Skip to main content
Log in

Retinal oxygen distribution and the role of neuroglobin

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

The retina is the tissue layer at the back of the eye that is responsible for light detection. Whilst equipped with a rich supply of oxygen, it has one of the highest oxygen demands of any tissue in the body and, as such, supply and demand are finely balanced. It has been suggested that the protein neuroglobin (Ngb), which is found in high concentrations within the retina, may help to maintain an adequate supply of oxygen via the processes of transport and storage. We construct mathematical models, formulated as systems of reaction–diffusion equations in one-dimension, to test this hypothesis. Numerical simulations show that Ngb may play an important role in oxygen transport, but not in storage. Our models predict that the retina is most susceptible to hypoxia in the regions of the photoreceptor inner segment and inner plexiform layers, where Ngb has the potential to prevent hypoxia and increase oxygen uptake by 30–40 %. Analysis of a simplified model confirms the utility of Ngb in transport and shows that its oxygen affinity (\(P_{50}\) value) is near optimal for this process. Lastly, asymptotic analysis enables us to identify conditions under which the piecewise linear and quadratic approximations to the retinal oxygen profile, used in the literature, are valid.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

References

  • Alder VA, Cringle SJ, Constable IJ (1983) The retinal oxygen profile in cats. Invest Ophthalmol Vis Sci 24(1):30-36

    Google Scholar 

  • Anderson B (1968) Ocular effects of changes in oxygen and carbon dioxide tension. Trans Am Ophthalmol Soc 66:423-474

    Google Scholar 

  • Anderson B, Saltzman HA (1964) Retinal oxygen utilization measured by hyperbaric blackout. Arch Ophthalmol 72(6):792-795

    Article  Google Scholar 

  • Bender CM, Orszag SA (1999) Advanced mathematical methods for scientists and engineers I: asymptotic methods and perturbation theory. Springer, Berlin

    Book  MATH  Google Scholar 

  • Bentmann A, Schmidt M, Reuss S, Wolfrum U, Hankeln T, Burmester T (2005) Divergent distribution in vascular and avascular mammalian retinae links neuroglobin to cellular respiration. J Biol Chem 280(21):20660-20665

    Article  Google Scholar 

  • Birol G, Wang S, Budzynski E, Wangsa-Wirawan ND, Linsenmeier RA (2007) Oxygen distribution and consumption in the macaque retina. Am J Physiol Heart Circ Physiol 293(3):H1696-H1704

    Article  Google Scholar 

  • Braun RD, Linsenmeier RA, Goldstick TK (1995) Oxygen consumption in the inner and outer retina of the cat. Invest Ophthalmol Vis Sci 36(3):542-554

    Google Scholar 

  • Burmester T, Hankeln T (2004) Neuroglobin: a respiratory protein of the nervous system. News Physiol Sci 19(3):110-113

    Google Scholar 

  • Burmester T, Hankeln T (2009) What is the function of neuroglobin? J Exp Biol 212(10):1423-1428

    Article  Google Scholar 

  • Burmester T, Weich B, Reinhardt S, Hankeln T (2000) A vertebrate globin expressed in the brain. Nature 407(6803):520-523

    Article  Google Scholar 

  • Burmester T, Ebner B, Weich B, Hankeln T (2002) Cytoglobin: a novel globin type ubiquitously expressed in vertebrate tissues. Mol Biol Evol 19(4):416-421

    Article  Google Scholar 

  • Chan G, Balaratnasingam C, Yu PK, Morgan WH, McAllister IL, Cringle SJ, Yu DY (2012) Quantitative morphometry of perifoveal capillary networks in the human retina. Invest Ophthalmol Vis Sci 53(9):5502-5514

    Article  Google Scholar 

  • Costa LE, Mendez G, Boveris A (1997) Oxygen dependence of mitochondrial function measured by high-resolution respirometry in long-term hypoxic rats. Am J Physiol 273(3):C852-C858

    Google Scholar 

  • Cringle SJ, Yu DY (2002) A multi-layer model of retinal oxygen supply and consumption helps explain the muted rise in inner retinal PO(2) during systemic hyperoxia. Comp Biochem Physiol 132(1):61-66

    Article  Google Scholar 

  • Dollery CT, Bulpitt CJ, Kohner EM (1969) Oxygen supply to the retina from the retinal and choroidal circulations at normal and increased arterial oxygen tensions. Invest Ophthalmol Vis Sci 8(6):588-594

    Google Scholar 

  • Fago A, Hundahl C, Malte H, Weber RE (2004b) Functional properties of neuroglobin and cytoglobin. insights into the ancestral physiological roles of globins. IUBMB Life 56(11-12):689-696

    Article  Google Scholar 

  • Gillies MC, Su T, Naidoo D (1995) Electrical resistance and macromolecular permeability of retinal capillary endothelial cells in vitro. Curr Eye Res 14(6):435-442

    Article  Google Scholar 

  • Goldman D (2008) Theoretical models of microvascular oxygen transport to tissue. Microcirculation 15(8):795-811

    Article  Google Scholar 

  • Hamdane D, Kiger L, Dewilde S, Green BN, Pesce A, Uzan J, Burmester T, Hankeln T, Bolognesi M, Moens L, Marden MC (2003) The redox state of the cell regulates the ligand binding affinity of human neuroglobin and cytoglobin. J Biol Chem 278(51):51713-51721

    Article  Google Scholar 

  • Hardarson SH, Stefánsson E (2010) Oxygen saturation in central retinal vein occlusion. Am J Ophthalmol 150(6):871-875

    Article  Google Scholar 

  • Hardarson SH, Basit S, Jonsdottir TE, Eysteinsson T, Halldorsson GH, Karlsson RA, Beach JM, Benediktsson JA, Stefansson E (2009) Oxygen saturation in human retinal vessels is higher in dark than in light. Invest Ophthalmol Vis Sci 50(5):2308-2311

    Article  Google Scholar 

  • Haugh L, Linsenmeier R, Goldstick T (1990) Mathematical models of the spatial distribution of retinal oxygen tension and consumption, including changes upon illumination. Ann Biomed Eng 18:19-36

    Article  Google Scholar 

  • Jürgens KD, Peters T, Gros G (1994) Diffusivity of myoglobin in intact skeletal muscle cells. Proc Natl Acad Sci 91(9):3829-3833

    Article  Google Scholar 

  • Keener J, Sneyd J (1998) Mathematical physiology. Springer, Berlin

    MATH  Google Scholar 

  • Kiger L, Uzan J, Dewilde S, Burmester T, Hankeln T, Moens L, Hamdane D, Baudin-Creuza V, Marden MC (2004) Neuroglobin ligand binding kinetics. IUBMB Life 56(11-12):709-719

    Article  Google Scholar 

  • Kohen R, Nyska A (2002) Invited review: oxidation of biological systems: oxidative stress phenomena, antioxidants, redox reactions, and methods for their quantification. Toxicol Pathol 30(6):620-650

    Article  Google Scholar 

  • Kur J, Newman EA, Chan-Ling T (2012) Cellular and physiological mechanisms underlying blood flow regulation in the retina and choroid in health and disease. Prog Retin Eye Res 31(5):377-406

    Article  Google Scholar 

  • Linsenmeier RA (1986) Effects of light and darkness on oxygen distribution and consumption in the cat retina. J Gen Physiol 88(4):521-542

    Article  Google Scholar 

  • Linsenmeier RA, Braun RD (1992) Oxygen distribution and consumption in the cat retina during normoxia and hypoxemia. J Gen Physiol 99(2):177-197

    Article  Google Scholar 

  • Linsenmeier RA, Padnick-Silver L (2000) Metabolic dependence of photoreceptors on the choroid in the normal and detached retina. Invest Ophthalmol Vis Sci 41(10):3117-3123

    Google Scholar 

  • Linsenmeier RA, Yancey CM (1989) Effects of hyperoxia on the oxygen distribution in the intact cat retina. Invest Ophthalmol Vis Sci 30(4):612-618

    Google Scholar 

  • McGuire BJ, Secomb TW (2001) A theoretical model for oxygen transport in skeletal muscle under conditions of high oxygen demand. J Appl Physiol 91(5):2255-2265

    Google Scholar 

  • Ockendon J, Howison S, Lacey A, Movchan A (2003) Applied partial differential equations, revised edn. Oxford University Press, Oxford

    MATH  Google Scholar 

  • Ostojić J, Sakaguchi DS, de Lathouder Y, Hargrove MS, Trent JT, Kwon YH, Kardon RH, Kuehn MH, Betts DM, Grozdanić S (2006) Neuroglobin and cytoglobin: oxygen-binding proteins in retinal neurons. Invest Ophthalmol Vis Sci 47(3):1016-1023

    Article  Google Scholar 

  • Ostojić J, Grozdanić SD, Syed NA, Hargrove MS, Trent JT, Kuehn MH, Kwon YH, Kardon RH, Sakaguchi DS (2008) Patterns of distribution of oxygen-binding globins, neuroglobin and cytoglobin in human retina. Arch Ophthalmol 126(11):1530-1536

    Article  Google Scholar 

  • Oyster CW (1999) The human eye: structure and function. Sinauer Associates Inc, Sunderland

    Google Scholar 

  • Padnick-Silver L, Linsenmeier RA (2003) Effect of acute hyperglycemia on oxygen and oxidative metabolism in the intact cat retina. Invest Ophthalmol Vis Sci 44(2):745-750

    Article  Google Scholar 

  • Pournaras CJ, Rungger-Brändle E, Riva CE, Hardarson SH, Stefansson E (2008) Regulation of retinal blood flow in health and disease. Prog Retin Eye Res 27(3):284-330

    Article  Google Scholar 

  • Rajendram R, Rao NA (2007) Neuroglobin in normal retina and retina from eyes with advanced glaucoma. Br J Ophthalmol 91(5):663-666

    Article  Google Scholar 

  • Richmond KN, Shonat RD, Lynch RM, Johnson PC (1999) Critical PO(2) of skeletal muscle in vivo. Am J Physiol Heart Circ Physiol 277(5):H1831-H1840

    Google Scholar 

  • Roh HD, Goldstick TK, Linsenmeier RA (1990) Spatial variation of the local tissue oxygen diffusion coefficient measured in situ in the cat retina and cornea. Adv Exp Med Biol 277:127-136

    Article  Google Scholar 

  • Schmidt M, Giessl A, Laufs T, Hankeln T, Wolfrum U, Burmester T (2003) How does the eye breathe? J Biol Chem 278(3):1932-1935

    Article  Google Scholar 

  • Schmidt M, Laufs T, Reuss S, Hankeln T, Burmester T (2005) Divergent distribution of cytoglobin and neuroglobin in the murine eye. Neurosci Lett 374(3):207-211

    Article  Google Scholar 

  • Stefánsson E (1988) Retinal oxygen tension is higher in light than dark. Pediatr Res 23:5-8

    Article  Google Scholar 

  • Swaroop A, Kim D, Forrest D (2010) Transcriptional regulation of photoreceptor development and homeostasis in the mammalian retina. Nat Rev Neurosci 11:563-576

    Article  Google Scholar 

  • Tan PEZ, Yu PK, Balaratnasingam C, Cringle SJ, Morgan WH, McAllister IL, Yu DY (2012) Quantitative confocal imaging of the retinal microvasculature in the human retina. Invest Ophthalmol Vis Sci 53(9):5728-5736

    Article  Google Scholar 

  • Törnquist P, Alm A, Bill A (1990) Permeability of ocular vessels and transport across the blood-retinal-barrier. Eye 4(2):303-309

    Article  Google Scholar 

  • Trent JT, Hargrove MS (2002) A ubiquitously expressed human hexacoordinate hemoglobin. J Biol Chem 277(22):19538-19545

    Article  Google Scholar 

  • Wangsa-Wirawan ND, Linsenmeier RA (2003) Retinal oxygen: fundamental and clinical aspects. Arch Ophthalmol 121(4):547-557

    Article  Google Scholar 

  • Wilson DF, Rumsey WL, Green TJ, Vanderkooi JM (1988) The oxygen dependence of mitochondrial oxidative phosphorylation measured by a new optical method for measuring oxygen concentration. J Biol Chem 263(6):2712-2718

    Google Scholar 

  • Yu DY, Cringle SJ (2001) Oxygen distribution and consumption within the retina in vascularised and avascular retinas and in animal models of retinal disease. Prog Retin Eye Res 20(2):175-208

    Article  Google Scholar 

  • Yu DY, Cringle SJ (2002) Outer retinal anoxia during dark adaptation is not a general property of mammalian retinas. Comp Biochem Physiol 132(1):47-52

    Article  Google Scholar 

  • Yu DY, Cringle SJ (2005) Retinal degeneration and local oxygen metabolism. Exp Eye Res 80(6):745-751

    Article  Google Scholar 

  • Yu DY, Cringle SJ, Alder VA, Su EN (1994) Intraretinal oxygen distribution in rats as a function of systemic blood pressure. Am J Physiol Heart Circ Physiol 267(6):H2498-H2507

    Google Scholar 

  • Yu DY, Cringle SJ, Alder V, Su EN (1999) Intraretinal oxygen distribution in the rat with graded systemic hyperoxia and hypercapnia. Invest Ophthalmol Vis Sci 40(9):2082-2087

    Google Scholar 

  • Yu DY, Cringle SJ, Su EN, Yu PK (2000) Intraretinal oxygen levels before and after photoreceptor loss in the RCS rat. Invest Ophthalmol Vis Sci 41(12):3999-4006

    Google Scholar 

  • Yu DY, Cringle S, Valter K, Walsh N, Lee D, Stone J (2004) Photoreceptor death, trophic factor expression, retinal oxygen status, and photoreceptor function in the P23H rat. Invest Ophthalmol Vis Sci 45(6):2013-2019

    Article  Google Scholar 

  • Yu DY, Cringle SJ, Su EN (2005) Intraretinal oxygen distribution in the monkey retina and the response to systemic hyperoxia. Invest Ophthalmol Vis Sci 46(12):4728-4733

    Article  Google Scholar 

Download references

Acknowledgments

We gratefully acknowledge the Engineering and Physical Sciences Research Council (EPSRC) in the UK for funding through a studentship at the Systems Biology programme of the University of Oxford’s Doctoral Training Centre P.A.R.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Paul A. Roberts.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (pdf 396 KB)

Appendices

Appendix 1: Placing a bound on the concentration of pentacoordinate neuroglobin

It was found in all single and eight layer model simulations that the concentration of pentacoordinate Ngb, n, at any given point in space is significantly lower than that of hexacoordinate Ngb, \(n_h\), or oxygen bound Ngb, \(n_o\), at that point. In this appendix we derive a bound on n to explain why this is the case.

Beginning with the time-dependent equation for n, Eq. (10), we can re-write it in the form

$$\begin{aligned} \frac{\partial n}{\partial t} = D\frac{\partial ^2 n}{\partial x^2} + f(x,t) - \left( \frac{1}{\kappa } + g(x,t)\right) n, \end{aligned}$$
(42)

where \(1/\kappa = k_3\), \(f(x,t) = k_4n_h + \alpha k_2n_o \ge 0\) and \(g(x,t) = \alpha k_1 c \ge 0\). Whether we are using the single or the eight layer model, this equation will be defined on a finite domain, which we can denote as \(x \in [0,L]\) without loss of generality. As usual, we impose zero-flux boundary conditions at either end of the domain.

Consider the function \(M = M(t)\) which satisfies

$$\begin{aligned} \frac{{\mathrm {d}} M}{{\mathrm {d}} t} = \left( \max _{x,t}(f(x,t)) + \beta \right) - \frac{1}{\kappa }M, \end{aligned}$$
(43)

where the constant \(\beta > 0\) is arbitrarily small. This can be written as

$$\begin{aligned} \frac{{\mathrm {d}} M}{{\mathrm {d}} t} = F - \frac{1}{\kappa }M, \end{aligned}$$
(44)

where \(F = (\max _{x,t}(f(x,t)) + \beta ) > f(x,t) \ge 0\) is a constant. We close the system by imposing the initial condition

$$\begin{aligned} M(0) = \max _{x}(n(x,0)). \end{aligned}$$
(45)

Solving this initial-value problem using the integrating factor method we obtain

$$\begin{aligned} M(t) = e^{-\frac{t}{\kappa }}M(0) + \kappa F(1 - e^{-\frac{t}{\kappa }}). \end{aligned}$$
(46)

Taking the limit as t tends to infinity we find that

$$\begin{aligned} \lim _{t \rightarrow \infty }M(t) = \kappa F. \end{aligned}$$
(47)

Let \(u(x,t) := n(x,t) - M(t)\), then

$$\begin{aligned} \frac{\partial u}{\partial t} - D\frac{\partial ^2 u}{\partial x^2} = (f - F) - \frac{u}{\kappa } - gn, \end{aligned}$$
(48)

from Eqs. (42) and (44). Since \((f - F) < 0\) and \(gn \ge 0\) we have that

$$\begin{aligned} \frac{\partial u}{\partial t} - D\frac{\partial ^2 u}{\partial x^2} + \frac{u}{\kappa } < 0, \end{aligned}$$
(49)

and therefore

$$\begin{aligned} \frac{\partial }{\partial t}\left( e^{\frac{t}{\kappa }}u\right) - D\frac{\partial ^2}{\partial x^2}\left( e^{\frac{t}{\kappa }}u\right) < 0. \end{aligned}$$
(50)

Defining \(v(x,t) := e^{\frac{t}{\kappa }}u(x,t)\), we have that

$$\begin{aligned} \frac{\partial v}{\partial t} - D\frac{\partial ^2 v}{\partial x^2}< 0, \end{aligned}$$
(51)

with initial condition

$$\begin{aligned} v(x,0) = n(x,0) - \max _{x}(n(x,0)), \end{aligned}$$
(52)

and zero-flux boundary conditions.

By the maximum principle for parabolic PDEs we have that v must achieve its maximum on one of the boundaries \(x = 0\), \(x = L\) or \(t = 0\) (Ockendon et al. 2003). Suppose that v takes its maximum value on \(x = 0\) for some \(t = t_1 > 0\). Since \(n_x(0,t_1) = 0\) by the zero-flux boundary condition and since \(M_x(t_1) = 0\), as M is independent of x, this implies that \(u_x(0,t_1) = 0\) and so \(v_x(0,t_1) = 0\). Therefore, in order for this point to be a maximum, we require both that \(v_t(0,t_1) = 0\) and \(v_{xx}(0,t_1) \le 0\). This implies that \(v_t(0,t_1) - Dv_{xx}(0,t_1) \ge 0\), which contradicts (51). Therefore, the maximum cannot lie on the boundary \(x = 0\). Using a similar argument we can show that the maximum cannot lie on the boundary \(x = L\). Therefore, the maximum must lie on the boundary \(t = 0\). As a check we can see that if the maximum lies on the boundary \(t = 0\) for some \(x = x_1\), where \(0 < x_1 < L\), then we must have that \(v_t(x_1,0) \le 0\), \(v_x(x_1,0) = 0\) and \(v_{xx}(x_1,0) \le 0\). Thus (51) may be satisfied, provided \(|v_t(x_1,0)| > D|v_{xx}(x_1,0)|\).

Since v(xt) takes its maximum value on \(t = 0\) this means that \(v(x,t) \le \max _{x}(v(x,0)) = \max _{x}(n(x,0) - \max _{x}(n(x,0))) = \max _{x}(n(x,0)) - \max _{x}(n(x,0)) = 0\). Therefore \(u \le 0\), which implies that \(n \le M\). Therefore, by (47), we have that at steady-state \(n \le \kappa F\). Taking the limit as \(\beta \) tends to zero, we find \(n \le \kappa \max _{x,t}(f(x,t)) = \kappa \max _{x,t}(k_4n_h + \alpha k_2n_o) \le \kappa \alpha k_2 n_T = \frac{\alpha k_2 n_T}{k_3} \approx 4n_T\times 10^{-4}\). This bound was found to be satisfied in all simulations.

Appendix 2: Asymptotic analysis

In the absence of Ngb, Eqs. (21)–(23) reduce to a single equation for oxygen:

$$\begin{aligned} 0 = \frac{{\mathrm {d}}^2c}{{\mathrm {d}}x^2} - \frac{Q_ic}{\gamma + c}, \end{aligned}$$
(53)

for \(Q_i\) \((1,\ldots ,7^*).\)

In layers 2 and 4, \(Q_2 = Q_4 = 0\), so that an exact analytical solution to Eq. (53) may be obtained. In all other layers, \(Q_i\) is strictly positive, so that an exact analytical solution cannot be derived. In these layers we look instead for a leading order solution. Since the value of \(Q_i\) is not, in general, continuous between layers, we cannot use standard matching techniques to construct a solution which is valid across all layers. Instead we use patching, ensuring that the solution in each layer satisfies the boundary conditions between it and the adjoining layers (see Bender and Orszag 1999, pgs. 335–336 for a discussion of patching).

We construct an asymptotic expansion for c(x):

$$\begin{aligned} c(x) \sim c_0(x) + \varepsilon c_1(x) + O(\varepsilon ^2) , \end{aligned}$$
(54)

choosing \(\varepsilon = 0.1\), so as to provide a clear separation between the various scales, and introduce \(\gamma ^* = \varepsilon ^{-2} \gamma \) so that \(\gamma ^* = O(1)\). We also note that \(Q_i = O(1)\) for \(i \in \{1,3,5,6^*,7^*\}\) and that \(Q_2 = Q_4 = 0\). Applying the scaling on \(\gamma \), Eq. (53) becomes

$$\begin{aligned} 0 = \frac{{\mathrm {d}}^2c}{{\mathrm {d}}x^2} - \frac{Q_ic}{\varepsilon ^2 \gamma ^* + c}. \end{aligned}$$
(55)

We now consider the leading order solution to Eq. (55) within each layer, grouping layers that have the same scaling.

1.1 Layer 1

In layer 1, \(c = O(1)\), as can be seen in Fig. 9, and Eq. (55) supplies, at leading order,

$$\begin{aligned} c(x) \sim c_0(x) = \frac{Q_1x^2}{2} + A_1x + B_1, \end{aligned}$$
(56)

where \(A_1\) and \(B_1\) are constants.

1.2 Layers 2 and 4

In layers 2 and 4, \(Q_2 = Q_4 = 0\), so that Eq. (55) can be solved exactly to yield

$$\begin{aligned} c(x) = A_ix + B_i, \end{aligned}$$
(57)

for \(i = 2, 4,\) where \(A_i\) and \(B_i\) are constants.

1.3 Layers 3, 5 and 7\(^*\)

In layers 3, 5 and 7\(^*\), \(c = O(\varepsilon )\), as can be seen in Fig. 9. Rescaling oxygen as \(c = \varepsilon c^*\), we then rescale x as \(x = \varepsilon ^{1/2}x_i^* + L_{i-1}\) in order to achieve a dominant balance, so that, after returning to our original scaling on c and x we have, at leading order,

$$\begin{aligned} c(x) \sim \frac{Q_ix^2}{2} + A_ix + B_i, \end{aligned}$$
(58)

for \(i \in \{3,5,7^*\}\), where \(A_i\) and \(B_i\) are constants.

1.4 Layer 6\(^*\)

The situation in layer 6\(^*\) is less straightforward. On the left- and right-hand sides of layer 6\(^*\), \(c = O(\varepsilon )\), whereas, toward the centre of the layer, in a neighbourhood around the local minimum, \(c = O(\varepsilon ^2)\). Where \(c = O(\varepsilon )\), we could scale c and x as in layers 3, 5 and 7\(^*\); however, in the central region, where \(c = O(\varepsilon ^2)\), we would regain Eq. (53) in the dominant balance after dropping the stars. Since seeking leading order solutions does not allow us to avoid dealing with Eq. (53), we instead retain Eq. (53) across the whole of layer 6\(^*\) and use quadrature methods to derive approximate analytical solutions for the oxygen profile in this layer.

We will derive separate approximations to the oxygen profile for the left-hand side, centre and right-hand side of layer 6\(^*\), in order to account for the variation in c between \(O(\varepsilon )\) and \(O(\varepsilon ^2)\). We will also derive approximations to the minimum oxygen concentration in layer 6\(^*\), \(c_{min}\), and its position, \(x_{min}\).

Multiplying Eq. (53) by \({\mathrm {d}}c/{\mathrm {d}}x\) and integrating between \(x_{min}\) and x, we find that

$$\begin{aligned} \frac{{\mathrm {d}}c}{{\mathrm {d}}x} = \pm \sqrt{2Q_{6^*}}\sqrt{c - c_{min} - \gamma \log {\left( \frac{\gamma + c}{\gamma + c_{min}}\right) }}, \end{aligned}$$
(59)

where we take the positive (negative) root to the right (left) of \(x_{min}\), since the gradient of the oxygen profile is positive (negative) there. We note that the values of \(c_{min}\) and \(x_{min}\) are unknown at this stage.

We begin by seeking the left-hand and right-hand approximations near \(L_5\) and \(L_{6^*}\) respectively. For the left-hand approximation, we integrate Eq. (59) between x and \(L_5\) to obtain

$$\begin{aligned} \int _{c(x)}^{c_L}\left( s - c_{min} - \gamma \log \left( \frac{\gamma + s}{\gamma + c_{min}}\right) \right) ^{-\frac{1}{2}}{\mathrm {d}}s = \sqrt{2Q_{6^*}}(x - L_5), \end{aligned}$$
(60)

where \(c_L = c(x=L_5)\) is unknown at this stage. Since \(c \approx c_L\) in the left-hand region, where \(c_L = O(\varepsilon )\), and since \(c_{min} = O(\varepsilon ^2)\) and \(\gamma = O(\varepsilon ^2)\) (see Fig. 9 for values of \(c_L\) and \(c_{min}\)), we may approximate the integrand by \(s^{-\frac{1}{2}}\) to obtain the left-hand approximation as

$$\begin{aligned} c(x) \approx \frac{Q_{6^*}}{2}(L_5 - x)^2 + \sqrt{2Q_{6^*}c_L}(L_5 - x) + c_L. \end{aligned}$$
(61)

In a similar way, we obtain the right-hand approximation:

$$\begin{aligned} c(x) \approx \frac{Q_{6^*}}{2}(x - L_{6^*})^2 + \sqrt{2Q_{6^*}c_R}(x - L_{6^*}) + c_R, \end{aligned}$$
(62)

where \(c_R = c(x=L_{6^*})\) is unknown at this stage. Both \(c_L\) and \(c_R\) may be found by applying the boundary conditions as described at the end of this section.

To derive the central approximation, valid in the neighbourhood of \(x_{min}\), we integrate Eq. (59) between \(x_{min}\) and x. Writing \(\log ((\gamma + s)/(\gamma + c_{min})) = \log (1 + (s - c_{min})/(\gamma + c_{min}))\), we expand the integrand (which is the same as in Eq. (60)) about \(s = c_{min}\), in powers of \((s - c_{min})/(\gamma + c_{min})\), where s is the variable of integration and \(|(s - c_{min})/(\gamma + c_{min})| \ll 1\). Retaining only the first term and neglecting higher order terms, we obtain the central solution:

$$\begin{aligned} c(x) \approx \left( 1 + \frac{Q_{6^*}(x_{min} - x)^2}{2(\gamma + c_{min})}\right) c_{min}, \end{aligned}$$
(63)

in which \(x_{min}\) and \(c_{min}\) are presently unknown.

To determine \(x_{min}\), we integrate Eq. (59) between the limits \(x_{min}\) and \(L_5\), and \(x_{min}\) and \(L_{6^*}\), to obtain the following pair of equations:

$$\begin{aligned} \int _{c_{min}}^{c_L}\left( s - c_{min} - \gamma \log \left( \frac{\gamma + s}{\gamma + c_{min}}\right) \right) ^{-\frac{1}{2}}{\mathrm {d}}s= & {} \sqrt{2Q_{6^*}}(x_{min} - L_5), \end{aligned}$$
(64)
$$\begin{aligned} \int _{c_{min}}^{c_R}\left( s - c_{min} - \gamma \log \left( \frac{\gamma + s}{\gamma + c_{min}}\right) \right) ^{-\frac{1}{2}}{\mathrm {d}}s= & {} \sqrt{2Q_{6^*}}(L_{6^*} - x_{min}). \end{aligned}$$
(65)

Subtracting Eq. (64) from Eq. (65), and noting that \(\min (c_L,c_R) < s < \max (c_L,c_R)\) in the integrand, where \(c_L = O(\varepsilon )\), \(c_R = O(\varepsilon )\), \(c_{min} = O(\varepsilon ^2)\) and \(\gamma = O(\varepsilon ^2)\) (see Fig. 9 for values of \(c_L\), \(c_R\) and \(c_{min}\)), we may approximate the integrand by \(s^{-\frac{1}{2}}\) to obtain

$$\begin{aligned} x_{min} \approx \frac{L_5 + L_{6^*}}{2} - \left( \frac{\sqrt{c_R} - \sqrt{c_L}}{\sqrt{2Q_{6^*}}}\right) . \end{aligned}$$
(66)

We obtain an implicit expression for \(c_{min}\), by substituting for \(x_{min}\) from Eq. (66) into Eq. (64):

$$\begin{aligned} \int _{c_{min}}^{c_L}\left( s - c_{min} - \gamma \log \left( \frac{\gamma + s}{\gamma + c_{min}}\right) \right) ^{-\frac{1}{2}}{\mathrm {d}}s \nonumber \\ \quad \approx \sqrt{\frac{Q_{6^*}}{2}}(L_{6^*} - L_5) - (\sqrt{c_R} - \sqrt{c_L}), \end{aligned}$$
(67)

where the integral must be calculated numerically. With \(x_{min}\) and \(c_{min}\) specified by (66) and (67) respectively, we can use Eq. (63) to calculate the central solution.

1.5 Using iteration to improve accuracy

We can iteratively improve the accuracy of our approximation using the exact derivative, given by Eq. (59), in the oxygen-flux boundary conditions at \(x=L_{6^*}\) and \(x=L_{7^*}\), given by Eq. (20), to yield:

$$\begin{aligned} 0= & {} \sqrt{2Q_{6^*}}\sqrt{c_L - c_{min} - \gamma \log \left( \frac{\gamma + c_L}{\gamma + c_{min}}\right) } + \nonumber \\&\displaystyle \quad (Q_5L_5 + A_5(c_L)) - {\hat{h}}_5(c_v - c_L),\nonumber \\ \end{aligned}$$
(68)
$$\begin{aligned} 0= & {} \sqrt{2Q_{6^*}}\sqrt{c_R - c_{min} - \gamma \log \left( \frac{\gamma + c_R}{\gamma + c_{min}}\right) } + \nonumber \\&\displaystyle \quad Q_{7^*}(L_{7^*} - L_{6^*}) - {\hat{h}}_{6^*}(c_v - c_R).\nonumber \\ \end{aligned}$$
(69)

The constant \(c_{min}\) takes the value calculated by solving (67), using the original values of \(c_L\) and \(c_R\), whilst the constant \(A_5\) is a function of \(c_L\), given by:

$$\begin{aligned} A_5(c_L) = \frac{1}{L_5}\left[ c_L - c_c - \frac{1}{2}\left( Q_5(L_5^2 + L_4^2) + Q_3(L_2^2-L_3^2) - Q_1L_1^2\right) \right] . \end{aligned}$$
(70)

Equations (68)–(69) can be solved numerically, using the Matlab routine fsolve, to find updated values for \(c_L\) and \(c_R\). The updated \(c_L\) and \(c_R\) can then be used to calculate an updated value for \(c_{min}\), using Eq. (67). We may then repeat the iteration, using the updated value of \(c_{min}\) in Eqs. (68)–(70), to find new values for \(c_L\) and \(c_R\). Once the solution has converged, the final values of \(c_L\) and \(c_R\) can then be used to calculate improved values for \(A_1,\ldots ,A_5,B_{7^*},x_{min}\) and \(c_{min}\), by applying the boundary conditions to Eq. (41) as described in Sect. 4.3 and using Eqs. (66) and (67).

It was found that iteration results in a small improvement in the accuracy of the approximation and that the solution does not change significantly (by \(O(10^{-4})\) or greater) after the third iteration. Therefore, we use the parameters generated by the third iteration for the approximate solution in Sect. 4.3.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Roberts, P.A., Gaffney, E.A., Luthert, P.J. et al. Retinal oxygen distribution and the role of neuroglobin. J. Math. Biol. 73, 1–38 (2016). https://doi.org/10.1007/s00285-015-0931-y

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-015-0931-y

Keywords

Mathematics Subject Classification

Navigation