Rancillac, J. Rossier, and A. Boccara, “High-speed wave-mixing laser doppler imaging in vivo,” Opt. Lett. 33, 842–844 (2008). [CrossRef]   [PubMed]  

50. N. Warnasooriya, F. Joud, P. Bun, G. Tessier, M. Coppey-Moisan, P. Desbiolles, M. Atlan, M. Abboud, and M. Gross, “Imaging gold nanoparticles in living cell environments using heterodyne digital holographic microscopy.” Opt. Express 18, 3264–3273 (2010). [CrossRef]   [PubMed]  

51. F. Verpillat, F. Joud, P. Desbiolles, and M. Gross, “Dark-field digital holographic microscopy for 3d-tracking of gold nanoparticles,” Opt. Express 19, 26044–26055 (2011). [CrossRef]  

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

53. J. Gao, J. A. Lyon, D. P. Szeto, and J. Chen, “In vivo imaging and quantitative analysis of zebrafish embryos by digital holographic microscopy,” Biomed. Opt. Express 3, 2623–2635 (2012). [CrossRef]   [PubMed]  

Supplementary Material (3)

Media 1: AVI (1539 KB)     
Media 2: AVI (1367 KB)     
Media 3: AVI (7296 KB)     

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 (8)

Fig. 1
Fig. 1 Heterodyne digital holographic microscopy experimental arrangement (a) Injection part of the interferometer. HWP: half wave plate; PBS: polarizing beam splitter; AOM1, AOM2: acousto optic modulators (Bragg cells). (b) Classical upright microscope used for the off-axis recombinaison of reference and object beams. BS: cube beam splitter; CCD: CCD camera.
Fig. 2
Fig. 2 Detection of zebrafish embryo blood flow in transmission geometry. (a) Illustration of the spectral broadening of light encountering moving scattering objects (here, red blood cells). (b) Principles of the spectral measurement by heterodyne digital holography.
Fig. 3
Fig. 3 Spatial filtering and reconstruction principle. Holograms H(x, y) (a), H1(kx, ky) (b), H2(kx, ky) (c) and H3(x, y, z = 0) (d): see Media 1. The display is made in arbitrary Log scale for the average intensity 〈|HX|2〉. This average is calculated made by making the reconstruction with I0..I3, with I1..I4... and with I28..I31, and by averaging the resulting |HX|2. Images (a) and (d) show a side view of a 6 days old zebrafish embryo. Ventral side in on the right and anterior is up.
Fig. 4
Fig. 4 Two phase detection efficiency |η|2 as a function of the LO versus signal frequency offset: x. Calculation is made by Eq. 6 with TCCD = 100 ms and Texp = 50 ms.
Fig. 5
Fig. 5 Reconstructed hologram H3(x, y, z = 0) made for (ωLOωI)/(2π) = 0 Hz (a): see also Media 2, 10 Hz (b) ... to 80 Hz (i). Images (a) to (i) are displayed with the same Log scale for the average intensity 〈|H3|2〉. The sample is the same as the one in Fig. 3.
Fig. 6
Fig. 6 Dependance of the Doppler holographic signal 〈|H3(x, y)|2〉 with the frequency off-set (ωLOωI)/(2π) for different location A to F of the reconstructed image of Fig. 3(a). Curves are drawn with linear (a) and logarithmic (b) scales.
Fig. 7
Fig. 7 Fourier space reconstructed hologram H2(kx, ky) made without (a) and with (b,c) selection of the scattered wave vector kS. In (b) the selected zone is oriented toward uφ=0. In (c), three zones oriented along uφ with φ = 0 (blue), 2π/3 (green) and 4π/3 (red) are selected. Display is made in arbitrary log scale for 〈|H2|2〉. Frequency shift is (ωLOωI) = 0.
Fig. 8
Fig. 8 Colored reconstructed hologram H3(x, y, z = 0) made for (ωLOωI)/(2π) = 0 Hz (a), 10 Hz (b) ... to 60 Hz (g), and for −20 Hz (h) and + 20 Hz (i). Images (a) to (g) are made with same zebrafish sample and same viewpoint as Fig. 3 and Fig. 5. Images (h) to (i) correspond to another fish embryo. Images (a) to (g) (and (h) and (i)) are displayed with the same color RGB Log scale for the average intensity 〈|H3|2〉. Media 3 corresponds to images (h) and (i).

Equations (6)

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

ω S ω I = q.v
H ( x , y ) = [ I 0 ( x , y ) I 2 ( x , y ) ] + j [ I 1 ( x , y ) I 3 ( x , y ) ]
H 1 ( k x , k y ) = FFT [ H ( x , y ) e j | k | ( x 2 + y 2 ) / 2 d ]
H 3 ( x , y , z ) = FFT 1 [ H 2 ( k x , k y ) e j ( k x 2 + k y 2 ) / 2 z ]
Δ ω = ω L O ω I = m ω CCD H ( x , y ) = I 0 ( x , y ) I 1 ( x , y )
η ( x ) = 1 2 T exp k = 0 1 ( 1 ) k t = k T CCD T exp / 2 k T CCD + T exp / 2 e j 2 π x t d t = 1 2 sinc ( π x T exp ) [ 1 e j 2 π x T CCD ]
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.