Open Access
1 January 2003 Blood pO2 and blood flow at the optic disc
Stephane R. Chamot, Stephen D. Cranstoun, Benno L. Petrig, Constantin J. Pournaras, Charles E. Riva
Author Affiliations +

1.

Introduction

Compromised delivery of oxygen (O 2) to the tissues of the eye fundus has been implicated in a variety of ocular diseases such as diabetic and other proliferative retinopathies1 2 3 4 and possibly glaucoma.5 Studies that have investigated the oxygenation of the ocular tissues have been largely performed using O 2 -sensitive microelectrodes inserted into the eye to determine the tissue partial pressure (pO 2,tissue ) of O 2 in locally defined regions of the retina6 7 8 9 10 or the optic nerve head11 12 13 14 15 16 17 (ONH). Except for a limited number of studies investigating the pO 2 in the vitreous of human eyes during ocular surgery,18 19 20 this invasive technique has been mainly applied in animals.

The recently developed technique of phosphorescence quenching (PQ) technique by O 2 enables the measurement of the intravascular partial pressure of O 2 in blood (pO 2,blood ) in a noninvasive manner.21 With PQ, 2-D maps of ONH and retinal pO 2,blood were obtained in the eye of cats and piglets at rest and in response to hyperoxia and increased intraocular pressure22 23 24 (IOP).

As revealed by recent studies in the brain, transient changes of pO 2,blood in the cortex begin less than 1 s after a flash-induced neuronal stimulation and last25 for less than 10 s. Recent studies suggest that the neural tissues of the ONH may also exhibit such stimulus-induced variations26 27 of the pO 2,blood . The 2-D imaging method just mentioned would be too slow, however, to investigate such changes since, at the present time, at least 10 s are needed to acquire the data for calculating a pO 2,blood map.22 23

With the goal of investigating the transient pO 2,blood changes at the ONH in response to various physiological stimuli, we have, therefore, extended the application of the PQ technique to obtain a local measurement of pO 2,blood with a time constant of approximately 2.5 s. This measure was also combined with blood flow measurement by laser Doppler flowmetry (LDF).

2.

Materials and Methods

2.1.

Phosphorescence Quenching by O 2

For in vivo measurements of pO 2,blood , metalloporphyrin molecules capable of phosphorescing are bound to serum albumin before being injected intravenously into the blood stream. A brief flash illuminates the tissue of interest and excites the porphyrin molecules to a triplet state. The excited molecules return to their ground state either through a radiative process (phosphorescence) or by transfer of energy to other molecules (quenching). When measured from a solution containing a large number of quenching molecules, the phosphorescence intensity decays exponentially with a time constant (τ), which represents the time at which the initial intensity is reduced by 1/e. The only significant quenching agent present in the blood being28 O 2, τ is related to pO 2,blood through the Stern–Volmer equation: τ0/τ=1+kQτ0pO 2,blood , where τ0 is the probe lifetime in the absence of a quenching agent, and kQ the quenching constant.

2.2.

Phosphor

The oxygen probe used for PQ measurements is a Pd-meso-tetra (4-carboxyphenyl) porphine (Oxygen Enterprises, Pennsylvania). This compound must be bound to serum albumin before systemic injection. Therefore, blood is drawn from the animal (∼4 ml/kg) and the serum is extracted by centrifugation at 8000 rpm for 10 min. The probe (20 mg/kg of animal weight) is mixed with the serum, then the pH is adjusted to 7.4, and the serum/probe solution filtered with 0.2-μm sterile filter. The solution is infused intravenously (∼100 ml/h) and allowed to equilibrate with the blood before measurements proceed. The binding of the phosphorescent probe to albumin ensures that the parameters τ0 and kQ are constant over the range of physiological pH and temperature. The probe’s sensitivity to O 2 is also enhanced and its self-quenching is reduced. Since albumin does not diffuse across the wall of the retinal and ONH blood vessels due to the blood retinal barrier, PQ measurements from these tissues provide only data on O 2 dissolved in blood.

In saline solution, the probe absorbs maximally at 412 and 528 nm and its emission spectrum is centered at 706 nm. In the physiological range of pH and blood temperatures (i.e., 7.2 to 7.4 pH unit and 36 to 38 °C, respectively), τ0 decreases by ∼0.8/°C and ∼1 per 0.1 pH unit, whereas kQ increases by ∼3/°C and ∼1 per 0.1 pH unit.29 Thus, keeping the pH and temperature within the ranges already specified induces a maximum error of ∼6 when evaluating pO 2,blood based on phosphorescence lifetimes greater than 25 μs (pO 2,blood <100 mm Hg). In this study, we used29 the values of τ0 and kQ corresponding to a pH of 7.4 and a temperature of 38 °C (i.e., 637 μs and 381 mm Hg−1 s−1, respectively).

2.3.

Ocular Fundus Phosphorometer

Figure 1 schematically shows the phosphorometer built to measure pO 2,blood in the fundus of the eye. The optics of this device have been mounted on a fundus camera (Zeiss, Germany). To illuminate the fundus with the excitation light, the flash lamp normally supplied for fundus photography was replaced by a flash lamp (F) with shorter pulse width. This unit consisted of a 60-W, 2-μs flash width at half-maximal intensity (Oxygen Enterprises, Ltd., Pennsylvania). An interference filter (IF, XM-530, Corion™, United States; center wavelength=530 nm, FWMH=25 nm) placed at the lamp output transmitted the spectral portion of the flash light that matched the absorption spectrum of the phosphorescent probe. Each flash delivered an energy of ∼25 μJ/cm2 over a 30-deg area at the posterior pole. Diffuse illumination of the fundus for alignment or light adaptation of the eye was achieved through the conventional illumination path of the camera starting at the 50-W tungsten lamp (L) and traversing the beamsplitter (BS). The irradiance of the white illumination light measured at the eye fundus was 2 mW/cm2.

Figure 1

Zeiss fundus camera modified to incorporate the various optical components necessary for PQ and laser Doppler flowmeter (LDF) measurements: flash source for dye excitation (F), interference filter (IF) for spectral selection of the excitation light, fundus illumination lamp (L), beamsplitter (BS) in the illumination pathway, 670-nm laser diode (D) for LDF measurements, motorized mirror (M) to select PQ or LDF detection pathway (M), long-pass filter (LP) for spectral selection of the phosphorescent light, pinhole (PH) for spatial detection of the phosphorescent light, photomultiplier tube (PMT) for detection of the phosphorescent light, optical fiber (OF) for spatial selection of the laser light, avalanche photodiode (APD) for detection of the laser light.

005301j.1.jpg

The phosphorescent emission was collected from a localized area (Ø≅200 μm) of the optic disk through the optics used for fundus observation. This area was spatially defined by a pinhole (PH) placed in a retinal image plane of the fundus camera. The phosphorescence was detected by a photomultiplier (PMT) having its maximum sensitivity in the red portion of the spectrum (R928, Hamamatsu, Japan). A long-pass filter (LP, RG630, Schott, Germany) was mounted in front of PH to enhance the phosphorescence detection. Its transmission range was 650 to 800 nm.

The signal at the output of PMT was amplified, then digitized at 1 MHz and fed into an IBM PC compatible computer. Labview (National Instrument, Austin, Texas) software was developed to control the flash triggering, acquire and analyze the phosphorescence data and calculate the pO 2,blood . To eliminate potentially undesirable contributions to the pO 2,blood resulting from the tail of this emission, the recording of the digitized photocurrent started 40 μs after and stopped 2360 μs after triggering the excitation flash. After removing all data points below the dark noise level of the detector, the pO 2,blood was obtained based on the average of 10 to 30 successive phosphorescence decay curves obtained at rate of 30 Hz. A linear regression optimized for noise correction (weighted linear fit)30 was applied to the natural logarithm of all data points to calculate τ and pO 2,blood , which were then displayed on the Labview stripchart. The total acquisition and computation time for each value of pO 2,blood was approximately 2.5 s.

2.4.

Blood Flow Measurements

For the quasisimultaneous measurement of blood flow and pO 2,blood at the same location of the optic disk, an LDF was incorporated into the phosphorometer. The LDF technique has been previously described in detail.31 Briefly, a laser beam (670 nm) with long optical coherence is focused on a small area (Ø≅150 μm) of the optic disk, away from the larger vessels. A photomultiplier detects the portion of the laser light scattered by the red blood cells and the surrounding tissue that exits the eye pupil. The photocurrent is amplified, digitized and analyzed by a computer that calculates the relative flow (F ONH ) of the red blood cells in the ONH volume sampled by the laser beam.

The LDF system was incorporated into the phosphorometer by having the beam from a visible laser (LDM115, Imatronic, United Kingdom, wavelength=670 nm) follow the optical path of the fundus illumination system of the camera and focusing it precisely at the center of the area from which the phosphorescence was detected. The light scattered by the red blood cells and surrounding tissue in the volume sampled by the laser beam was collected by an optical fiber (OF) placed at the image plane of the illuminated site and guided to an avalanche photodiode (APD). The photocurrent was analyzed using dedicated software running on a NeXT™ computer system to calculate the LDF parameters.31

A mirror (M) could be flipped from position 1, which reflected the phosphorescence toward the PH, to position 2 that enabled the scattered laser light to reach the optic fiber. Both PH and OFs were positioned in conjugated retinal image planes and “looked” at the same retinal location. Thus, the tip of the optical fiber indicated the location where pO 2,blood was measured. The proximity of laser and peak phosphorescence wavelengths did not allow simultaneous LDF and PQ recordings.

2.5.

Animals

Experiments were performed on 11 miniature pigs (∼10 kg). All procedures conformed to the Association for Research in Vision and Ophthalmology (ARVO) statement for the use of animals in ophthalmic and vision research. After intramuscular injection of a tranquilizer for pigs (Stresnil™, Cilag Chemie AG, Switzerland), anesthesia was induced with 25 to 30 mg of Hypnodil™ (chlorhydrate of metomidate) injected into the ear vein. After arterial and venous catheterization, the pigs were curarized (Tubocurarine™, 1 mg/10 kg), intubated and artificially ventilated. During the experiment, anesthesia was maintained by continuous perfusion of Nabutal™, 4 ml/h, and Tubocurarine™, 0.1 mg/h. The animals were ventilated at approximately 18 strokes/min, with a continuous flow of 20 O 2+80 N 2 O using a variable volume respirator (Siemens™ volumeter). The head was secured to avoid movements. Systolic and diastolic blood pressures were monitored via the femoral artery using a transducer (Mingograph™, Siemens-Elema, Switzerland). The systemic arterial pO 2, pCO 2, and pH (pO 2,a, pCO 2,a, and pH a, respectively) were measured intermittently from the same artery with a blood gas analyzer (AVL, automatic gas system 940). Adjustments of ventilatory rate, stroke volume, and the composition of the inspired gas maintained values of pCO 2,a of 30 to 35 mm Hg and pO 2,a of 100 to 110 mm Hg. A rectal thermometer was used to measure the animal’s temperature, which was maintained between 36 and 38°C by means of a thermoblanket with temperature control unit.

2.6.

Protocols

All PQ and LDF measurements were performed at a rim location of the optic disk and care was taken to avoid as well as possible the large retinal vessels visible at the disk surface.

In the 11 animals, pO 2,blood measurements were collected intermittently during 7:25±1:45 (standard deviation) hours after dye injection at an average of rate of ∼5 min−1. Between PQ measurements that were performed in total darkness, the diffuse white fundus illumination (2 mW/cm2) was occasionally turned on for alignment.

The PQ and LDF techniques were employed alternatively to record the pO 2,blood and F ONH response to three different ventilation mixtures, each of which was tested on one animal. After a recording of the baseline pO 2,blood and F ONH with the animal ventilated with air (20 O 2+80 N 2 O) for approximately 1 min, the breathing conditions were changed. One animal was given 100 N 2 O to breathe until death. The second animal received 100 O 2 for 10 min and the third one breathed a CO 2 enriched air mixture (20 O 2+5 CO 2+75 N 2 O) for 12 min. The pO 2,blood and F ONH were recorded alternatively.

3.

Results

Figure 2(a) shows a typical time course of the logarithm of the phosphorescence intensity and associated linear fit (regression coefficient=0.975, p<0.001) to the 810 data points above the dark noise level. The lifetime τ of the phosphorescence was found to be 87±3 μs, a value which, according to the Stern–Volmer equation, corresponds to a pO 2,blood of 26±1 mm Hg. A typical recording of pO 2,blood values obtained with the phosphorometer at a 2.5-s sampling rate is shown in Figure 2(b).

Figure 2

(a) Typical signal obtained with the phosphorometer by averaging 20 phosphorescence decays; 468 data points above dark noise level were used for analysis. The solid line represents the weighted linear fit of the data logarithm (R=0.975,p<0.001), with τ=87±3 μs. The pO 2,blood calculated by using the Stern–Volmer equation with the appropriate τ0 and kQ constant is 26±1 mm Hg. (b) Typical microvascular pO 2,blood recording in the ONH achieved a rate of one per 2.5 s.

005301j.2.jpg

3.1.

pO 2,blood Time Course

All pO 2,blood values collected in the 11 animals were grouped according to the time of their measurement after the phosphor injection (30-min groups). For each group, the mean pO 2,blood (± standard deviation) was calculated and plotted versus time (Figure 3). A significant decrease of pO 2,blood (ANOVA, p<0.0001) was observed, although during the first 120 min of the PQ measurements (period I), the pO 2,blood did not change (ANOVA, LSD post hoc test, p>0.05) and had an average value of 24±1.4 mm Hg. Thereafter (period II), it decreased steadily to a lower level of 7±0.9 mm Hg after 400 min (ANOVA, LSD post hoc test, p<0.05) . In most animals, observations of the eye fundus at this time of the experiment revealed constrictions and even closure of some arteries and veins in the area that had been repeatedly excited for PQ measurements. The retina and ONH tissues were edematous. A breakdown of the blood-retinal barrier was revealed by the leakage of fluorescein observed on angiograms (not shown). In one animal, the contralateral eye, which remained in complete darkness during the 375 min of the PQ investigations on the fellow eye, was prepared for PQ measurements. It appeared entirely normal and pO 2,blood averaged over different sites of the disk was 19±2.2 mm Hg. Fluorescein angiography performed on this eye did not show any early or late leakage of the dye.

Figure 3

All the pO 2,blood values measured in 11 animals were grouped according to their delay after the phosphor injection (30-min groups). The average value ± standard deviation is plotted here as a function of time. The pO 2,blood did not vary significantly [analysis of variance (ANOVA), p>0.05] during the first 100 min (period I) but then (period II) dropped significantly (ANOVA; p<0.001) .

005301j.3.jpg

3.2.

Measurements of pO 2,blood and F ONH

In the animal that was ventilated with 100 N 2 O. The pO 2,a dropped continuously, which produced a F ONH increase of 75 and a pO 2,blood decrease of 60 (Figure 4).

Figure 4

In one animal, pO 2,blood (dots) and F ONH (line) were simultaneously measured at the peripheral ONH during an acute hypoxemia. At time 0, a ventilation with 0 O 2 was used. After 2 min, pO 2,blood was decreased by 60 while F ONH increased by 75.

005301j.4.jpg

In the animal ventilated for 10 min with 100 O 2, blood gases taken during normoxia and after 5 min of hyperoxia showed an increase of pO 2,a from 109 to 356 mm Hg. Consecutively, a 25 pO 2,blood increase and a 23 reduction of F ONH were measured (Figure 5). Both pO 2,blood and F ONH returned to baseline levels after the breathing was back to normal.

Figure 5

Recording of pO 2,blood (dots) and F ONH (line) at the peripheral ONH during hyperoxia (one animal). After 90 s of baseline, the ventilation was switched to 100 O 2. The F ONH decreased by 22.8 while pO 2,blood increased by 25. The ventilation was switched to normal after 700 s and both parameters went back to their baseline levels.

005301j.5.jpg

The animal ventilated with 20 O 2+5 CO 2+75 N 2 O became hypercapnic since the pCO 2,a increased from 34.3 to 45.6 mm Hg. Simultaneously, pO 2,blood and F ONH increased by 24 and 12, respectively (Figure 6).

Figure 6

Recording of pO 2,blood (dots) and F ONH (line) at the peripheral ONH during hypercapnia (one animal). After 2 min of baseline, the ventilation was switched to 20 O 2+5 CO 2+75 N 2. The pO 2,blood increased by 12 and F ONH by 25.

005301j.6.jpg

4.

Discussion

This paper shows that measurements of pO 2,blood and F ONH can be obtained quasisimultaneously at the ONH disk with a fundus camera modified for PQ/LDF. The pO 2,blood and F ONH measurements were achieved with a time resolution of 2.5-s (Figure 2) and 47 ms,31 respectively.

4.1.

Range of pO 2,blood

Several phosphorometers have been developed for studying pO 2 in various biological tissues (muscles, brain, intestine, kidney, tumors).21 32 33 34 35 36 37 However, the device presented here is the first that enables focal measurements of pO 2,blood in the ocular fundus. As in the brain, the endothelial cells covering the inner surface of the retinal and ONH vessels in the healthy eye present tight junctions, which prevent the diffusion of soluble molecules such as albumin from the blood into the interstitial tissues (blood-retinal barrier). Therefore, our technique of phosphorescence quenching by O 2 assesses only the phosphorescence emitted by the dye inside the vessels. Values of pO 2,blood in the ocular fundus were already measured in pigs and cats with an imaging technique using the phosphorescence quenching principle.22 24 These studies reported ONH pO 2,blood values of 35±9 and 33 mm Hg, respectively, which were higher than those obtained at period I with our phosphorometer in 11 miniature pigs (24±1.4 mm Hg). Even if these two imaging techniques and our phosphorometer applied the same phosphorescence quenching principle, specificity of each instrumental setup or data processing could account for the different pO 2,blood range. As pointed out by Golub et al.;,38 39 the temporal characteristic of the flash impulse function influences the shape of the phosphorescence emission (convolution) and consequently its lifetime. Because the excitation light delivered by our instrumental setup reaches its peak intensity after 2 μs, whereas the flash used in these two studies peaked at about 10 μs, the phosphorescence decay emitted by a given tissue sample certainly differs when obtained with different instruments. Furthermore, with the ideal impulse function for time-resolved phosphorescence measurements being a Dirac function, we believe that our pO 2,blood measurements obtained with a shorter flash impulse are more accurate than those obtained with flashes of longer duration. Difference in the processing of the phosphorescence emission may also explain the differences between the pO 2,blood values obtained with various devices. Whereas Blumenro¨der et al.; used a Levenberg–Marquardt fitting routine for calculating the lifetime of the phosphorescence decay22 and Shonat et al.; a linear fitting routine,24 we applied a fast, weighted linear fit optimized for noise correction to extract30 τ. Another important element that has to be considered when comparing the various pO 2,blood values is that these were obtained in different species (minipigs, pigs and cats).

As the diffusion of O 2 from the blood vessels into the interstitial ONH tissue depends on (1) the intravascular blood O 2 content, (2) the hemodynamics, (3) the O 2 consumption of the vessel wall, (4) the architecture of the ONH microvascular network, and (5) the metabolic tissue consumption in the ONH, to establish the relation between pO 2,tissue and intravascular pO 2,blood a diffusion model is required. For the optic nerve, with its complex multilayered vascular systems, such a model is lacking. The pO 2,blood values obtained at the ONH in our study are not at odds with pO 2,tissue values obtained by others that used O 2 -sensitive microelectrodes. Thus, Bouzas et al.;12 found a prepapillary pO 2 of 33.1±3.9 mm Hg at juxta-arteriolar locations and 16.6±1.4 mm Hg at intervascular locations, both with the probe tip positioned at 50 μm in front of the ONH surface. Stefansson et al.;,17 using large microelectrodes with a 100-μm tip, reported an intervascular prepapillary pO 2 of 24.1±11.6 mm Hg. At 50 and 200 μm under the surface of the ONH cup, pO 2,tissue was 9.6±1.2 and 9.4±0.8 mm Hg, respectively.12 At similar depths in the ONH rim, pO 2,tissue was12 10.8±1.9 and 8.2±1.8 mm Hg. In comparison, pO 2,tissue measured between 0- and 200-μm depth into the optic disk of the cat range11 from 15 to 10 mm Hg. These values measured in the depths of the nerve, compared to the higher values found at the surface of the disc, suggest that our phosphorometer predominantly measures pO 2,blood in the retinal layer of the nerve.

4.2.

Origin of the Phosphorescence

Due to the topography of the ONH, the depth of tissue sampled with the PQ technique is difficult to evaluate. The green excitation light used for PQ measurements, because of its absorption and scattering by the ONH tissue and blood, should predominantly interact with the dye molecules located in the vessels close to the ONH surface. However, measurements by Koelle et al.; using sections of the cat ONH suggest that green light backscattered at a depth of 400 μm from the ONH surface still contributes to the signal detection.40 In their paper on PQ in the pig, Blumenro¨der et al.;22 concluded from the fact that the lamina cribrosa cannot be seen by green light ophthalmoscopy that the light does not reach this layer. This argument, however, does not take into consideration that most of the light penetrating the ONH tissue is scattered, a process that may prevent the observation of a clear image of the lamina cribrosa. Regarding the vascular origin of the phosphorescence, placing the pin hole away from the larger vessels at the surface favored detection of phosphorescence from dye circulating in the small arterioles, capillaries, and venules, vessels that are usually not visible with a fundus camera. Therefore, the detected phosphorescence is a superposition of contributions from arterioles, capillaries, and venules. It is expected that, since the venous blood volume exceeds the arterial and capillary blood volumes, the major contribution of the measured pO 2,blood comes from venules. This contribution is further emphasized by the fact that quenching is weaker for the vessels with low pO 2,blood than for those with high pO 2,blood .

4.3.

Time Course of pO 2,blood during Prolonged PQ Measurements

The decrease of pO 2,blood during the period II of the PQ measurements is probably due to a decreased blood flow, as suggested by the marked constriction of the vasculature and white appearance of the ONH tissue in the region illuminated by the excitation light that was observed toward the end of this period. As the maximum irradiance used for each pO 2,blood measurements (30 flashes×25 μJ/cm 2=750 μJ/cm 2 at 530 nm) is below the maximum permissible ocular exposure41 for a continuous exposure of 1 s (30 flashes delivered at 30 Hz), the damage observed at the ocular fundus is not the consequence of a thermic process, but likely results from the photoactivation of our phosphor. Light activation of intravascularly administered fluorochromes is known to generate toxic oxygen species, including singlet oxygen and superoxide anions. Various deleterious microvascular effects such as platelet thrombus formation, arteriolar constriction, or postcapillary venular leakage have been reported and attributed to the luminal generation of oxidant species.42 These effects provide the basis of the photodynamic treatment of tumors.43 As the quenching of an activated phosphor by an O 2 molecule also produces a singlet oxygen molecule, extended pO 2,blood measurements with the phosphorescence quenching technique may lead to vascular damage.

In one animal, the contralateral eye was also investigated at period II (direct observation, pO 2,blood measurements, fluorescein angiography). In this eye, which remained in darkness for hours before, the ocular fundus looked normal and pO 2,blood was approximately 19 mm Hg. It shows that the porphyrin dye, if not optically excited, does not affect the pO 2,blood level, which supports the preceeding hypothesis on alterations due to photoactivation of the phosphorescent probe.

In our system, two sources of phosphorescence excitation susceptible to induce photoactivation were present. One was the flash for excitation of the dye, the other the white light used for fundus observation. According to the data of Shonat et al.;,33 35 the energy density delivered to obtain a pO 2,blood measurement (∼750 mJ/cm2) should not affect the tissue. We presume that it is the visible light used to occasionally observe the fundus that caused the most severe photoactivation and produced tissue alterations resulting in the drop of the pO 2,blood in period II.

Future studies will benefit from any attempt to reduce the likelihood of alterations due to photoactivation, such as improving the phosphorescence detection sensitivity and reducing the number of flashes needed for the average phosphorescence decay curve, decreasing the dye concentration and removing the spectral component of the observation light likely to induce photoactivation.

4.4.

Modification of the Ventilation Conditions

Previous PQ or LDF measurements at the ONH during hypoxemia reported that pO 2,blood drops,22 and F ONH increases44 by 67 to 89. Our measurement, albeit in only one animal, is in agreement with these results. It also illustrates the feasibility of quasisimultaneous measurements at the ONH rim of pO 2,blood and F ONH , since 100 N 2 O breathing induced a 60 decrease and a 75 increase of these parameters, respectively. This measurement illustrates the regulation capabilities of the ONH vasculature, which compensates the drop of the arterial blood O 2 content by increasing blood flow.

LDF measurements performed during hyperoxia (100 O 2 breathing) showed blood flow decreases of 39 to 49 in cats44 and 25 to 38 in humans.45 46 Using the PQ imaging technique, pO 2,blood was also measured at the optic disc during hyperoxia in 6 piglets.22 In that study, pO 2,blood rose from 35±9 to 44±14 mm Hg, while pO 2,a was increased from 100 to 120 mm Hg with an O 2 -rich ventilation. Our measurement during hyperoxia (100 O 2 breathing) shows that pO 2,blood increased by 25, whereas pO 2,a raised by 226. The difference between the increase of pO 2,a and pO 2,blood indicates a regulation mechanism that prevents excessive amounts of O 2 to reach the ocular microcirculation.

CO 2 has a vasodilating effect on the blood vessels, which can enhance the blood perfusion. Our F ONH measurement is in accordance with a study by Harris,46 who reported a similar 28 increase of the ONH blood flow when using the same gas mixture. A higher perfusion apparently contributes to increase the supply of O 2 to the capillaries since pO 2,blood increased by approximately 12.

The good agreement of the results obtained during hypoxemia, hyperoxia, and hypercapnia with the literature shows that pO 2,blood and blood flow measurements can be reliably measured quasisimultaneously in the microcirculation of the prelaminar optic nerve. This instrument is specifically adapted for investigating with a good time resolution the time course of transient pO 2,blood changes that could occur during various physiological perturbations.

Acknowledgments

The authors wish to thank Jean-Luc Munoz for the preparation of the animals. This work was supported in part by Grant No. 3100-057045 and 3200-61685 from the Swiss National Foundation for Scientific Research and a grant from the Loterie Romande.

REFERENCES

1. 

P. A. D’Amore , “Mechanisms of retinal and choroidal neovascularization,” Invest. Ophthalmol. Visual Sci. , 35 (12), 3974 –3979 (1994). Google Scholar

2. 

C. J. Pournaras , “Retinal oxygen distribution. Its role in the physiopathology of vasoproliferative microangiopathies,” Retina , 15 (4), 332 –347 (1995). Google Scholar

3. 

E. Stefansson , M. B. Landers III , and M. L. Wolbarsht , “Oxygenation and vasodilatation in relation to diabetic and other proliferative retinopathies,” Ophthalmic Surg. , 14 (3), 209 –226 (1983). Google Scholar

4. 

G. N. Wise , “Retinal neovascularization,” Trans. Am. Ophthalmol. Soc. , 54 (729), 826 –835 (1956). Google Scholar

5. 

S. S. Hayreh , “The pathogenesis of optic nerve lesions in glaucoma,” Trans. Am. Acad. Ophthalmol. Otolaryngol. , 81 (2), 197 –213 (1976). Google Scholar

6. 

V. A. Alder and S. J. Cringle , “Vitreal and retinal oxygenation,” Graefe's Arch. Clin. Exp. Ophthalmol. , 228 151 –157 (1990). Google Scholar

7. 

R. A. Linsenmeier , “Effects of light and darkness on oxygen distribution and consumption in the cat retina,” J. Gen. Physiol. , 88 521 –542 (1986). Google Scholar

8. 

C. E. Riva , C. J. Pournaras , and M. Tsacopoulos , “Regulation of local oxygen tension and blood flow in the inner retina during hyperoxia,” J. Appl. Physiol. , 61 (2), 592 –598 (1986). Google Scholar

9. 

J. Ahmed , R. D. Braun , R. Dunn Jr. , and R. A. Linsenmeier , “Oxygen distribution in the macaque retina,” Invest. Ophthalmol. Visual Sci. , 34 (3), 516 –521 (1993). Google Scholar

10. 

S. J. Cringle, D. Y. Yu, V. A. Alder, and E. N. Su, “Oxygen tension and blood flow in the retina of normal and diabetic rats,” in Oxygen Transport to Tissue XIV, W. Erdmaann and D. F. Bruley, Eds., pp. 787–791, Plenum, New York (1992).

11. 

J. Ahmed , R. A. Linsenmeier , and R. Dunn , “The oxygen distribution in the prelaminar optic nerve head of the cat,” Exp. Eye Res. , 59 457 –466 (1994). Google Scholar

12. 

E. A. Bouzas , G. Donati , and C. J. Pournaras , “Distribution and regulation of the optic nerve head tissue PO2, ,” Surv. Ophthalmol. , 42 (2), 1 –8 (1997). Google Scholar

13. 

D. G. Buerk, D. N. Atochin, and C. E. Riva, “Simultaneous tissue pO2, nitric oxide, and laser Doppler blood flow measurements during neuronal activation of optic nerve head,” in Oxygen Transport to Tissue XX, A. Hudetz and D. F. Bruley, Eds., pp. 159–164, Plenum, New York (1998).

14. 

J. T. Ernest , “Autoregulation of optic-disc oxygen tension,” Invest. Ophthalmol. , 13 (2), 101 –106 (1974). Google Scholar

15. 

J. T. Ernest , “Optic disc oxygen tension,” Exp. Eye Res. , 24 271 –278 (1977). Google Scholar

16. 

C. E. Riva , C. J. Pournaras , C. L. Poitry-Yamate , and B. L. Petrig , “Rhythmic changes in velocity, volume, and flow of blood in the optic nerve head tissue,” Microvasc. Res. , 40 (1), 36 –45 (1990). Google Scholar

17. 

E. Stefansson , P. K. Jensen , T. Eysteinsson , K. Bang , J. F. Kiilgaard , J. Dollerup , E. Scherfig , and M. la Cour , “Optic nerve oxygen tension in pigs and the effect of carbonic anhydrase inhibitors,” Invest. Ophthalmol. Visual Sci. , 40 2756 –2762 (1999). Google Scholar

18. 

H. Sakaue , Y. Tsukahara , A. Negi , N. Ogino , and Y. Honda , “Measurement of vitreous oxygen tension in human eyes,” Jpn. J. Ophthalmol. , 33 (2), 199 –203 (1989). Google Scholar

19. 

H. Sakaue , A. Negi , and Y. Honda , “Comparative study of vitreous oxygen tension in human and rabbit eyes,” Invest. Ophthalmol. Visual Sci. , 30 (9), 1933 –1937 (1989). Google Scholar

20. 

E. Stefansson , R. Machemer , E. de Juan Jr. , B. W. McCuen , and J. Peterson , “Retinal oxygenation and laser treatment in patients with diabetic retinopathy,” Am. J. Ophthalmol. , 113 (1), 36 –38 (1992). Google Scholar

21. 

J. M. Vanderkooi , G. Maniara , T. J. Green , and D. F. Wilson , “An optical method for measurement of dioxygen concentration based upon quenching of phosphorescence,” J. Biol. Chem. , 262 (12), 5476 –5482 (1987). Google Scholar

22. 

S. Blumenro¨der , A. J. Augustin , and F. H. J. Koch , “The influence of intraocular pressure and systemic oxygen tension on the intravascular pO2 of the pig retina as measured with phosphorescence imaging,” Surv. Ophthalmol. , 42 (1), S118 –S126 (1997). Google Scholar

23. 

R. D. Shonat , D. F. Wilson , C. E. Riva , and S. D. Cranstoun , “Effect of acute increases in intraocular pressure on intravascular optic nerve oxygen tension in cats,” Invest. Ophthalmol. Visual Sci. , 33 (11), 3174 –3180 (1992). Google Scholar

24. 

R. D. Shonat , D. F. Wilson , C. E. Riva , and M. Pawlowski , “Oxygen distribution in the retinal and choroidal vessels of the cat as measured by a new phosphorescence imaging method,” Appl. Opt. , 31 (19), 3711 –3718 (1992). Google Scholar

25. 

D. Malonek , U. Dirnagl , U. Lindauer , K. Yamada , I. Kanno , and A. Grinvald , “Vascular imprints of neuronal activity: Relationships between the dynamics of cortical blood flow, oxygenation, and volume changes following sensory stimulation,” Proc. Natl. Acad. Sci. U.S.A. , 94 14826 –14831 (1997). Google Scholar

26. 

D. G. Buerk , C. E. Riva , and S. D. Cranstoun , “Frequency and luminance-dependent blood flow and K+ ion changes during flicker stimuli in cat optic nerve head,” Invest. Ophthalmol. Visual Sci. , 36 2216 –2227 (1995). Google Scholar

27. 

C. E. Riva and D. G. Buerk , “Dynamic coupling of blood flow to function and metabolism in the optic nerve head,” J. Neuroophthalmol. , 20 (2), 45 –54 (1998). Google Scholar

28. 

D. F. Wilson , A. Pastuszko , J. E. DiGiacomo , M. Pawlowski , R. Schneiderman , and M. Delivoria-Papadopoulos , “Effect of hyperventilation on oxygenation of the brain cortex of newborn piglets,” J. Appl. Physiol. , 70 (6), 2691 –2696 (1991). Google Scholar

29. 

L.-W. Lo , C. J. Koch , and D. F. Wilson , “Calibration of oxygen-dependent quenching of the phosphorescence of Pd-meso-tetra (4-carboxyphenyl) porphine: a phosphor with general application for measuring oxygen concentration in biological systems,” Anal. Biochem. , 236 153 –160 (1996). Google Scholar

30. 

W. H. Press, B. P. Flannery, S. A. Teukolsky, and W. T. Vetterling, in Numerical Recipes in C: The Art of Scientific Computing, pp. 523–528, Cambridge University Press, Cambridge (1988).

31. 

B. L. Petrig and C. E. Riva, “Optic nerve head laser Doppler flowmetry: principles and computer analysis,” in Ocular Blood Flow, H. J. Kaiser, J. Flammer, and P. Hendrickson, Eds., pp. 120–127, Karger, Basel (1996).

32. 

T. J. Green , D. F. Wilson , J. M. Vanderkooi , and S. P. DeFeo , “Phosphorimeters for analysis of decay profiles and real time monitoring of exponential decay and oxygen concentrations,” Anal. Biochem. , 174 (1), 73 –79 (1988). Google Scholar

33. 

R. D. Shonat , K. N. Richmond , and P. C. Johnson , “Phosphorescence quenching and the microcirculation: an automated, multipoint oxygen tension measuring instrument,” Rev. Sci. Instrum. , 66 (10), 5075 –5084 (1995). Google Scholar

34. 

L. Zheng , A. S. Golub , and R. N. Pittman , “Determination of pO2 and its heterogeneity in single capillaries,” Am. J. Physiol. , 271 (40), H365 –H372 (1996). Google Scholar

35. 

R. D. Shonat , E. S. Wachman , W. Niu , A. P. Koretsky , and D. L. Farkas , “Near-simultaneous hemoglobin saturation and oxygen tension maps in mouse brain using AOTF microscope,” Biophys. J. , 73 1223 –1231 (1997). Google Scholar

36. 

F. Torres , I. and M. Intaglietta , “Microvessel PO2 measurements by phosphorescence decay method,” Am. J. Physiol. , 265 (4), H1434 –H1438 (1993). Google Scholar

37. 

M. Sinaasappel , C. Donkersloot , J. van Bommel , and C. Ince , “ PO2 measurements in the rat intestinal microcirculation,” Am. J. Physiol. , 276 (6), G1515 –G1520 (1999). Google Scholar

38. 

A. S. Golub , A. S. Popel , L. Zheng , and R. N. Pittman , “Analysis of phosphorescence decay for nonuniform oxygen tension. Consequences of finite excitation flash duration,” Adv. Exp. Med. Biol. , 471 649 –659 (1999). Google Scholar

39. 

A. S. Golub , A. S. Popel , L. Zheng , and R. N. Pittman , “Analysis of phosphorescence decay in heterogeneous systems: consequences of finite excitation flash duration,” Photochem. Photobiol. , 69 (6), 624 –632 (1999). Google Scholar

40. 

J. S. Koelle , C. E. Riva , B. L. Petrig , and S. D. Cranstoun , “Depth of tissue sampling in the optic nerve head using laser Doppler flowmetry,” Laser Med. Sci. , 8 49 –54 (1993). Google Scholar

41. 

American National Standard for Safe Use of Lasers, p. 85, Laser Institute of America, Orlando (2000).

42. 

V. H. Fingar , “Vascular effects of photodynamic therapy,” J. Clin. Laser Med. Surg. , 14 (5), 323 –328 (1996). Google Scholar

43. 

T. J. Wieman and V. H. Fingar , “Photodynamic therapy,” Surg. Clin. North Am. , 72 (3), 609 –622 (1992). Google Scholar

44. 

C. E. Riva , S. Harino , B. L. Petrig , and R. D. Shonat , “Laser Doppler flowmetry in the optic nerve,” Exp. Eye Res. , 55 499 –506 (1992). Google Scholar

45. 

A. Lietz , P. Hendrickson , J. Flammer , S. Orgu¨l , and I. O. Haefliger , “Effect of carbogen, oxygen and intraocular pressure on Heidelberg retina flowmeter parameter “flow” measured at the papilla,” Ophthalmologica , 212 149 –152 (1998). Google Scholar

46. 

A. Harris , “Laser Doppler flowmetry measurement of changes in human optic nerve head blood flow in response to blood gas perturbations,” J. Glaucoma , 5 258 –265 (1996). Google Scholar
©(2003) Society of Photo-Optical Instrumentation Engineers (SPIE)
Stephane R. Chamot, Stephen D. Cranstoun, Benno L. Petrig, Constantin J. Pournaras, and Charles E. Riva "Blood pO2 and blood flow at the optic disc," Journal of Biomedical Optics 8(1), (1 January 2003). https://doi.org/10.1117/1.1527935
Published: 1 January 2003
Lens.org Logo
CITATIONS
Cited by 13 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Phosphorescence

Blood

Mercury

Tissues

Blood circulation

Eye

Molecules

Back to Top