Open Access
1 July 2006 Linear response range characterization and in vivo application of laser speckle imaging of blood flow dynamics
Bernard Choi, Julio C. Ramírez-San-Juan, Justin Lotfi, J. Stuart Nelson M.D.
Author Affiliations +
Abstract
Noninvasive blood flow imaging can provide critical information on the state of biological tissue and the efficacy of approaches to treat disease. With laser speckle imaging (LSI), relative changes in blood flow are typically reported, with the assumption that the measured values are on a linear scale. A linear relationship between the measured and actual flow rate values has been suggested. The actual flow rate range, over which this linear relationship is valid, is unknown. Herein we report the linear response range and velocity dynamic range (VDR) of our LSI instrument at two relevant camera integration times. For integration times of 1 and 10 ms, the best case VDR was 80 and 60 dB, respectively, and the worst case VDR was 20 and 50 dB. The best case VDR values were similar to those reported in the literature for optical Doppler tomography. We also demonstrate the potential of LSI for monitoring blood flow dynamics in the rodent dorsal skinfold chamber model. These findings imply that LSI can provide accurate wide-field maps of microvascular blood flow rate dynamics and highlight heterogeneities in flow response to the application of exogenous agents.

1.

Introduction

Noninvasive blood flow imaging can provide critical information on the state of biological tissue and the efficacy of approaches to treat disease. Fluorescein angiography is used routinely to study blood flow dynamics in the retina.1 Laser Doppler flowmetry and laser Doppler imaging have been applied in numerous preclinical and clinical studies on the brain,2, 3 retina,4, 5 skin,6, 7 and joints.8 Principles of Doppler shift imaging have been combined with optical coherence tomography to provide high-resolution cross-sectional images of blood flow. 9, 10, 11, 12, 13, 14, 15, 16, 17

In 1981, Fercher and Briers18 proposed a laser speckle imaging (LSI) approach as an alternative to laser Doppler imaging. On the basis of this study, it was concluded that variations in speckle contrast can be used to provide directly a wide-field velocity distribution map. LSI has since been employed for wide-field blood flow imaging of the brain, 2, 19, 20, 21, 22 retina,18, 23 and skin, 6, 24, 25, 26 with high temporal resolution. Briers27 has demonstrated that LSI and laser Doppler imaging provide equivalent assessments of blood flow.

With LSI, relative changes in blood flow are typically reported, with the assumption that the measured values are on a linear scale. A linear relationship between the measured and actual flow rate values has been suggested.2, 19 The actual flow rate range over which this linear relationship is valid is unknown. Knowledge of the relationship between measured and actual blood flow values is important because researchers would then know to which in vitro and in vivo flow models LSI could be applied to quantify relative flow changes.

The primary objective of this paper is to provide insight into the relationship between LSI measurements of speckle flow index with actual flow rate. The secondary objective is to demonstrate application of LSI to an in vivo rodent dorsal skinfold model and to show its potential as a wide-field microvascular imaging modality in a widely used animal model. Based on the analysis of Yuan, 21 the central hypothesis was that image integration time would affect the overall velocity dynamic range and signal-to-noise ratio of LSI.

2.

Materials and Methods

2.1.

Relationship Between Speckle Flow Index and Actual Flow Rate

The objective of this study was to establish the actual flow rate range over which a linear relationship existed with measured speckle flow index (SFI). To accomplish this objective, we developed a simple in vitro flow phantom to simulate blood flow in the rodent dorsal skinfold model. In the accompanying experiments, we varied flow rate and camera integration time.

2.1.1.

Experimental Design

LSI instrument

The instrument consists of a 30-mW, 633-nm helium-neon laser (Edmund Industrial Optics, Barrington, New Jersey), plano-convex lens, beam steering mirrors, digital charge-coupled device (CCD) camera (Retiga EXi, QImaging, Burnaby, British Columbia, Canada) equipped with a macro lens, and desktop Personal Computer (PC). Collimated light emitted from the laser became divergent after passing through the lens. To make the instrument more compact, mirrors were used to steer the diverging beam to the target plane. The resultant speckle pattern was imaged on the CCD array at 1:1 magnification and acquired to the PC memory. The f -stop of the macro lens was set to ensure a speckle size equivalent to the pixel dimensions (6.45×6.45μm2) of the camera.

In vitro flow phantom

To generate scattering agar gels, we heated a solution consisting of 100mL of deionized water and 10mL of glycerol to boiling. Glycerol was used to improve the mechanical integrity of the resultant gels. We added simultaneously 0.3g of TiO2 (Sigma, St. Louis, Missouri) and 2g of agar to the heated solution. The former was used to increase the scattering coefficient of the otherwise clear gels, and the agar added was deemed appropriate to simulate the reduced scattering coefficient of skin, based on previous frequency domain photon migration experiments.28 The solution was then poured into various molds (e.g., Petri dishes, sandwiched glass slides) to achieve the desired thickness ( 150μm to several millimeters) and allowed sufficient time to solidify. In specific gels, an 550-μm inner diameter glass capillary tube (Chase Instruments Corp., Glen Falls, New York) was embedded into the mold prior to solidification. A syringe-based infusion pump (Harvard Apparatus, Holliston, Massachusetts) was used to inject fluid into the flow tube. Tygon tubing delivered the fluid from the filled syringe mounted on the pump to the tube embedded in the gel. Five hundred mL of indated whole blood was obtained from the San Diego Blood Bank (San Diego, California) and used in this study.

Experiments

The infusion pump was set to achieve flow rates up to 21mms . The flow phantom tube was imaged with the LSI instrument. Camera integration times of 1 and 10ms were selected as representative values based on those used in previous studies. 2, 6, 19, 20, 21, 22, 25, 26

LSI image analysis

The image processing algorithm has been described previously in detail.2, 24 Briefly, the recorded image sequence was converted to speckle contrast images by applying a 7×7 sliding window to each 1392×1040 image. Although the 7×7 sliding window reduces the spatial resolution of our system, such a technique is necessary to preserve the assumption of first order speckle statistics.27 Because the resultant speckle contrast image is essentially oversampled, relatively high spatial frequency details remain preserved in the speckle contrast and subsequent flow images. At each window position, the mean graylevel intensity (I) and standard deviation (σ) were determined, and the speckle contrast (K) of the center pixel in the window can be computed as

Eq. 1

K=σI.
Assuming a Lorentzian flow velocity distribution,29 the correlation time (τc) of the intensity fluctuations can be calculated from:

Eq. 2

K={τc2T[1exp(2Tτc)]},
where T is the camera integration time. Maple software (Maplesoft, Ontario, Canada) was used to obtain an analytic series expansion for τc as a function of K , for the values of T used in this study. Relative flow images were obtained by calculating 1τc at each image pixel; a higher pixel value was assumed to be analogous to higher blood flow.

Because we were not concerned with real-time speckle flow imaging in this study, we performed the image processing offline. We selected a central 260×260μm2 (i.e., 40×40pixels ) region of interest and computed the mean SFI. We then determined the mean SFI from images acquired at an actual flow rate of zero, corresponding to a background contribution from Brownian motion of scatterers. For further analysis, we used the difference between mean SFI values at nonzero actual flow rates and the background values.

2.2.

In Vivo Application of LSI

The objective of this study was to demonstrate in vivo application of LSI. To accomplish this objective, we applied chemical agents to the microvascular network of the rodent dorsal skinfold chamber model. 25, 30, 31, 32, 33, 34, 35, 36, 37

2.2.1.

Experimental Design

Rodent dorsal skinfold chamber model. Surgical installation of a dorsal skinfold window permitted observation of full-thickness skin from both the epidermal and subdermal sides (Fig. 1 ). The animal was anesthetized with a combination of Ketamine and Xylazine (4:3 ratio, 0.1g100g body weight) administered by intraperitoneal injection. A section for window placement on the back of the animal was selected, surgically scrubbed, shaved, and depilated. Sutures attached to a temporary mount retracted the dorsal skin away from the animal’s body. A circular section with an approximate diameter of 1cm was cut from one side of the symmetrical skinfold, thus exposing blood vessels in the underlying skin. An aluminum chamber was sutured to both sides of the skin and the sutures cut to release the skin from the mount. To prevent dehydration, saline was applied periodically to the subdermal skin. The surgery was performed as defined in a protocol approved by the University of California, Irvine, Animal Use Committee.

Fig. 1

Representative subdermal image of a dorsal skinfold window chamber preparation.

041129_1_041604jbo1.jpg

Chemical agents

We were interested in studying microvascular flow during the application of a recently developed polyurethane agent designed to reduce temporarily optical scattering in biological tissues such as skin.38 The agent was applied directly to the subdermal microvascular network and LSI images acquired over a 15-min period from the subdermal side. Due to lateral diffusion of the agent away from the imaged microvascular network after application, the agent was reapplied on a regular basis during the image acquistion period. In separate experiments, each on a different animal, phosphate buffered saline and 13.6M glycerol were applied as negative and positive controls, respectively, for inducing microvascular flow changes.

Image acquisition and processing

Raw speckle images were acquired and SFI images computed as described above. White light reflectance images were also acquired to serve as anatomical reference images. To visualize changes in blood flow induced by agent application, the SFI images were converted to binary images using the iterative selection method.39 This automated method was used to determine a threshold SFI value based on an initial estimate SFI0 . The mean SFI values of all pixels below (SFIbelow) and above (SFIabove) SFI0 were calculated and a new estimate SFInew computed as the arithmetic mean of SFIbelow and SFIabove . This process was repeated with each new estimate until SFIbelow was equal to SFIabove .

With each experiment, a threshold SFI value was determined for the SFI image acquired immediately after agent application and was applied to binarize SFI images acquired at later time points. To visualize better arteriolar and venular blood flow dynamics, the binary SFI images were superimposed on the corresponding white light anatomical images.

3.

Results

Camera integration time impacts the relationship between measured SFIs and actual flow rate (Fig. 2 )

With an increase in integration time from 1 to 10ms , the following trends were observed with the in vitro flow phantom: (1) the linear response range decreased from 2 to 20mms , to 0 to 5mms ; (2) the best case velocity resolution, defined as the reciprocal of the slope of a linear fit to the data, decreased from 2 to 5μms ; and (3) as shown by the correlation coefficient of the linear fits, the amount of scatter in the data decreased. Collectively, these data provide us with knowledge on the actual flow rate range over which our LSI instrument has a linear response.

Fig. 2

Effect of camera integration time on the linear response range between SFI and actual flow rate. Data acquired with integration times of (a) 1 and (b) 10ms are presented. The data show that an increase in integration time decreases the linear response velocity range and velocity resolution and increases the linear regression fit to the discrete data points. In (a), the full set of acquired data is shown, and the linear fit is restricted to actual flow rate values of 2 to 20mms .

041129_1_041604jbo2.jpg

Wide-field, arteriolar, and venular blood flow dynamics are readily monitored during exogenous chemical agent application

Representative white light, SFI, and superimposed white light–binary SFI images of the rodent dorsal skinfold window chamber are shown in Fig. 3 . A good correspondence between visible blood vessels is evident from a comparison of the white light and SFI images and in the superimposed image. By applying the same threshold SFI value to all images acquired in a time-resolved sequence, blood flow dynamics are readily visualized in a qualitative fashion. With direct application of saline (i.e., negative control) to the microvascular network for 16min , a slight increase in blood flow was observed, primarily in the smaller arteriolar and venular branches (Fig. 4 ). With direct application of glycerol (i.e., positive control) to the microvascular network, a gradual decrease in blood flow was observed (Fig. 5 ). From the superimposed white light–binary SFI image, we observed an immediate decrease in venular flow and a delayed decrease in arteriolar flow (between 12 and 16min after glycerol application). With direct application of the polyurethane agent to the microvascular network, a slight increase in blood flow was observed in the smaller vessel branches (Fig. 6 ), similar to that observed with saline application. Collectively, these data demonstrate the potential of LSI to allow wide-field monitoring of heterogeneous blood flow dynamics.

Fig. 3

Representative (a) white light reflectance, (b) SFI, and (c) superimposed white light–binary SFI images of the microvascular network in the rodent dorsal skinfold chamber. Image dimensions are 9×6.8mm2 .

041129_1_041604jbo3.jpg

Fig. 4

Representative (a) white light reflectance and (b, c) superimposed white light–binary SFI images at 0 and 16min , respectively, after direct saline application to the microvascular network of the rodent dorsal skinfold chamber. The images show a slight increase in blood flow in the smaller blood vessel branches [(arrows in (c)]. Image dimensions are 9×6.8mm2 .

041129_1_041604jbo4.jpg

Fig. 5

Representative (a) white light reflectance and (b to f) superimposed white light–binary SFI images at various time points after direct glycerol application to the microvascular network of the rodent dorsal skinfold chamber. The images show a heterogeneous decrease in blood flow, with an initial decrease in venular flow [“V” in (a)] and a delayed decrease in arteriolar flow [“A” in (a)]. Image dimensions are 9×6.8mm2 .

041129_1_041604jbo5.jpg

Fig. 6

Representative (a) white light reflectance and (b to f) superimposed white light–binary SFI images at various time points after direct polyurethane application to the microvascular network of the rodent dorsal skinfold chamber. The images show a slight increase in blood flow in the smaller blood vessel branches. Image dimensions are 9×6.8mm2 .

041129_1_041604jbo6.jpg

4.

Discussion

We are primarily interested in studying hemodynamics in in vivo microvascular network animal models, such as the dorsal skinfold window chamber preparation. Before we could use LSI reliably as a quantitative tool to assess blood flow in these models, we first needed to study the relationship between SFI and actual flow rate. In this study, we identified a range over which the relationship is linear, for two commonly used camera integration times. Specifically, we have shown that, for integration times of 1 and 10ms , the linear relationship holds for flow rates of 2 to 20 and 0 to 5mms , respectively (Fig. 2). There is slightly more scatter in the 1-ms integration time data than that at 10ms and hence more deviation from the linear regression fit. A similar trend in improved speckle contrast-to-noise ratio was observed by Yuan 21

Our data suggest that camera integration time should be selected with the anticipated velocity range to be evaluated. Vargas 15 used optical Doppler tomography to measure blood flow velocities of 4mms in arterioles and venules in the rodent dorsal skinfold model. From blood flow rates measured in arterioles and venules of the rodent cranial window model, velocities of less than 5mms were anticipated.32 To ensure that a linear relationship is maintained between measured speckle flow index values and actual flow rate, our data suggest an integration time of approximately 10ms should be used to monitor flow rate dynamics directly in arterioles and venules. From the analysis of Yuan, 21 the speckle contrast to noise ratio is only 8% less at an integration time of 10ms than the maximum ratio at 5ms , which we believe is an acceptable compromise to ensure measurements are acquired over a linear response range.

The data obtained in this study can be used to estimate a velocity dynamic range (VDR) for our LSI instrument10

Eq. 3

VDR=20log(VzΔVz),
where VDR is in units of decibels, Vz is the velocity range, and ΔVz is the velocity resolution. It is instructive to study two VDR limits, focusing on the linear response velocity range only: best case [i.e., ΔVz equals the inverse of the slope of the regression line between SFI and actual flow rate (Fig. 2)]; and worst case (i.e., ΔVz equals the quotient of the largest deviation from the regression line and the slope of the line). For integration times of 1 and 10ms , the best case VDR is 80 and 60dB , respectively, and the worst case VDR is 20 and 50dB . The best case VDR values are similar to those reported by Yang 10 for their optical Doppler tomography system. Thus, for blood flow characterization in superficial vessels such as those found in the rodent dorsal skinfold chamber, chick chorioallantoic membrane, and rodent cranial window, the performance of LSI is on par with optical coherence tomography.

We performed a series of experiments (Figs. 4 to 6) to demonstrate the potential of LSI for wide-field monitoring of blood flow dynamics in an in vivo microvascular network model. We observed a heterogeneous decrease in blood flow after glycerol application (Fig. 5), with an initial decrease in venules and a delayed decrease in arterioles. A similar trend was observed by Vargas 15 with optical Doppler tomography monitoring of a single arteriole-venule pair. A potential advantage of LSI over such a technique is the ability to interrogate multiple vessels and associated branches simultaneously.

A surprising finding was that the blood flow rate increased slightly with application of phosphate buffered saline (Fig. 4) and the polyurethane agent (Fig. 6). The thermoregulatory aspect of cutaneous blood flow may have led to this increase, because the agent temperature (25°C) was slightly lower than the in vivo skinfold temperature (28°C) . Another possibility is that the reduced access of the subdermal tissue to ambient oxygen due to the addition of saline may have led to a compensatory increase in blood flow to reestablish steady-state oxygenation conditions. The strong flow reduction effect of glycerol was probably sufficient to overcome any compensatory increase in blood flow due to either effect.

Our current LSI instrument is capable of raw speckle image acquisition rates of 100 frames per s, which corresponds to a temporal resolution of 10ms . The intent of the current in vivo imaging study was to compare our data (Figs. 4 to 6) with previously published data,15 so we acquired image sequences during agent application at a considerably slower frame rate of 0.5 frames per min. Previously, we have acquired raw speckle images within seconds after pulsed dye laser irradiation of select blood vessels,25 demonstrating the potential of LSI to study rapid transients in blood flow.

With our optical imaging configuration, we acquired images at unity magnification to provide a raw speckle image lateral resolution of 6.45μm . To achieve this spatial resolution, we selected the f-number of the macro lens so that the speckle size was equivalent to that of a single pixel2

Eq. 4

w=1.2(M+1)λ(f#),
where w is the speckle size, M is the optical magnification, and f# is the f -number. Although the intent of the current study was to perform wide-field blood flow imaging of the dorsal skinfold window chamber preparation, it is possible to increase the optical magnification by selecting the appropriate zoom lens with a decreased f# setting. The current hypothesis is that the speckle size should be equal to the size of a single pixel to achieve appropriate sampling of the speckle pattern.24 Equation 4 sets an upper limit on the image magnification due primarily to the available f# settings on commercially available lenses. To the best of our knowledge, a quantitative theoretical description to support the current hypothesis has not been proposed, so it may be possible to perform high-magnification LSI (i.e., “laser speckle microscopy”) with an optical configuration that violates Eq. 4 We have planed a set of experiments to test the current hypothesis and will report our findings in a future publication.

Theoretical27 and experimental2 studies have established that LSI and laser Doppler imaging provide equivalent relative flow values. Thus, if the same laser wavelength is used, both methods interrogate perfusion within equivalent volumes of tissue. The primary advantages of LSI include its relative simplicity (i.e., no need for scanning mechanical components) and its image acquisition rate. LSI images can be acquired at the maximum camera frame rate (typically 30+ frames per s for 300+ kilopixel images) and processed offline. With laser Doppler imaging, acquisition times of several minutes are required to achieve a 65-kilopixel image.6 Such prolonged acquisition times are associated with a considerably higher risk of motion artifact. Laser Doppler imaging does provide information on the direction of blood flow, a capability that is not present with conventional LSI. These considerations should be taken into account during selection of the appropriate wide-field blood flow imaging method.

The LSI velocity dynamic range and signal-to-noise ratio reported here provide an imaging platform that can be used to study normal and tumor microvasculature in the rodent dorsal skinfold model. Angiogenesis associated with tumor growth has been studied extensively with this animal model.31, 32, 37 In such studies, blood flow dynamics are typically monitored with point laser Doppler probes or direct red blood cell counting algorithms. Coupling of LSI with other optical imaging modalities, such as reflectance,22 fluorescence, and multiphoton microscopy, may enhance knowledge of critical factors such as heterogeneities in tumor blood flow and local oxygen consumption dynamics.

In summary, we have shown that SFI and actual flow rate have a linear relationship at typically used LSI integration times and at blood flow rate values typical of arterioles, venules, and capillaries. With an increase in integration time, the VDR decreases and signal-to-noise ratio increases. These findings imply that LSI can provide accurate wide-field maps of microvascular blood flow rate dynamics and highlight heterogeneities in flow response to the application of exogenous agents.

Acknowledgments

The authors acknowledge financial support from the American Society for Laser Medicine and Surgery, Inc. (B.C.); the Arnold and Mabel Beckman Fellows Program (B.C.); the Beckman Foundation; the Fondo de Repatriaciones CONACYT-México (J.C.R.S.J.); and National Institutes of Health [Grants No. AR047551, EB02495, and AR048458, (J.S.N.); and the Laser Microbeam and Medical Program (LAMMP), a NIH Biomedical Technology Resource, Grant No. P41-RR01192, at the University of California, Irvine]. The authors thank Dr. Roger Chiu-Zarate (Beckman Laser Institute and Medical Clinic) for assistance with the LSI instrument software.

References

1. 

Y. Tao, Y. R. Jiang, X. X. Li, C. Y. Yin, and M. Yao, “Fundus and histopathological study of radial optic neurotomy in the normal miniature pig eye,” Arch. Ophthalmol. (Chicago), 123 1097 –1101 (2005). 0003-9950 Google Scholar

2. 

A. K. Dunn, T. Bolay, M. A. Moskowitz, and D. A. Boas, “Dynamic imaging of cerebral blood flow using laser speckle,” J. Cereb. Blood Flow Metab., 21 195 –201 (2001). 0271-678X Google Scholar

3. 

S. A. Sheth, M. Nemoto, M. W. Guiou, M. A. Walker, and A. W. Toga, “Spatiotemporal evolution of functional hemodynamic changes and their relationship to neuronal activity,” J. Cereb. Blood Flow Metab., 25 830 –841 (2005). 0271-678X Google Scholar

4. 

R. D. Ferguson, D. X. Hammer, A. E. Elsner, R. H. Webb, and S. A. Burns, “Wide-field retinal hemodynamic imaging with the tracking scanning laser ophthalmoscope,” Opt. Express, 12 5198 –5208 (2004). https://doi.org/10.1364/OPEX.12.005198 1094-4087 Google Scholar

5. 

C. E. Riva, E. Logean, and B. Falsini, “Visually evoked hemodynamical response and assessment of neurovascular coupling in the optic nerve and retina,” Prog. Retin Eye Res., 24 183 –215 (2005). 1350-9462 Google Scholar

6. 

K. R. Forrester, J. Tulip, C. Leonard, C. Stewart, and R. C. Bray, “A laser speckle imaging technique for measuring tissue perfusion,” IEEE Trans. Biomed. Eng., 51 2074 –2084 (2004). 0018-9294 Google Scholar

7. 

A. Boignard, M. Salvat-Melis, P. H. Carpentier, C. T. Minson, L. Grange, C. Duc, F. Sarrot-Reynauld, and J. L. Cracowski, “Local hyperhemia to heating is impaired in secondary Raynaud’s phenomenon,” Arthritis Res. Ther., 7 R1103 –R1112 (2005). 1478-6354 Google Scholar

8. 

C. G. Egan, J. C. Lockhart, and W. R. Ferrell, “Pathophysiology of vascular dysfunction in a rat model of chronic joint inflammation,” J. Physiol. (London), 557 635 –643 (2004). 0022-3751 Google Scholar

9. 

H. W. Ren, Z. H. Ding, Y. H. Zhao, J. J. Miao, J. S. Nelson, and Z. P. Chen, “Phase-resolved functional optical coherence tomography: simultaneous imaging of in situ tissue structure, blood flow velocity, standard deviation, birefringence, and Stokes vectors in human skin,” Opt. Lett., 27 1702 –1704 (2002). 0146-9592 Google Scholar

10. 

V. X. D. Yang, M. L. Gordon, B. Qi, J. Pekar, S. Lo, E. Seng-Yue, A. Mok, B. C. Wilson, and I. A. Vitkin, “High speed, wide velocity dynamic range Doppler optical coherence tomography (Part I): System design, signal processing, and performance,” Opt. Express, 11 794 –809 (2003). 1094-4087 Google Scholar

11. 

B. J. Vakoc, S. H. Yun, J. F. de Boer, G. J. Tearney, and B. E. Bouma, “Phase-resolved optical frequency domain imaging,” Opt. Express, 13 5483 –5493 (2005). https://doi.org/10.1364/OPEX.13.005483 1094-4087 Google Scholar

12. 

J. K. Barton and S. Stromski, “Flow measurement without phase information in optical coherence tomography images,” Opt. Express, 13 5234 –5239 (2005). https://doi.org/10.1364/OPEX.13.005234 1094-4087 Google Scholar

13. 

V. X. D. Yang, Y. X. Mao, N. Munce, B. Standish, W. Kucharczyk, N. E. Marcon, B. C. Wilson, and I. A. Vitkin, “Interstitial Doppler optical coherence tomography,” Opt. Lett., 30 1791 –1793 (2005). 0146-9592 Google Scholar

14. 

C. J. Pedersen, S. Yazdanfar, V. Westphal, and A. M. Rollins, “Phase-referenced Doppler optical coherence tomography in scattering media,” Opt. Lett., 30 2125 –2127 (2005). 0146-9592 Google Scholar

15. 

G. Vargas, A. Readinger, S. S. Dozier, and A. J. Welch, “Morphological changes in blood vessels produced by hyperosmotic agents and measured by optical coherence tomography,” Photochem. Photobiol., 77 541 –549 (2003). https://doi.org/10.1562/0031-8655(2003)077<0541:MCIBVP>2.0.CO;2 0031-8655 Google Scholar

16. 

X. J. Wang, T. E. Milner, and J. S. Nelson, “Characterization of fluid flow velocity by optical Doppler tomography,” Opt. Lett., 20 1337 –1339 (1995). 0146-9592 Google Scholar

17. 

Z. P. Chen, T. E. Milner, D. Dave, and J. S. Nelson, “Optical Doppler tomographic imaging of fluid flow velocity in highly scattering media,” Opt. Lett., 22 64 –66 (1997). 0146-9592 Google Scholar

18. 

A. F. Fercher and J. D. Briers, “Flow Visualization by Means of Single-Exposure Speckle Photography,” Opt. Commun., 37 326 –330 (1981). https://doi.org/10.1016/0030-4018(81)90428-4 0030-4018 Google Scholar

19. 

H. Y. Cheng, Q. M. Luo, S. Q. Zeng, S. B. Chen, J. Cen, and H. Gong, “Modified laser speckle imaging method with improved spatial resolution,” J. Biomed. Opt., 8 559 –564 (2003). https://doi.org/10.1117/1.1578089 1083-3668 Google Scholar

20. 

H. Y. Cheng, Q. M. Luo, S. Q. Zeng, S. B. Chen, W. H. Luo, and H. Gong, “Hyperosmotic chemical agent’s effect on in vivo cerebral blood flow revealed by laser speckle,” Appl. Opt., 43 5772 –5777 (2004). https://doi.org/10.1364/AO.43.005772 0003-6935 Google Scholar

21. 

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). https://doi.org/10.1364/AO.44.001823 0003-6935 Google Scholar

22. 

A. K. Dunn, A. Devor, A. M. Dale, and D. A. Boas, “Spatial extent of oxygen metabolism and hemodynamic changes during functional activation of the rat somatosensory cortex,” Neuroimage, 27 279 –290 (2005). 1053-8119 Google Scholar

23. 

M. Hirao, H. Oku, W. Goto, T. Sugiyama, T. Kobayashi, and T. Ikeda, “Effects of adenosine on optic nerve head circulation in rabbits,” Exp. Eye Res., 79 729 –735 (2004). 0014-4835 Google Scholar

24. 

J. D. Briers, G. Richards, and X. W. He, “Capillary blood flow monitoring using laser speckle contrast analysis (LASCA),” J. Biomed. Opt., 4 164 –175 (1999). https://doi.org/10.1117/1.429903 1083-3668 Google Scholar

25. 

B. Choi, N. M. Kang, and J. S. Nelson, “Laser speckle imaging for monitoring blood flow dynamics in the in vivo rodent dorsal skinfold model,” Microvasc. Res., 68 143 –146 (2004). 0026-2862 Google Scholar

26. 

T. K. Smith, B. Choi, J. C. Ramirez-San-Juan, J. S. Nelson, K. Osann, and K. M. Kelly, “Microvascular blood flow dynamics associated with photodynamic therapy and pulsed dye laser irradiation,” Lasers Surg. Med., 38 532 –539 (2006). 0196-8092 Google Scholar

27. 

J. D. Briers, “Laser Doppler and time-varying speckle: A reconciliation,” J. Opt. Soc. Am. A Opt. Image Sci. Vis, 13 345 –350 (1996). 1084-7529 Google Scholar

28. 

S. MerriH, Google Scholar

29. 

J. W. Goodman, Statistical Optics, John Wiley & Sons, Inc., New York (1985). Google Scholar

30. 

J. K. Barton, A. J. Welch, and J. A. Izatt, “Investigating pulsed dye laser-blood vessel interaction with color Doppler optical coherence tomography,” Opt. Express, 3 251 –256 (1998). 1094-4087 Google Scholar

31. 

M. W. Dewhirst, S. Shan, Y. T. Cao, B. Moeller, F. Yuan, and C. Y. Li, “Intravital fluorescence facilitates measurement of multiple physiologic functions and gene expression in tumors of live animals,” Dis. Markers, 18 293 –311 (2002). 0278-0240 Google Scholar

32. 

R. K. Jain, L. L. Munn, and D. Fukumura, “Dissecting tumour pathophysiology using intravital microscopy,” Nat. Rev. Cancer, 2 266 –276 (2002). 1474-175X Google Scholar

33. 

Q. Chen, S. Tong, M. W. Dewhirst, and F. Yuan, “Targeting tumor microvessels using doxorubicin encapsulated in a novel thermosensitive liposome,” Mol. Cancer Ther., 3 1311 –1317 (2004). Google Scholar

34. 

P. Cabrales, A. G. Tsai, and M. Intaglietta, “Alginate plasma expander maintains perfusion and plasma viscosity during extreme hemodilution,” Am. J. Physiol. Heart Circ. Physiol., 288 H1708 –H1716 (2005). 0363-6135 Google Scholar

35. 

N. Isaka, T. P. Padera, J. Hagendoorn, D. Fukumura, and R. K. Jain, “Peritumor lymphatics induced by vascular endothelial growth factor-C exhibit abnormal function,” Cancer Res., 64 4400 –4404 (2004). 0008-5472 Google Scholar

36. 

P. Babilas, G. Shafirstein, W. Baumler, J. Baier, M. Landthaler, R. M. Szeimies, and C. Abels, “Selective photothermolysis of blood vessels following flashlamp-pumped pulsed dye laser irradiation: In vivo results and mathematical modelling are in agreement,” J. Invest. Dermatol., 125 343 –352 (2005). 0022-202X Google Scholar

37. 

B. S. Sorg, B. J. Moeller, O. Donovan, Y. Cao, and M. W. Dewhirst, “Hyperspectral imaging of hemoglobin saturation in tumor microvasculature and tumor hypoxia development,” J. Biomed. Opt., 10 044004 (2005). https://doi.org/10.1117/1.2003369 1083-3668 Google Scholar

38. 

M. Khan, B. Choi, S. Chess, K. Kelly, J. McCullough, and J. S. Nelson, “Optical clearing of in vivo human skin: Implications for light-based diagnostic imaging and therapeutics,” Lasers Surg. Med., 34 83 –85 (2004). https://doi.org/10.1002/lsm.20014 0196-8092 Google Scholar

39. 

T. W. Ridler and S. Calvard, “Picture thresholding using an iterative selection method,” IEEE Trans. Syst. Man Cybern., 8 630 –632 (1978). 0018-9472 Google Scholar
©(2006) Society of Photo-Optical Instrumentation Engineers (SPIE)
Bernard Choi, Julio C. Ramírez-San-Juan, Justin Lotfi, and J. Stuart Nelson M.D. "Linear response range characterization and in vivo application of laser speckle imaging of blood flow dynamics," Journal of Biomedical Optics 11(4), 041129 (1 July 2006). https://doi.org/10.1117/1.2341196
Published: 1 July 2006
Lens.org Logo
CITATIONS
Cited by 96 scholarly publications and 6 patents.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Blood circulation

Speckle

In vivo imaging

Doppler tomography

Cameras

Skin

Animal model studies

Back to Top