Open Access
1 May 2007 Illumination and fluorescence collection volumes for fiber optic probes in tissue
Dean Tai, Darren Hooks, John D. Harvey, Bruce H. Smaill, Christian Soeller
Author Affiliations +
Abstract
Optical fibers can deliver light to, and collect it from, regions deep in tissue. However, reported illumination and fluorescence collection volumes adjacent to the fiber tip have been inconsistent, and systematic data on this topic are not available. Illumination and fluorescence collection profiles were characterized with high spatial resolution for different optical fibers in tissue and various fluids using two-photon flash photolysis and excitation. We confirm that illumination and fluorescence collection volumes for optical fibers are near identical. Collection volume is determined by the core dimensions and numerical aperture (NA) of the fiber and the scattering properties of the medium. For a multimode optical fiber with 100 μm core diam and NA=0.22, 80% of the total fluorescence is collected from a depth of 170 μm in tissue and 465 μm in nonscattering fluid. A semiempirical mathematical description of photon flux adjacent to the fiber tip was also developed and validated. This was used to quantify the extent of temporal blurring associated with propagation of a wavefront of altered fluorescence emission across the region addressed by fiber optic probes. We provide information that will facilitate the design of optical probes for tissue imaging or therapeutic applications.

1.

Introduction

Over the past two decades, optical fibers have been employed in biophotonics applications in a variety of novel ways. Fiber-based measurement of fluorescence from functional dyes has yielded important information about biological processes and pathology. For instance, optical fibers have been employed in the detection of subsurface tumors in rat models.1 Recently, a new field of light-activated cancer therapy has emerged,2 in which optical fibers are used to deliver light for treatment of specific tissue regions.3 Fiber optic components have been used to record electrical activity in cardiac tissue employing potential-sensitive dyes.4, 5 Moreover, simultaneous recordings of blood volume, oxygenation, and intracellular calcium levels have been made via optical fibers on the surface of the brain with a single indicator employing reflectance and fluorescence.6 A particular advantage of fiber optic–based systems is that they provide a minimally invasive means of applying these techniques deep in tissue. Fiber optic probes have been used to record electrical activity at single7, 8 and multiple intramural sites in the intact heart. 9, 10, 11, 12

In order to interpret the data collected with optical fiber-based systems or to design appropriate fiber optic probes, information about fiber illumination and fluorescence collection volumes is required. For instance, action potential upstrokes obtained in cardiac tissue using optical mapping techniques are generally much slower than those in corresponding intracellular recordings.12, 13 This is not due to the response of the dyes employed. Optical potentials recorded from very small collection areas in isolated or confluent cell cultures4, 14 have upstroke velocities very similar to those measured intracellularly. It has been argued that the blurring of the upstroke is due to the spread of electrical activation through the region addressed by the imaging system.13, 15, 16 On this basis, the capacity of a fluorescence imaging system to follow the activation upstroke should depend on propagation velocity and effective fluorescence collection volume.13, 15, 16

To date, no systematic study of the collection volumes of a range of fiber probes has been performed, and no one method for the determination of these volumes is widely accepted. Zhu and Yappert have presented a simple model for estimating the effective collection volumes of optical fibers17 in nonabsorbing, nonscattering media. However, an experimental approach to this problem is more appropriate, particularly for biological tissues that exhibit complex absorption and scattering properties.18, 19 Byars characterized the fluorescence collection volume of a 240μm diam, multimode (MM) optical fiber in heart tissue,10 but the approach employed had limited spatial resolution, and it is difficult to extrapolate from their findings to optical probes with different geometry.

The objective of the work described in this study was to characterize both illumination and fluorescence collection volumes in optical fibers of different diameters in nonscattering and scattering fluids and in tissue. Novel techniques utilizing two-photon (2P) microscopy and 2P flash photolysis have been developed for this purpose. In addition, a simple semiempirical model of fluorescence collection for an optical fiber has been developed as a means of interpreting experimental data and to aid optical probe design.

2.

Methods

2.1.

Definition of Effective Fluorescence Collection Volume

The effective fluorescence detection volume for an optical fiber that is used to supply excitation light to, and collect fluorescent emission from, a region adjacent to its tip is determined by both its illumination and collection volumes. The illumination intensity distribution adjacent to the tip of a flat-cleaved fiber is assumed to be axisymmetric with respect to a cylindrical polar coordinate system (r,z) , in which z is coincident with the axis of the fiber and the radial coordinate r is transverse to it. Following the analysis proposed by Zhu and Yappert,17 the illumination efficiency is therefore represented as Ie(r,z) , where 0<Ie<1 , and the fluorescence collection efficiency is similarly represented as Ic(r,z) , where 0<Ic<1 . Provided that illumination intensities are sufficiently low to avoid saturation, the fluorescence signal (dF) collected from a small volume (dV), at location (r,z) , can be expressed as:

Eq. 1

dF(r,z)=ηKC(r,z)Ic(r,z)Ie(r,z)dV,
where dV, the annular volume, is given by:

Eq. 2

dV=2πrdrdz,
and η and K are parameters proportional to the fluorescence quantum yield and absorptivity of the fluorophores, respectively, while C(r,z) describes the concentration of fluorophore. F(z) , the total fluorescence signal collected from within a depth z , can be determined by integrating dF across that volume. In the following estimate, it is assumed that fluorophore distribution is constant:

Eq. 3

F(z)=2πηKC0z0Ic(r,z)Ie(r,z)rdrdz.

Hence, the signal depends on the product of illumination and collection fields. The product

Eq. 4

Ieff(r,z)=Ie(r,z)×Ic(r,z),
describes the spatial properties of the fiber as a fluorescence probe and Ieff(r,z) is defined as the effective collection field.

The relative fluorescence signal

Eq. 5

R(z)=F(z)F(),
specifies the fraction of total fluorescence collected within a distance z from the fiber tip.

2.2.

Measurement of Fluorescence Collection Fields

Characterization of fiber response fields was performed by placing a fiber tip into the sample chamber (containing tissue or test solution) of a combined 2P/confocal laser scanning microscope (Zeiss LSM 410, Jena, Germany).20 A mode-locked Titanium:Sapphire (Ti:S) laser (Mira 700, Coherent, Santa Clara, California) tuned to 750nm provided excitation for 2P excitation of fluorophores or three-dimensionally resolved photorelease of caged fluoroscein. In the former cases, fluorescence emission was collected by the optical fiber under investigation, and measured with a photomultiplier tube (PMT) after passing through a bandpass emission filter (centered at 535nm and 50-nm wide, Chroma, Brattleboro, Vermont), see Fig. 1 . The microscope was used in 2P mode to create points of fluorescence emission by focusing the light of the Ti:S laser through a 10× objective into the sample. Since the fluorochrome concentration within the sample was uniform and the excitation profile resulting from 2P excitation was small compared to the length scale over which the collection field changed (i.e., it approximated a delta function), the fluorescence incident on the PMT was proportional to the value of Ic(r,z) when the 2P excitation spot was placed at the coordinates (r,z) [see Eq. 3]. In other words, the collection field was sampled by creating point sources of fluorescence of constant intensity at specified locations. To reconstruct the axisymmetric collection field of a flat-cleaved fiber, it was sufficient to scan only in the plane containing the fiber axis. On the other hand, to characterize the asymmetric collection field of an angle-cleaved optrode,9 it was necessary to collect a stack of such images.

Fig. 1

Schematic diagram of the imaging system. The system consists of a 2P confocal microscope, a PMT with a selectable emission filter F and an external laser source. For two-photon excitation studies, the 2P laser in the confocal microscope was used to stimulate fluorophores in the focal volume adjacent to an optical fiber (see inset A), and fluorescent emission collected by the fiber was detected by the PMT. Laser scanning was used to reconstruct fluorescence collection fields. For flash photolysis measurements, the 2P laser in the confocal microscope was used to release caged fluorophores in the same focal volume. For studies of fiber illumination profiles, continuous illumination was routed through the fiber from the external laser source, and the time-course of fluorescent emission by the uncaged fluorophore was monitored by the confocal microscope. A similar flash photolysis approach was used to provide additional characterization of fiber collection profiles (see Sec. 2 for details).

034033_1_035703jbo1.jpg

The scan time for a single 512×512 (sampling resolution 2.5μm ) 2P image was typically 16s , and the image signal-to-noise ratio (SNR) was improved by averaging over 16 images. For quantitative analysis, residual background was subtracted using control images recorded in the absence of 2P illumination.

2.2.1.

Measurement of collection fields with minimal fluorescence reabsorption

Although the preceeding technique yielded data with good spatial resolution, it is possible that fluorescent emission from fluorophores experiences reabsorption by surrounding fluorophore. Therefore, a variation of the technique described earlier was used to determine IC while minimizing attenuation of the fluorescence signal due to reabsorption by dye in the sample solution. A point source of fluorophore was provided by 2P excited flash photolysis of caged fluoroscein.21 Short bursts (200-ms duration) of light from the Ti:S laser were focused into a sample containing 1.67mM of caged fluoroscein using a 40×1.2 numerical aperture (NA) water immersion objective. The full-width half-maximum (FWHM) of the point spread function (PSF) of photorelease was 0.5μm in the xy plane and 1.4μm in the z -axis direction (coinciding with the axis of the objective).20 The released fluorophore was excited through the same objective using 488-nm light from the argon ion laser in the confocal microscope, which was aligned confocal with the stationary photolysis spot. A more detailed account of these methods is given elsewhere.20 During and after photolysis, the fluorescence received by the fiber was recorded while a line through the photolysis spot was repeatedly scanned with 488-nm light. For quantitative analysis, the line scan image was averaged temporally ( 50ms , after termination of the 2P release flash) and spatially (over a 10μm -wide region—see also Fig. 2 ) to obtain the collection efficiency at the photolysis spot. Due to diffusion, the full width of the fluorescence distribution at half maximal intensity was 10μm at the time of signal analysis (Fig. 2). While it was impractical to sample collection fields fully using 2P flash photolysis, this technique provides a means of characterizing their principal spatial features (see Sec. 3). The collection field adjacent to the fiber tip was characterized by performing 2P flash photolysis measurements at points 10 to 50μm apart in the axial direction and at selected transverse distances from the fiber axis. Since the fluorochrome concentration was insignificant at sites other than the photolysis spot, fluorescence reabsorption within the sample was negligible.

Fig. 2

Schematic diagram illustrating the flash photolysis technique. The 2P laser was used to release caged fluorophores in the focal volume adjacent to the fiber, which were then stimulated (see Fig. 1). Continuous line scans were acquired transverse to the fiber axis at this location. The 2P flash duration was 200ms , and the time-course of fluorescent emission for a typical such recording is shown at the center of the figure. Fluorescent emission at the focal point is shown in (a) as function of time, while (b) is a line scan acquired at the termination of the 2P flash. The relative fluorescence intensity at the point addressed by the 2P flash is estimated by integrating across the 10μm×50ms space-time window indicated by the white rectangle (see Sec. 2 for further details). Note that the traces shown in (a) and (b) have been averaged to improve signal quality.

034033_1_035703jbo2.jpg

2.3.

Determination of Illumination Fields

Fiber illumination fields were also determined by utilizing 2P excited flash photolysis of caged fluoroscein. The illumination field for excitation of the released fluorophore was provided for the fiber under investigation by coupling the light from the 488-nm line of an argon ion laser (N57-796, Edmund Optics Singapore PTE Ltd.) into the fiber entrance (see Fig. 1). Fluorescence emission from fluoroscein released at the stationary 2P photolysis spot was measured through the microscope objective using one of the PMTs in the confocal microscope. The intensity of the fluorescence signal measured directly after release of the caged fluorophore (see Fig. 2) was thus proportional to the illumination efficiency of the fiber at the location of the photolysis spot. This follows from Eq. 1, since C(r,z) was considered to be a delta function and the collection efficiency of the recording objective was constant. The illumination profile of the fiber was reconstructed from sequential intensity measurements at different sites along the plane containing the fiber axis.

2.4.

Optical Fibers

The detection volumes of several different optical fiber-types were characterized. The flat-cleaved fibers used were: a single-mode (SM) fiber with core diameter 4.0μm and NA of 0.12 (FS-SN-3224, 3M), multimode (MM) graded index fibers with core diameters of 50 and 105μm and NA of 0.22 (AFS50/125Y, AFS105/125Y, Thorlabs, Inc.), and an MM step-index fiber with core diameter 100μm and NA of 0.22 (F-MCB-T, Newport Corp.). A more complex optical probe designed for multisite intratissue fluorescence mapping9 was also characterized. This probe consists of angle-cleaved fibers (F-MCB-T, Newport Corp.) enclosed in a glass micropipette.

2.5.

Solution and Tissue Preparation

The probes were characterized in both nonscattering and scattering solutions, as well as in stained tissue. Nonscattering solution containing 500μML rhodamine was used for most two-dimensional (2-D) confocal scanning studies, because the SNR was optimized at this fluorophore concentration. For flash photolysis studies, 1.67mML of 10,000 MW 4,5-dimethoxy-2-nitrobenzyl (DMNB)-caged fluoroscein in dextran (Molecular Probes Corp.) were used. Scattering solutions were prepared by adding nonfluorescent 1μmdiam microspheres at a concentration of 5.06×109beadsmL to these aqueous solutions, resulting in scattering properties that mimic brain tissue.22 In addition, 2-D confocal scanning studies were carried out for a flat-cleaved AFS105/125Y fiber in nonscattering solution at rhodamine concentrations ranging from 50 to 5000μML to determine the effects of fluorophore concentration on fluorescence collection fields.

The tissue used in this study was taken from the hearts of 20-month-old Wistar Kyoto (WKY) rats. The rats were anaesthetized with halothane, decapitated, and the hearts were rapidly removed. Tissue slices ( 1mm thick) were cut from the free wall of the left ventricle (LV), oblique to the epicardial surface, and stained with rhodamine (1mML) for 24h at 4°C . Subsequently, fiber optic probes were inserted manually into the section thickness and 2P excitation studies were carried out. All procedures were approved by the University of Auckland Animal Ethics Committee.

2.6.

Flux Distribution Model for Illumination Field

One of the objectives of this study is to develop a model that can be used to estimate the illumination and effective fluorescence collection volumes in tissue for MM optical fibers of varying diameter and NA. In order to address this problem, we have set up a semiempirical model similar to that presented by Zhu and Yappert.17

The illumination intensity distribution at a given wavelength is represented as:

Eq. 6

Ie(r,z)=E(r,z)×D(z),
where energy loss is lumped in the axial decay component, D(z) and E(r,z) is a source term for which energy conservation applies. Thus, for all z

Eq. 7

0E(r,z)rdr=constant,
we assume that the decay is exponential and is expressed as:

Eq. 8

D(z)=exp(μt,z),
where the effective absorption coefficient, μt , is a wavelength-dependent space constant that reflects the loss of photons due to absorption and scattering of photons from within the measurement region.

The illumination source function E(r,z) is the convolution of a simple Gaussian distribution S1(r,z) , representing a single-mode illumination field, with a circular function S2(r,z) and is expressed as:

Eq. 9

E(r,z)=S1(r,z)*rS2(r,z),
where *r indicates that the convolution is performed only over the radial coordinate. The Gaussian function

Eq. 10

S1(r,z)=12πδ2exp(r22δ2),
has a standard deviation δ that varies with z such that most of the light is contained within a cone defined by the NA of the fiber. Note that the boundaries set by the NA have a threshold level of 5% of the peak intensity as specified by the fiber manufacturer (see Fig. 3 ). Therefore, the standard deviation δ is defined as:

Eq. 11

δ=[(ztanθ)22ln(0.05)]12,
where θ is determined from the NA of the fiber (NA) and the refractive index of the medium (n) :

Eq. 12

θ=sin1(NAn).

Fig. 3

Cross-sectional representations of the simple Gaussian function S1(r,z) and the spatial source term S2(r,z) that are convolved to form the illumination source function E(r,z) . Note that functions are normalized to ensure flux conservation.

034033_1_035703jbo3.jpg

The function S2(r,z) describes the circular area over which the Gaussian function S1(r,z) is convolved (in the radial coordinate) to generate the radial distribution of illumination at each axial position, and it incorporates a scale factor that ensures that energy is conserved as specified in Eq. 7.

Eq. 13

S2(r,z)={R02(r1+R0)2,ifr<r1+R0,0,otherwise,
where R0 is the radius of the fiber and r1=ztanθb .

When θb=0 , S2 is independent of z and E corresponds to a distribution of essentially independent single modes originating at the face of the MM fiber. Under these circumstances, the predicted axial profile of the illumination intensity E(r,z) is constant in the near-field, when the effective absorption coefficient μt=0 . This does not match the fall in axial intensity in the near-field observed either in air or in nonscattering fluids (see Sec. 3), due, presumably, to redistribution of light in the radial direction that is not encompassed by the spread of a Gaussian function alone. To more closely match the fall of axial illumination in the near-field, we have allowed, as a first-order approximation, the radius of S2 to vary linearly with z . Note that, in our simulations, θb is relatively small in comparison with θ .

A similar model should also describe the fluorescence collection efficiency at a given emission wavelength. However, because fluorescent emission is distributed across a range of wavelengths, the fluorescent light collected from a point (r,z) will be affected by wavelength variation in both the intensity of emission and the decay component D(z) . To account for this, the collection efficiency Ic should be evaluated as follows:

Eq. 14

Ic(r,z)=kλlλuE(r,z)×F(λ)×D(z,λ)dλ,
where k is a scaling factor, and λu and λl are the upper and lower bounds, respectively, of the effective emission spectrum F(λ) after filtering. It can be shown that this expression reduces to a form very similar to that in Eq. 6:

Eq. 15

Ic(r,z)=E(r,z)×D*(z),
where scaling constant and wavelength dependence are lumped in the weighted axial component D*(z) .

This model has been used to interpret and analyze experimental data and to investigate systematically the effects of fiber diameter and NA on both illumination and effective fluorescence collection volumes. Effective fluorescence collection volumes are estimated as the product of illumination and fluorescence collection fields.

The model has also been used to study the extent to which a wave of altered fluorescence emission propagating through the volume addressed by a fiber will be blurred by varying contributions of regions within the collection field. This process is simulated by allowing a plane-wave impulse to propagate at various angles and propagation speeds through the effective collection field defined by the model. At any time, the fluorescence signal, dF(r,z) , is equal to the effective collection field [see Eq. 4] along the impulse wavefront and is zero elsewhere. The total fluorescence signal was estimated by integrating across the entire volume as indicated in Eq. 3. This process was repeated sequentially to reconstruct the time-course of the optical signal collected as the plane-wave impulse moves across the volume addressed by the fiber.

3.

Results

3.1.

Comparison of Illumination and Fluorescence Collection Profiles

Flash photolysis was used to characterize illumination and collection fields under identical experimental conditions for a flat-cleaved fiber (F-MCB-T) in nonscattering and scattering media. Illumination and collection profiles mapped out along the fiber axis using 2P flash photolysis are superimposed in Fig. 4a on the axial collection profile measured in nonscattering rhodamine solution (100μML) using 2P confocal microscopy. Note that these data were obtained with a 488-nm excitation source and that the fluorescence emission was collected at 535nm . It is evident from Fig. 4a that there is no difference between axial illumination and collection profiles obtained with flash photolysis and that the axial collection data obtained with 2P confocal scanning in comparable conditions are consistent with these results.

Fig. 4

Comparison of illumination and fluorescence collection profiles for an F-MCB-T optical fiber in nonscattering fluid. (a) Axial illumination and fluorescence collection measurements: illumination () and collection () profiles obtained with 2P flash photolysis are compared with collection profile obtained with 2P confocal scanning (—). (b) Radial illumination and fluorescence collection measurements: illumination () profile obtained with 2P flash photolysis is compared with collection profile obtained with 2P confocal scanning (—) under comparable conditions.

034033_1_035703jbo4.jpg

The radial illumination profile in Fig. 4b, measured with 2P flash photolysis, exhibits a flat top over the central region and decays toward the periphery, almost exactly matching the corresponding collection profile obtained with 2P confocal scanning with the same fiber in nonscattering fluid. Results observed in scattering fluids were also similar to those presented here. SNR was relatively low in these experiments. This was due to decreased signal levels, but traces of incompletely caged fluorophore also increased the background noise.

3.2.

Fluorescence Collection Fields for Flat-Cleaved Optical Fibers

Fluorescence collection fields of optical probes were measured with 2P confocal microscopy as described earlier. A typical 2-D collection field along the axis of an MM flat-cleaved fiber in nonscattering fluid is shown in Fig. 5a . The axial intensity profile is shown in Fig. 5b, while radial intensity profiles at z=50 , 200, and 350μm are superimposed in Fig. 5c. The axial intensity falls progressively with distance from the fiber tip. However, the rate of axial decay appears to be less in the near-field than the far-field. The radial intensity distribution has a flat top adjacent to the tip of the fiber. This plateau is reduced as light spreads radially with increasing axial distance from the fiber tip. In the far-field, the radial intensity profile resembles a Gaussian distribution. The transition between these forms takes place around 250μm from the fiber tip in this case, and in general, the axial distance at which this occurs is directly related to the fiber radius.

Fig. 5

Fluorescence collection field for an F-MCB-T optical fiber characterized using 2P excitation. (a) Measured 2-D intensity map and the corresponding model fitting. (b) Axial intensity profile. (c) Radial intensity profiles at axial locations indicated in (b) 50, 200, and 350μm from the fiber tip. Note that results were obtained with nonscattering 500μML rhodamine solution. The solid lines in (b) and (c) were fitted using the model.

034033_1_035703jbo5.jpg

Typical collection fields for a MM fiber in scattering fluid and tissue are shown in Fig. 6a . Corresponding axial intensity profiles in nonscattering and scattering fluids and in stained tissue are overlaid in Fig. 6b. Note that the data here are smoothed using a moving average filter. There is significant attenuation of fluorescence collection intensity in scattering fluid and in tissue. However, the flat top of the radial intensity distribution adjacent to the fiber tip was preserved in all cases. It should be noted that the axial intensity profile in scattering medium exhibits considerably more decay than the corresponding profile in the nonscattering fluid.

Fig. 6

Fluorescence collection field for an F-MCB-T optical fiber characterized using 2P excitation in different media. (a) 2-D intensity maps for scattering fluid and stained tissue and the corresponding model fittings. (b) Axial intensity profiles for tissue, nonscattering fluids, and scattering fluids. Note that results obtained with nonscattering and scattering fluids contained 500μML rhodamine, while the tissue was stained with 1000μML rhodamine. The solid lines in (b) are the corresponding fitting results using the model.

034033_1_035703jbo6.jpg

3.3.

Model Validation

The semiempirical model described in the Sec. 2 was validated by comparing predicted collection profiles with data obtained experimentally. Because fluorescent emission was collected across a relatively narrow wavelength window (535±25nm) , it was assumed that the weighted axial component D*(z) in Eq. 15 could be represented by a simple exponential expression as in Eq. 8. For a given MM fiber, the parameter θb that controls the spread of the function S2(r,z) [see Eq. 13] was set as follows. Near-field axial intensity profiles observed in nonscattering fluid were matched to the model predictions by adjusting θb , having first estimated the effective absorption coefficient μt in this medium from measured photon loss. The effective absorption coefficient μt that best fit the model to experimental data was then determined for scattering fluid and tissue. For the AFS105/125Y fibers, this process was also carried out for nonscattering fluids at a series of rhodamine concentrations ranging from 50 to 1000μML . Best-fit values for the effective absorption coefficient μt are given in Table 1 .

Table 1

Values of effective absorption coefficient that best fit model to experimental data. The average RMS error for the fit between model and data was 0.058±0.025 and 0.072±0.023 , for axial and radial profiles, respectively. The parameter θb was set to 2deg in all cases. The value indicated by * is the concentration of dye used to stain the tissue, not the actual dye concentration in the tissue.

Fiber diameter (μm) Medium typeDye concentration (μM∕L) Effective absorption coefficient, μt (mm−1)
50Scattering fluid5003.4
50Nonscattering fluid5001.7
100Scattering fluid5003.4
100Nonscattering fluid5001.6
100Tissue 1000* 3.5
105Scattering fluid5003.5
105Nonscattering fluid500.7
105Nonscattering fluid1000.8
105Nonscattering fluid5001.6
105Nonscattering fluid10002.2

Similar results were observed in the other flat-cleaved MM fibers tested. The average root-mean-square (RMS) error for the fit between the model and experimental data was 0.058±0.025 and 0.072±0.023 for axial and radial profiles, respectively. Background due to dark current was evident at low dye concentrations, but this was not subtracted from the experimental data.

3.4.

Quantification of Effective Collection Volumes

The effective fluorescence collection efficiency Ieff is defined in Eq. 4 as the product of fluorescence collection and illumination efficiencies. Because illumination and fluorescence collection fields are very similar, Ieff can be approximated as the square of the measured collection efficiency distribution Ic and used to construct the relative fluorescence signal R(z) —see Eq. 5.

Figure 7 shows relative fluorescence signals for a flat-cleaved F-MCB-T fiber in nonscattering and scattering fluids and in tissue. The parameters Z50 and Z80 , which are defined as the axial depths at which R=50 and 80%, respectively, are used to quantify the effective collection volume of the fiber. Values for Z50 in stained tissue, scattering fluid, and nonscattering fluid are 75, 95, and 220μm , respectively. Corresponding values of Z80 are 170, 195, and 465μm .

Fig. 7

Relative fluorescence signals for an F-MCB-T optical fiber in different media. Z50 in stained tissue, scattering fluids, and nonscattering fluids are 75, 90, and 220μm , respectively. Corresponding values of Z80 are 175, 195, 90, and 465μm . The fluorescence emission was measured at 535±25nm .

034033_1_035703jbo7.jpg

Table 2 presents Z50 and Z80 for MM fibers of various diameters in nonscattering fluid. The radial FWHM of the effective collection profile at Z50 is also given. The results indicate that effective fluorescence collection volume of an MM fiber is directly related to its core diameter. While the data in Table 2 were acquired in nonscattering fluids at an emission wavelength of 535±25nm , comparable results were also obtained in scattering fluid and at different emission frequencies. Z50 increased by less than 10% at 600nm and was approximately halved in scattering fluid.

Table 2

Effective collection dimensions for MM optical fibers of different diameter in nonscattering fluid at a rhodamine concentration of 500μM∕L . Note that fluorescence emission was measured at 535±25nm .

Fiber Diameter (μm) Medium type Z50 (μm) Z80(μm) FWHM at Z50 (μm)
50Nonscattering fluid12524040
100Nonscattering fluid22046595
105Nonscattering fluid240550100

3.5.

Effective Fluorescence Collection Volume of Purpose-Built Optical Probe (Optrode)

The fluorescence collection profile of a purpose-built optical probe (optrode), in which hexagonally packed, angle-cleaved F-MCB-T fibers are enclosed in a glass micropipette, was also measured.

The fluorescence collection field in nonscattering fluid for a single optrode channel is presented in Fig. 8a . The collection field is oriented radial to the probe axis and is more complex (and less symmetric) than for a flat-cleaved MM fiber. In Fig. 8b, axial profiles for a single optrode channel in nonscattering fluid and tissue are compared with the profile in nonscattering fluid for a flat-cleaved F-MCB-T fiber. Effective collection dimensions for a single optrode channel were estimated from three-dimensional (3-D) two-photon excitation data. These results are given in Table 3 .

Fig. 8

Fluorescence collection field for a purpose-built fiber optic probe (optrode) characterized using 2P excitation. (a) 2-D intensity map in nonscattering fluid. (b) Axial intensity profiles for optrode in tissue and nonscattering fluid compared with profile for a flat-cleaved F-MCB-T optical fiber in nonscattering fluid. Note that the optrode is constructed from angle-cleaved F-MCB-T fibers (for further details see Sec. 2).

034033_1_035703jbo8.jpg

3.6.

Effective Collection Volume and Propagated Impulse Response

The volume from which fluorescence is collected determines the spatial resolution of functional fluorescence measurements. However, when propagating events such as calcium waves or electrical depolarization are being mapped, the time-course of the fluorescence signal may be blurred by the spread of the wave across the collection volume. To investigate this, we have simulated the temporal response of an impulse of fluorescence emission propagating as a plane-wave across optical fiber collection volumes of varying dimensions (see Sec. 2). The results of this study are presented in Fig. 9 .

Fig. 9

Simulated fluorescent response of impulse propagating as a plane-wave across a fiber collection volume. (a) Impulse response for propagation at 0deg , 45deg , and 90deg with respect to the fiber axis ( Z50=80μm , fiber diam=100μm , and propagation velocity=0.4ms ). (b) FWHM of the impulse response as a function of fiber diameter ( Z50=80μm and propagation velocity=0.4ms ). (c) FWHM of the impulse response as a function of Z50 (fiber diam=100μm and propagations velocity=0.4ms ). Results for the propagation directions 0deg and 90deg are superimposed in both (b) and (c).

034033_1_035703jbo9.jpg

Table 3

Effective collection dimensions for a single optrode channel in nonscattering fluid and tissue.

Fiber typeMedium type Z50 (μm) Z80 (μm) FWHM at Z50 (μm)
OptrodeNonscattering fluid10523065
OptrodeTissue7215260

Figure 9a shows a simulated fluorescent signal propagating across a fiber collection volume at 0deg , 45deg , and 90deg with respect to the fiber axis. Propagation across the collection volume broadens the fluorescent signal, and this blurring varies with impulse direction. The broadening is quantified by estimating the FWHM of the impulse response. FWHM is presented as a function of fiber diameter in Fig. 9b ( Z50=80μm and propagation velocity=0.4ms ) and as a function of Z50 in Fig. 9c (fiber diameter=100μm and propagation velocity=0.4ms ). Results for the propagation directions 0deg and 90deg are superimposed in both cases.

As expected, these results show that the blurring introduced by propagated fluorescent activity increases as the collection volume is enlarged. However, for a propagation velocity of 0.4ms (a velocity typical for spread of electrical excitation through cardiac tissue), the FWHM of the impulse response was less than 0.5ms across the range of fiber diameters and Z50 values tested. Last, sensitivity to the direction of propagation depends on the effective collection dimensions. With a fiber diameter of 100μm and Z50 equal to 80μm , the effective collection volume is approximately isotropic, and the fact that the FWHM of the impulse response shown in Fig. 9a is not affected significantly by the propagation direction under these circumstances reflects this.

4.

Discussion

The primary objective of this study was to identify both the illumination and fluorescence collection volumes of optical fibers in different fluid media, as well as in tissue. Systematic analyses were carried out with a variety of optical fibers employing novel methods that provide high spatial resolution. These include 2P excitation of rhodamine to measure fluorescence collection fields and 2P release of caged fluoroscein combined with confocal scanning microscopy to estimate illumination fields. Fluorescence collection fields have been characterized for flat-cleaved optical fibers and related to core diameter, NA, and optical properties (scattering and absorption) of the medium. We have also demonstrated that illumination and fluorescence collection profiles for MM fibers are essentially the same at comparable wavelengths. A semiempirical model was developed to describe the illumination and fluorescence collection fields for flat-cleaved MM fibers, and this has been used to interpret the experimental data. For an MM optical fiber with a core diameter of 100μm and NA=0.22 in tissue, 80% of the total fluorescence in a wavelength window 535±25nm excited by 488-nm illumination is collected from a volume that is approximately 170μm deep and 105μm in diameter. In conjunction with the model, these data provide a basis for designing fiber optic probes, with specific excitation and effective collection profiles. As far as we are aware, this is the first time that illumination and collection profiles of optical fibers have been characterized systematically at high spatial resolution in a range of media, including soft tissue.

4.1.

Characterization of Illumination and Fluorescence Collection Fields

The experimental techniques employed in this study to determine the illumination and fluorescence collection profiles incorporate a number of novel aspects that add to the precision of this study. First, 2P excitation illuminates an ellipsoidal volume that is small (<1%) in comparison to the diameter of the optical fiber and can therefore be considered to be a point source. Because the 3-D location of the 2P focal volume can be precisely controlled, it is possible to reconstruct the fluorescence collection fields of optical fibers and fiber optical probes with high spatial resolution. Second, the 2P excitation wavelength (750nm) is well separated from the spectral range of fluorescent emission for both rhodamine and fluoroscein. Last, with 2P release of caged fluorophores, it is possible to sample fiber illumination profiles in a fashion similar to that used to characterize fluorescence collection fields. Moreover, the use of flash photolysis also means that confounding effects of light absorption by the fluorophore are minimized, since caged fluorophores outside the region addressed by the 2P focal volume do not respond to excitation. In any case, this study demonstrates that the three-dimensionally resolved nature of 2P excitation allows accurate characterization of illumination and collection fields of arbitrary probes or other optical elements as long as the fields are moderately larger than achievable 2P point spread functions (<1μm) .

The ellipsoidal shape of the 2P focal spot may affect the resolution with which illumination and fluorescence collection volumes can be characterized experimentally. With a 40× 0.9 NA objective, the FWHM of the 2P focal volume was 0.5μm in the xy plane but 1.4μm in the z -axis direction, and for lower magnification objectives, where NA is reduced, elongation in the z direction was proportionately worse. This has little impact on the axially symmetric illumination and fluorescence collection fields associated with the individual optical fibers studied here. In this case, it can be shown that the xy resolution of the objective applies, except in the near-field. No significant differences were observed between images obtained with the 10× and a 40× objective, and we therefore used a 10× objective since it covered a significantly larger field of view. However, the spread of the 2P focal volume in the z direction does give rise to signal blurring when characterizing illumination and collection fields that are not axially symmetric. The resultant degradation of resolution can always be overcome by using high NA objectives and employing image tiling to cover the required field of view.

4.2.

Effective Collection Volume

The fact that measured illumination and fluorescence collection fields for flat-cleaved MM fibers were very similar in both nonscattering and scattering fluids is noteworthy. Absorption and scattering are wavelength-dependent, and the wavelengths of excitation light and fluorescent emission are different. In this study, however, fluorescence was collected over a relatively narrow wavelength window close to the excitation wavelength. Moreover, we have shown that collection volumes are strongly determined by scattering, which is more weakly dependent on wavelength than absorption (see the following discussion). Within this context, it is not surprising that the techniques employed could not resolve differences between illumination and collection volumes. We have also independently reconstructed the 3-D illumination field of an optical fiber in air by imaging light intensity scattered from a thin film transverse to it at a series of different axial locations. A confocal microscope in wide-field mode was used for this purpose. The results (not shown) match the detailed fluorescence collection fields obtained with 2P excitation when absorption and scattering are minimized. These results provide objective justification for the reconstruction of effective fluorescence collection volumes from measured fluorescence collection fields using the relationship [see Eq. 1]:

Eq. 16

dF(r,z)=ηKIc(r,z)2dV.

4.3.

Comparison with Previous Work

A number of studies of effective fluorescence collection dimensions have been published. 7, 9, 10, 13, 15, 23, 24, 25, 26 The effect of the propagation of cardiac electrical activation across the effective collection volume on measured optical potentials has also been considered, 13, 15, 25, 26 with reported collection depths in tissue ranging from 100μm (Ref. 9) to > 1mm (Ref. 25). In part, this variance reflects the different means of fluorescence excitation and collection used. Focusing on previous results that are most comparable with ours, Krauthamer 7 measured illumination intensity profiles for SM and MM optical fibers in air but report significantly greater axial attenuation than we do and observed no significant effect of fiber diameter on collection depth. Like us, Byars 10 characterized the effective collection volumes of MM optical fibers in cardiac tissue. Their results for a flat-cleaved fiber with core diameter 240μm and NA 0.51 are consistent with a value of Z50 around 800μm , while we would expect Z50 to be less than 300μm for an optical fiber with these specifications in tissue. Byars characterized axial attenuation by passing a fiber through an excised LV wall segment from the endocardial surface until it touched a layer of absorbent paper soaked with fluorophore attached to the epicardial surface. Measurements were then made at a number of points through the LV as the fiber was withdrawn. We note that their axial steps were very coarse (500μm) , and it is unknown to what extent the optical properties of tissue might have been affected by the track previously made by the fiber.

Ding 24 used Monte Carlo simulations to investigate the effects of illumination and collection diameters on effective collection depth. Their predicted Z50 for illumination and collection diameters of 100μm was around 200μm , which closely matches our results with optical fibers of comparable dimension. Hooks9, 27 measured effective collection volumes for flat-cleaved MM optical fibers in nonscattering fluid by recording the fluorescence collected as the fiber was moved toward the edge of a cuvette containing dye solution. The results— Z50 was 140μm and 224μm , respectively, for MM optical fibers with core diameters of 65μm and 100μm —are very similar to those given here. Knisley 23 employed a similar approach to estimate effective fluorescence collection depths in cardiac tissue. They used a 100μm laser beam to excite regions of different tissue thickness along a wedge of LV myocardium stained with di-4-ANEPPS and stated that the thickness of tissue contributing to fluorescence was around 300μm . In comparison, the Z80 reported here for a 100μm MM fiber is 170μm . This difference is due to the fact that Knisley collected fluorescence emission with a PMT, which has a much larger detection area than the optical fiber used in our case. When the effects of collection dimension are accounted for using the theoretical analysis outlined earlier,24 their results are entirely consistent with ours.

4.4.

Validation of the Model

The model outlined in this paper provides a semiempirical description of the illumination field adjacent to the tip of a flat-cleaved MM fiber. The approach employed has a number of similarities to a model previously presented by Zhu and Yappert.17 These investigators represented the distribution of light radial to the fiber axis as a Gaussian function with a standard deviation that was a linear function of axial position. As in our model, losses due to absorption and scattering were lumped as a single axial exponential decay. Such a model could not account for the complexity of the illumination field observed near the tip of an MM fiber, and Zhu and Yappert17 acknowledged that “for multi-mode fibers a multi-Gaussian distribution should be used.” In our model, we have constructed such a distribution by convolving a Gaussian function (representing single-mode propagation) with a circular source area transverse to the fiber axis. The diameter of the radial distribution source is identical to that of the fiber core at its tip but is also a linear function of axial position, which enabled us to accommodate redistribution of light in the radial direction that is not encompassed by the spread of a Gaussian function alone.

The model provided a good fit to the fluorescence collection fields measured using 2P excitation and fit the 2P flash photolysis data equally well. The radial distribution function accurately represented experimental intensity profiles transverse to the fiber axis in near- and far-fields. In particular, it reproduced the flat-topped radial intensity distribution adjacent to the fiber tip and the Gaussian distribution distant from it. It also predicted the axial distance at which the transition between these two profiles was observed. However, when μt was low, there were small but systematic differences between observed and predicted axial intensity profiles. Under these circumstances, the rate of axial attenuation was slightly less in the near-field than in the far-field, but this is not completely matched by the model [see Fig. 5b]. Residual asymmetries that are probably the result of modal fluctuation,28, 29, 30 and are most evident in radial intensity profiles, also cannot be reproduced by the model [see Fig. 5c]. For individual model fits, however, the RMS error was on the order of the regional variation introduced by modal fluctuation and optical scattering. Taken together, these observations confirm the robustness of the model.

4.5.

Scattering and Absorption

Tissue exhibits marked scattering of light and weaker frequency-dependent light absorption. Optical scattering is a complex process, and this is reflected by the fluorescence collection profiles presented in Fig. 6 for a flat-cleaved fiber (F-MCB-T) in scattering fluid and stained tissue. There is much greater regional variability in the fluorescence collection field for the scattering fluid [see Fig. 6a] than in equivalent results for a nonscattering fluid [see Fig. 5a]. It is also evident from these figures that fluorescent light originating from outside the NA of the fiber has a higher probability of being collected in the scattering fluid than in the nonscattering fluid. The scattering fluids prepared for this study provide an isotropic scattering environment,22 whereas scattering in tissue is generally highly anisotropic. For the data shown in Fig. 6a, at least, there appears to be greater regional variability for the stained tissue than for scattering fluid.

Despite the complexity of these processes, it has been suggested that optical scattering can be treated as an isotropic “absorption-like” event, provided that the dimensions of the volume of interest are greater than the scattering mean-free-path length.17, 19 According to Patterson and Pogue,19 the mean-free-path-length in biological tissues is 71μm at 500nm . On this basis, scattering was lumped with absorption as a single axial decay term in our model. While scattering markedly affects fluorescence collection characteristics and gives rise to substantial regional variability (Fig. 6), the model nonetheless provides a good estimate of the average intensity distribution. This validates the assumption that optical scattering in large volumes can be approximated as isotropic absorption.

While it is difficult to quantify actual fluorophore concentration in stained tissue, dye absorption will likely reduce effective collection volumes for such membrane potential sensitive dyes as di-4-ANNEPS, which is typically administered at relatively high concentrations.12 Effective collection volume is therefore reduced at high fluorophore concentrations, but this can be ignored at concentrations <100μML . Ding have measured the optical properties of di-4-ANEPPS stained rabbit heart and report values for absorption μa and scattering μs of 0.52±0.13mm1 and 23.0±2.5mm1 , respectively, at 488nm . Corresponding values at 669nm were μa=0.10±0.04mm1 and μs=21.8±2.5mm1 (Ref. 24). Our estimates of total absorption for cardiac tissue were 3.5mm1 , which is much larger than the value for absorption μa quoted earlier, but significantly less than quoted scattering coefficients indicating that some scattered light is collected by the fiber optic probe. This suggests that scattering plays an important role in the axial decay of effective fluorescence collection efficiency for fiber optic probes in tissue. This view is reinforced by the observation that very similar estimates of total absorption were made for probes in scattering fluid at the same dye concentration.

4.6.

Determinants of Effective Fluorescence Collection Volumes in Fiber Optic Probes

The experiments and modeling studies reported here demonstrate that the dimensions of the effective fluorescence collection volume for a flat-cleaved optical fiber are mainly determined by its core diameter and by the optical properties of the medium in which it is used. We have shown that a change in fiber diameter from 50 to 100μm gives rise to changes of similar magnitude in collection volume dimensions (see Table 1). However, we found that the influence of fiber diameter on effective collection volume was reduced when the axial decay constant was high, for instance, in cardiac tissue. These experimental observations are matched by the model, which also predicts similar trends when the fiber diameter is increased from 100 to 200μm . While the effect of altering NA was not considered experimentally, the model predicts a relatively modest reduction in Z50 when NA is doubled—less than 10% in a nonscattering fluid and less still in scattering fluids and tissue.

The fluorescence collection field of a purpose-built fiber optic probe (optrode) that consists of multiple angle-cleaved MM fibers enclosed in a glass micropipette9 was similar in shape but significantly smaller. We ascribe these differences to lensing effects as light passes through the curved surface of the fiber and also to scattering at interfaces in the probe, where there is a refractive index mismatch.

4.7.

Effects of Propagation on Fluorescence Impulse Response of Optical Probes

It has been noted previously that the upstroke of the cardiac action potential observed in optical recordings is generally much slower than rapid depolarization in corresponding intracellular measurements.12, 13 This has been attributed to the spread of electrical activation through the region addressed by the imaging system. 13, 15, 16, 31, 32 According to our analysis, propagation at 0.4ms produces blurring of the impulse response that is dependent on the direction of propagation and the dimensions of the effective collection volume. However, the FWHM for the impulse response in tissue was less than 0.5ms for an optical fiber with a core diameter of 100μm and Z50 of 80μm , regardless of propagation direction. It is noteworthy that directional sensitivity of fiber optic probes is minimized when the collection volume is relatively isotropic. However, it would be possible to design probes with specific directional sensitivity on the basis of the information provided in this paper. Electrical activation propagates through a normal ventricular myocardium at 0.2 to 1ms (Refs. 33, 34, 35), and rise-times of ventricular action potentials recorded using fiber optic probes with core diameters of 100μm range from 5 to 10ms (Refs. 7, 9, 10). Less than 1ms of this upstroke time can be explained by the spread of activation across the relatively small collection volume of these probes. This discrepancy is perplexing and should be subjected to more-detailed investigation in the future.

4.8.

Limitations of the Study

There are some obvious limitations to the results presented in this study, and these will now be discussed. First, the optical fibers tested all had the same NA. While the model suggests that NA has a limited affect on Z50 in tissue, this prediction has yet to be tested experimentally. Second, the experimental results presented here were obtained at a limited number of wavelengths. A variety of excitation wavelengths are used with functional fluorescence recordings, and emission spectra for most fluorophores occupy a relatively broad range of wavelengths. In addition, the absorption of light in tissue is strongly frequency dependent, whereas scattering is much less so.18, 19 However, we have demonstrated that effective collection volumes in tissue are dominated by scattering for fiber optic imaging systems (although this is less so for systems with larger collection dimensions). We have also shown that the variation of μt with wavelength is relatively small, and should therefore be possible to account for using a simple linear correction. A further difficulty is that estimates of μt at low dye concentrations are based on fluorescence recordings with low SNR. The reliability of the absolute μt values presented here may therefore be limited by confounding effects of dark-noise. Last, while the model used in this study was useful and relatively robust, it is nonetheless semiempirical. To represent illumination and fluorescence collection fields more completely, it will be necessary to use a representation that is based firmly on the fundamental principles of physical optics.

5.

Conclusions

In this study, we have used powerful new techniques to characterize the illumination and fluorescence collection volumes of optical fibers at much higher spatial resolution than previously reported. Systematic analyses were carried out with a variety of fiber optic probes in tissue and different fluid media. In addition, a semiempirical model was developed to describe the illumination and fluorescence collection fields for flat-cleaved optical fibers. This has been used to interpret the experimental data and also provides a foundation for the design of fiber optic probes. On the basis of this work, we have developed a firm understanding of the way in which fiber geometry, absorption, and scattering of light in the medium determine effective fluorescence collection fields and time resolution in propagation measurements.

Acknowledgments

This work was funded by the New Zealand Marsden Fund and by the Health Research Council of New Zealand. Dean Tai was supported by a University of Auckland Doctoral Scholarship and was also a recipient of a Richard Bates Memorial Scholarship from the Royal Society of New Zealand.

References

1. 

Y. Chen, G. Zheng, Z. H. Zhang, D. Blessington, M. Zhang, H. Li, Q. Liu, L. Zhou, X. Intes, S. Achilefu, and B. Chance, “Metabolism-enhanced tumor localization by fluorescence imaging: in vivo animal studies,” Opt. Lett., 28 2070 –2072 (2003). 0146-9592 Google Scholar

2. 

R. Ackroyd, C. Kelty, N. Brown, and M. Reed, “The history of photodetection and photodynamic therapy,” Photochem. Photobiol., 74 656 –669 (2001). https://doi.org/10.1562/0031-8655(2001)074<0656:THOPAP>2.0.CO;2 0031-8655 Google Scholar

3. 

F. Gad, T. Zahra, K. P. Francis, T. Hasan, and M. R. Hamblin, “Targeted photodynamic therapy of established soft-tissue infections in mice,” Photochem. Photobiol. Sci., 3 451 –458 (2004). https://doi.org/10.1039/b311901g 1474-905X Google Scholar

4. 

S. Rohr and J. P. Kucera, “Optical recording system based on a fiber optic image conduit: assessment of microscopic activation patterns in cardiac tissue,” Biophys. J., 75 1062 –1075 (1998). 0006-3495 Google Scholar

5. 

I. R. Efimov, V. P. Nikolski, and G. Salama, “Optical imaging of the heart,” Circ. Res., 95 21 –33 (2004). https://doi.org/10.1161/01.RES.0000130529.18016.35 0009-7330 Google Scholar

6. 

C. Du, A. P. Koretsky, I. Izrailtyan, and H. Benveniste, “Simultaneous detection of blood volume, oxygenation, and intracellular calcium changes during cerebral ischemia and reperfusion in vivo using diffuse reflectance and fluorescence,” J. Cereb. Blood Flow Metab., 25 1078 –1092 (2005). https://doi.org/10.1038/sj.jcbfm.9600102 0271-678X Google Scholar

7. 

V. Krauthamer, H. J. Bryant, C. C. Davis, and T. W. Athey, “Action potential-induced fluorescence changes resolved with an optical fiber carrying excitation light,” J. Fluoresc., 1 207 –213 (1991). https://doi.org/10.1007/BF00865245 1053-0509 Google Scholar

8. 

M. Neunlist, S. Z. Zou, and L. Tung, “Design and use of an optrode for optical recordings of cardiac action-potentials,” Pfluegers Arch., 420 611 –617 (1992). https://doi.org/10.1007/BF00374641 0031-6768 Google Scholar

9. 

D. A. Hooks, I. J. LeGrice, J. D. Harvey, and B. H. Smaill, “Intramural multisite recording of transmembrane potential in the heart,” Biophys. J., 81 2671 –2680 (2001). 0006-3495 Google Scholar

10. 

J. L. Byars, W. M. Smith, R. E. Ideker, and V. G. Fast, “Development of an optrode for intramural multisite optical recording of Vm in the heart,” J. Cardiovasc. Electrophysiol., 14 1196 –1202 (2003). https://doi.org/10.1046/j.1540-8167.2003.03203.x 1045-3873 Google Scholar

11. 

D. C. S. Tai, B. J. Caldwell, I. J. LeGrice, D. A. Hooks, A. J. Pullan, and B. H. Smaill, “Correction of motion artifact in transmembrane voltage-sensitive fluorescent dye emission in hearts,” Am. J. Physiol. Heart Circ. Physiol., 287 H985 –H993 (2004). 0363-6135 Google Scholar

12. 

B. J. Caldwell, I. J. LeGrice, D. A. Hooks, D. C. S. Tai, A. J. Pullan, and B. H. Smaill, “Intramural measurement of transmembrane potential in the isolated pig heart: validation of a novel technique,” J. Cardiovasc. Electrophysiol., 16 1001 –1010 (2005). https://doi.org/10.1111/j.1540-8167.2005.40558.x 1045-3873 Google Scholar

13. 

S. D. Girouard, K. R. Laurita, and D. S. Rosenbaum, “Unique properties of cardiac action potentials recorded with voltage-sensitive dyes,” J. Cardiovasc. Electrophysiol., 7 1024 –1038 (1996). https://doi.org/10.1111/j.1540-8167.1996.tb00478.x 1045-3873 Google Scholar

14. 

D. K. L. Cheng, L. Tung, and E. A. Sobie, “Nonuniform responses of transmembrane potential during electric field stimulation of single cardiac cells,” Am. J. Physiol. Heart Circ. Physiol., 277 H351 –H362 (1999). 0363-6135 Google Scholar

15. 

C. J. Hyatt, S. F. Mironov, M. Wellner, O. Berenfeld, A. K. Popp, D. A. Weitz, J. Jalife, and A. M. Pertsov, “Synthesis of voltage-sensitive fluorescence signals from three-dimensional myocardial activation patterns,” Biophys. J., 85 2673 –2683 (2003). 0006-3495 Google Scholar

16. 

D. L. Janks and B. J. Roth, “Averaging over depth during optical mapping of unipolar stimulation,” IEEE Trans. Biomed. Eng., 49 1051 –1054 (2002). https://doi.org/10.1109/TBME.2002.802057 0018-9294 Google Scholar

17. 

Z. Y. Zhu and M. C. Yappert, “Determination of effective depth and equivalent pathlength for a single-fiber fluorometric sensor,” Appl. Spectrosc., 46 912 –918 (1992). https://doi.org/10.1366/0003702924124411 0003-7028 Google Scholar

18. 

W. F. Cheong, S. A. Prahl, and A. J. Welch, “A review of the optical properties of biological tissues,” IEEE J. Quantum Electron., 26 2166 –2185 (1990). https://doi.org/10.1109/3.64354 0018-9197 Google Scholar

19. 

M. S. Patterson and B. W. Pogue, “Mathematical model for time-resolved and frequency-domain fluorescence spectroscopy in biological tissues,” Appl. Opt., 33 1963 –1974 (1994). 0003-6935 Google Scholar

20. 

C. Soeller and M. B. Cannell, “Construction of a two-photon microscope and optimisation of illumination pulse duration,” Pfluegers Arch., 432 555 –561 (1996). https://doi.org/10.1007/s004240050169 0031-6768 Google Scholar

21. 

M. B. Cannell, M. D. Jacobs, P. J. Donaldson, and C. Soeller, “Probing microscopic diffusion by 2-photon flash photolysis: measurement of isotropic and anisotropic diffusion in lens fiber cells,” Microsc. Res. Tech., 63 50 –57 (2004). https://doi.org/10.1002/jemt.10422 1059-910X Google Scholar

22. 

P. Theer, M. T. Hasan, and W. Denk, “Two-photon imaging to a depth of 1000μm in living brains by use of a Ti:Al2O3 regenerative amplifier,” Opt. Lett., 28 1022 –1024 (2003). 0146-9592 Google Scholar

23. 

S. B. Knisley, “Transmembrane voltage changes during unipolar stimulation of rabbit ventricle,” Circ. Res., 77 1229 –1239 (1995). 0009-7330 Google Scholar

24. 

L. Ding, R. Splinter, and S. B. Knisley, “Quantifying spatial localization of optical mapping using Monte Carlo simulations,” IEEE Trans. Biomed. Eng., 48 1098 –1107 (2001). https://doi.org/10.1109/10.951512 0018-9294 Google Scholar

25. 

W. T. Baxter, S. F. Mironov, A. V. Zaitsev, J. Jalife, and A. M. Pertsov, “Visualizing excitation waves inside cardiac muscle using transillumination,” Biophys. J., 80 516 –530 (2001). 0006-3495 Google Scholar

26. 

C. J. Hyatt, S. F. Mironov, F. J. Vetter, C. W. Zemlin, and A. M. Pertsov, “Optical action potential upstroke morphology reveals near-surface transmural propagation direction,” Circ. Res., 97 277 –284 (2005). https://doi.org/10.1161/01.RES.0000176022.74579.47 0009-7330 Google Scholar

27. 

D. A. Hooks, “Three-dimensional mapping of electrical propagation in the heart: experimental and mathematical model based analysis,” Univ. of Auckland, (2002). Google Scholar

28. 

D. B. S. Soh, J. Nilsson, S. Baek, C. Codemard, Y. C. Jeong, and V. Philippov, “Modal power decomposition of beam intensity profiles into linearly polarized modes of multimode optical fibers,” J. Opt. Soc. Am. A, 21 1241 –1250 (2004). https://doi.org/10.1364/JOSAA.21.001241 0740-3232 Google Scholar

29. 

R. Rokitski and S. Fainman, “Propagation of ultrashort pulses in multimode fiber in space and time,” Opt. Express, 11 1497 –1502 (2003). 1094-4087 Google Scholar

30. 

V. Doya, O. Legrand, F. Mortessagne, and C. Miniatura, “Speckle statistics in a chaotic multimode fiber,” Phys. Rev. E, 65 056223 (2002). https://doi.org/10.1103/PhysRevE.65.056223 1063-651X Google Scholar

31. 

S. F. Mironov, F. J. Vetter, and A. M. Pertsov, “Fluorescence imaging of cardiac propagation: spectral properties and filtering of optical action potentials,” Am. J. Physiol. Heart Circ. Physiol., 291 H327 –H335 (2006). 0363-6135 Google Scholar

32. 

M. J. Bishop, B. Rodriguez, J. Eason, J. P. Whiteley, N. Trayanova, and D. J. Gavaghan, “Synthesis of voltage-sensitive optical signals: application to panoramic optical mapping,” Biophys. J., 90 2938 –2945 (2006). https://doi.org/10.1529/biophysj.105.076505 0006-3495 Google Scholar

33. 

B. G. D. Duker, “Frequency-dependent effects of tocainide, quinidine, and flecainide on conduction as reflected in the rise time of the monophasic action-potential in the isolated guinea-pig heart,” Cardiovasc. Res., 25 217 –222 (1991). 0008-6363 Google Scholar

34. 

M. D. Halbach, U. Egert, J. Hescheler, and K. Banach, “Estimation of action potential changes from field potential recordings in multicellular mouse cardiac myocyte cultures,” Cell. Physiol. Biochem., 13 271 –284 (2003). 1015-8987 Google Scholar

35. 

A. O. Verkerk, M. W. Veldkamp, F. Abbate, G. Antoons, L. N. Bouman, J. H. Ravesloot, and A. C. G. van Ginneken, “Two types of action potential configuration in single cardiac Purkinje cells of sheep,” Am. J. Physiol. Heart Circ. Physiol., 277 H1299 –H1310 (1999). 0363-6135 Google Scholar
©(2007) Society of Photo-Optical Instrumentation Engineers (SPIE)
Dean Tai, Darren Hooks, John D. Harvey, Bruce H. Smaill, and Christian Soeller "Illumination and fluorescence collection volumes for fiber optic probes in tissue," Journal of Biomedical Optics 12(3), 034033 (1 May 2007). https://doi.org/10.1117/1.2750288
Published: 1 May 2007
Lens.org Logo
CITATIONS
Cited by 29 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Luminescence

Scattering

Tissue optics

Optical fibers

Light scattering

Absorption

Photolysis

Back to Top