Skip to main content
Log in

A Multiscale Analysis of Diffusions on Rapidly Varying Surfaces

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

Lateral diffusion of molecules on surfaces plays a very important role in various biological processes, including lipid transport across the cell membrane, synaptic transmission, and other phenomena such as exo- and endocytosis, signal transduction, chemotaxis, and cell growth. In many cases, the surfaces can possess spatial inhomogeneities and/or be rapidly changing shape. Using a generalization of the model for a thermally excited Helfrich elastic membrane, we consider the problem of lateral diffusion on quasi-planar surfaces, possessing both spatial and temporal fluctuations. Using results from homogenization theory, we show that, under the assumption of scale separation between the characteristic length and timescales of the membrane fluctuations and the characteristic scale of the diffusing particle, the lateral diffusion process can be well approximated by a Brownian motion on the plane with constant diffusion tensor \(D\) that depends on a highly nonlinear way on the detailed properties of the surface. The effective diffusion tensor will depend on the relative scales of the spatial and temporal fluctuations, and for different scaling regimes, we prove the existence of a macroscopic limit in each case.

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

Similar content being viewed by others

References

  • Abdulle, A., Schwab, C.: Heterogeneous multiscale FEM for diffusion problems on rough surfaces. Multiscale Model. Simul. 3(1), 195–220 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  • Aizenbud, B.M., Gershon, N.D.: Diffusion of molecules on biological membranes of nonplanar form. II. Diffusion anisotropy. Biophys. J. 48(4), 543–546 (1985)

    Article  Google Scholar 

  • Almeida, P.F.F., Vaz, W.L.C.: Lateral diffusion in membranes. Handb. Biol. Phys. 1, 305–357 (1995)

    Article  Google Scholar 

  • Ashby, M.C., Maier, S.R., Nishimune, A., Henley, J.M.: Lateral diffusion drives constitutive exchange of AMPA receptors at dendritic spines and is regulated by spine morphology. J. Neurosci. 26(26), 7046–7055 (2006)

    Article  Google Scholar 

  • Axelrod, D., Koppel, D.E., Schlessinger, J., Elson, E., Webb, W.W.: Mobility measurement by analysis of fluorescence photobleaching recovery kinetics. Biophys. J. 16(9), 1055–1069 (1976)

    Article  Google Scholar 

  • Bensoussan, A., Lions, J.L., Papanicolaou, G.: Asymptotic Analysis for Periodic Structures, vol. 5. North Holland, NY (1978)

    MATH  Google Scholar 

  • Berg, H.C.: Random Walks in Biology. Princeton University Press, Princeton (1993)

    Google Scholar 

  • Borgdorff, A.J., Choquet, D., et al.: Regulation of AMPA receptor lateral movements. Nature 417(6889), 649–653 (2002)

    Article  Google Scholar 

  • Bourgeat, A., Jurak, M., Piatnitski, A.L.: Averaging a transport equation with small diffusion and oscillating velocity. Math. Methods Appl. Sci. 26(2), 95–117 (2003)

    Article  MATH  MathSciNet  Google Scholar 

  • Brenner, S.C., Scott, L.R.: The Mathematical Theory of Finite Element Methods, vol. 15. Springer, Berlin (2008)

    MATH  Google Scholar 

  • Bressloff, P.C., Newby, J.M.: Stochastic models of intracellular transport. Rev. Mod. Phys. 85(1), 135 (2013)

    Article  Google Scholar 

  • Canham, P.B.: The minimum energy of bending as a possible explanation of the biconcave shape of the human red blood cell. J. Theor. Biol. 26(1), 61–81 (1970)

    Article  Google Scholar 

  • Coulibaly-Pasquier, K.A.: Brownian motion with respect to time-changing Riemannian metrics, applications to Ricci flow. In: Annales de l’Institut Henri Poincaré, Probabilités et Statistiques, vol. 47, pp. 515–538. Institut Henri Poincaré (2011)

  • Crank, J.: The Mathematics of Diffusion. Oxford University Press, Oxford (1979)

    MATH  Google Scholar 

  • Deckelnick, K., Dziuk, G., Elliott, C.M.: Computation of geometric partial differential equations and mean curvature flow. Acta Numer. 14, 139–232 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  • Doi, M., Edwards, S.F.: The Theory of Polymer Dynamics, vol. 73. Oxford University Press, Oxford (1988)

    Google Scholar 

  • Duncan, A.B.: Diffusion on Rapidly Varying Surfaces. Ph.D. thesis, University of Warwick (September 2013)

  • Dykhne, A.M.: Conductivity of a two-dimensional two-phase system. Sov. J. Exp. Theor. Phys. 32, 63 (1971)

    Google Scholar 

  • Dziuk, G., Elliott, C.M.: Finite element methods for surface PDEs. Acta Numer. 22, 289–396 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  • Festa, R., d’Agliano, E.G.: Diffusion coefficient for a Brownian particle in a periodic field of force: I. Large friction limit. Phys. A Stat. Mech. Appl. 90(2), 229–244 (1978)

    Article  Google Scholar 

  • Friedman, A.: Stochastic Differential Equations and Applications, dover ed. edition. Dover Books on Mathematics. Dover Publications, New York (2006)

    Google Scholar 

  • Garnier, J.: Homogenization in a periodic and time-dependent potential. SIAM J. Appl. Math. 57(1), 95–111 (1997)

    Article  MATH  MathSciNet  Google Scholar 

  • Gilbarg, D., Trudinger, N.S.: Elliptic Partial Differential Equations of Second Order, vol. 224. Springer, Berlin (2001)

    MATH  Google Scholar 

  • Gonzalez, O., Stuart, A.M.: A First Course in Continuum Mechanics. Cambridge University Press, Cambridge (2008)

    MATH  Google Scholar 

  • Gov, N.: Membrane undulations driven by force fluctuations of active proteins. Phys. Rev. Lett. 93(26), 268104 (2004)

    Article  Google Scholar 

  • Gov, N.S.: Diffusion in curved fluid membranes. Phys. Rev. E 73(4), 041918 (2006)

    Article  Google Scholar 

  • Granek, R.: From semi-flexible polymers to membranes: anomalous diffusion and reptation. J. Phys. II 7(12), 1761–1788 (1997)

    Google Scholar 

  • Gustafsson, S., Halle, B.: Diffusion on a flexible surface. J. Chem. Phys. 106, 1880 (1997)

    Article  Google Scholar 

  • Halle, B., Gustafsson, S.: Diffusion in a fluctuating random geometry. Phys. Rev. E 55(1), 680 (1997)

    Article  Google Scholar 

  • Helfrich, W., et al.: Elastic properties of lipid bilayers: theory and possible experiments. Z. Naturforsch. c 28(11), 693–703 (1973)

    Google Scholar 

  • Hsu, E.P.: Stochastic Analysis on Manifolds, vol. 38. American Mathematical Society, Providence, RI (2002)

    MATH  Google Scholar 

  • Jackson, J.L., Coriell, S.R.: Effective diffusion constant in a polyelectrolyte solution. J. Chem. Phys. 38, 959 (1963)

    Article  Google Scholar 

  • Karatzas, I.A., Shreve, S.E.: Brownian Motion and Stochastic Calculus, vol. 113. Springer, Berlin (1991)

    MATH  Google Scholar 

  • Keller, J.B.: Conductivity of a medium containing a dense array of perfectly conducting spheres or cylinders or nonconducting cylinders. J. Appl. Phys. 34(4), 991–993 (1963)

    Article  MATH  Google Scholar 

  • Khasminskii, R.Z., Yin, G.: Asymptotic series for singularly perturbed Kolmogorov–Fokker–Planck equations. SIAM J. Appl. Math. 56(6), 1766–1793 (1996)

    Article  MATH  MathSciNet  Google Scholar 

  • Kim, S., Karrila, S.J.: Microhydrodynamics: Principles and Selected Applications. Dover Publications, NY (1991)

    Google Scholar 

  • King, M.R.: Apparent 2-D diffusivity in a ruffled cell membrane. J. Theor. Biol. 227(3), 323–326 (2004)

    Article  Google Scholar 

  • Kohler, W., Papanicolaou, G.C.: Bounds for the effective conductivity of random media. In: Macroscopic Properties of Disordered Media, pp. 111–130. Springer, Berlin (1982)

  • Komorowski, T., Landim, C., Olla, S.: Fluctuations in Markov Processes: Time Symmetry and Martingale Approximation, vol. 345. Springer, Berlin (2012)

    Google Scholar 

  • Larsson, S., Thomée, V.: Partial Differential Equations with Numerical Methods, vol. 45. Springer, Berlin (2009)

    MATH  Google Scholar 

  • Leitenberger, S.M., Reister-Gottfried, E., Seifert, U.: Curvature coupling dependence of membrane protein diffusion coefficients. Langmuir 24(4), 1254–1261 (2008)

    Article  Google Scholar 

  • Lifson, S., Jackson, J.L.: On the self-diffusion of ions in a polyelectrolyte solution. J. Chem. Phys. 36, 2410 (1962)

    Article  Google Scholar 

  • Lin, L.C.L., Gov, N., Brown, F.L.H.: Nonequilibrium membrane fluctuations driven by active proteins. J. Chem. Phys. 124(7), 074903–074903 (2006)

    Article  Google Scholar 

  • Lin, L.C.L., Brown, F.L.H.: Dynamics of pinned membranes with application to protein diffusion on the surface of red blood cells. Biophys. J. 86(2), 764–780 (2004)

    Article  MathSciNet  Google Scholar 

  • Lindblom, G., Orädd, G.: NMR studies of translational diffusion in lyotropic liquid crystals and lipid membranes. Prog. Nucl. Magn. Reson. Spectrosc. 26, 483–515 (1994)

    Article  Google Scholar 

  • Matheron, G.: Eléments pour une théorie des milieux poreux. Masson, Paris (1967)

    Google Scholar 

  • Mattingly, J.C., Stuart, A.M.: Geometric ergodicity of some hypo-elliptic diffusions for particle motions. Markov Process. Relat. Fields 8(2), 199–214 (2002)

    MATH  MathSciNet  Google Scholar 

  • Mattingly, J.C., Stuart, A.M., Higham, D.J.: Ergodicity for SDEs and approximations: locally Lipschitz vector fields and degenerate noise. Stoch. Process. Appl. 101(2), 185–232 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  • Mendelson, K.S.: A theorem on the effective conductivity of a two-dimensional heterogeneous medium. J. Appl. Phys. 46(11), 4740–4741 (1975)

    Article  Google Scholar 

  • Meyn, S.P., Tweedie, R.L.: Markov Chains and Stochastic Stability. Cambridge University Press, Cambridge (2009)

    Book  MATH  Google Scholar 

  • Naji, A., Brown, F.L.H.: Diffusion on ruffled membrane surfaces. J. Chem. Phys. 126, 235103 (2007)

    Article  Google Scholar 

  • Papanicolaou, G.C.: Introduction to the asymptotic analysis of stochastic equations. Mod. Model. Contin. Phenom. (Ninth Summer Sem. Appl. Math., Rensselaer Polytech. Inst., Troy, NY, 1975) 16, 109–147 (1977)

    MathSciNet  Google Scholar 

  • Pardoux, É.: Homogenization of linear and semilinear second order parabolic pdes with periodic coefficients: a probabilistic approach. J. Funct. Anal. 167(2), 498–520 (1999)

    Article  MATH  MathSciNet  Google Scholar 

  • Pavliotis, G.A., Stuart, A.M., Zygalakis, K.C.: Homogenization for inertial particles in a random flow. Commun. Math. Sci. 5(3), 507–531 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  • Pavliotis, G.A., Stuart, A.M.: Multiscale Methods: Averaging and Homogenization. Springer, Berlin (2008)

    Google Scholar 

  • Poo, M., Cone, R.A., et al.: Lateral diffusion of rhodopsin in the photoreceptor membrane. Nature 247(441), 438 (1974)

    Article  Google Scholar 

  • Reister, E., Seifert, U.: Lateral diffusion of a protein on a fluctuating membrane. EPL (Europhys. Lett.) 71(5), 859 (2007)

    Article  Google Scholar 

  • Reister-Gottfried, E., Leitenberger, S.M., Seifert, U.: Hybrid simulations of lateral diffusion in fluctuating membranes. Phys. Rev. E 75(1), 011908 (2007)

    Article  Google Scholar 

  • Reister-Gottfried, E., Leitenberger, S.M., Seifert, U.: Diffusing proteins on a fluctuating membrane: analytical theory and simulations. Phys. Rev. E 81(3), 031903 (2010)

    Article  Google Scholar 

  • Risken, H.: The Fokker–Planck equation: methods of solution and applications, vol. 18. Springer, Berlin (1996)

    MATH  Google Scholar 

  • Saxton, M.J., Jacobson, K.: Single-particle tracking: applications to membrane dynamics. Annu. Rev. Biophys. Biomol. Struct. 26(1), 373–399 (1997)

    Article  Google Scholar 

  • Sbalzarini, I.F., Hayer, A., Helenius, A., Koumoutsakos, P.: Simulations of (an) isotropic diffusion on curved biological surfaces. Biophys. J. 90(3), 878 (2006)

    Article  Google Scholar 

  • Seifert, U.: Configurations of fluid membranes and vesicles. Adv. Phys. 46(1), 13–137 (1997)

    Article  Google Scholar 

  • Stuart, A.M.: Inverse problems: a Bayesian perspective. Acta Numer. 19(1), 451–559 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  • Van Den Berg, M., Lewis, J.T.: Brownian motion on a hypersurface. Bull. Lond. Math. Soc. 17(2), 144–150 (1985)

    Article  MATH  Google Scholar 

  • Van Kampen, N.G.: Stochastic Processes in Physics and Chemistry. North Holland, Amsterdam (2007)

    Google Scholar 

  • Zhikov, V.V., Kozlov, S.M., Oleinik, O.A.: Homogenization of Differential Operators and Integral Functionals. Springer, Berlin (1994)

    MATH  Google Scholar 

Download references

Acknowledgments

A.D. is grateful to EPSRC for financial support and thanks to the Centre for Scientific Computing @ Warwick for computational resources. G.P. acknowledges financial support from EPSRC Grant Nos. EP/J009636/1 and EP/H034587/1. A.M.S. is grateful to EPSRC and ERC for financial support.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. B. Duncan.

Additional information

Communicated by Paul Newton.

Appendix: Formal Multiscale Expansions

Appendix: Formal Multiscale Expansions

In this section, we formally derive the limiting equations in each of the scaling regimes using multiscale expansions.

1.1 Case I

To derive the homogenized equation in this regime, we make the ansatz that the solution \(v^\varepsilon \) of (35) is of the form

$$\begin{aligned} v^{\varepsilon } = v_{0}(x,y,t) + \varepsilon v_{1}(x,y,t) + \varepsilon ^{2} v_{2}(x,y,t) + \cdots , \end{aligned}$$
(108)

for smooth \(v_{i}:\mathbb {R}^{d}\times {\mathbb {T}}^{d} \times [0,T] \rightarrow \mathbb {R}^d\). Substituting (108) in (35) and identifying equal powers of \(\varepsilon \), we obtain the following equations

$$\begin{aligned}&\begin{array}{ll} O\left( \frac{1}{\varepsilon ^2}\right) : &{} \quad {\fancyscript{L}}_{0}v_{0}(x,y,t) = 0,\\ \end{array}\end{aligned}$$
(109)
$$\begin{aligned}&\begin{array}{lr} O\left( \frac{1}{\varepsilon }\right) :&{} \quad {\fancyscript{L}}_{0}v_{1}(x,y,t) = -{\fancyscript{L}}_{1}v_{0}(x,y,t),\\ \end{array}\end{aligned}$$
(110)
$$\begin{aligned}&\begin{array}{lr} O\left( 1\right) : &{}\quad {\fancyscript{L}}_{0}v_{2}(x,y,t) = \frac{\partial v_{0}}{\partial t} -{\fancyscript{L}}_{1}v_{1}(x,y,t) -{\fancyscript{L}}_{2}v_{0}(x,y,t),\\ \end{array} \end{aligned}$$
(111)

for \((x,y,t) \in \mathbb {R}^d \times {\mathbb {T}}^d \times (0,T]\).

Since the nullspace of \({\fancyscript{L}}_{0}\) contains only constants in \(y\), Eq. (109) thus implies that \(v_{0}\) is a function of \(x\) and \(t\) only. Equation (110) becomes

$$\begin{aligned} {\fancyscript{L}}_{0}v_{1}(x,y,t) = -F(y)\cdot \nabla _{x}v_{0}(x,t). \end{aligned}$$
(112)

Let \(\chi \in C^2({\mathbb {T}}^d; \mathbb {R}^{d})\) be the unique, mean-zero solution of the cell Eq. (39). If we choose \(v_1 = \chi \cdot \nabla _x v_0(x,t)\), then it is clear that \(v_{1}\) solves (110).

Finally, by the Fredholm alternative on \({\fancyscript{L}}_0\), a necessary condition for Eq. (111) to have a solution is that the RHS of (111) has mean zero with respect to the measure \(\rho \), that is,

$$\begin{aligned} \frac{\partial v_{0}(x,t)}{\partial t} = \frac{1}{Z}\int _{{\mathbb {T}}^{d}} {\fancyscript{L}}_{1}v_{1}(x,y,t) \rho (y)\,\hbox {d}y + \frac{1}{Z}\int _{{\mathbb {T}}^{d}}{\fancyscript{L}}_{2}v_{0}(x,t) \rho (y)\,\hbox {d}y. \end{aligned}$$

Substituting \(v_{0}\) and \(v_{1}\), we obtain

$$\begin{aligned} \frac{\partial v_{0}(x,t)}{\partial t}&= \frac{1}{Z}\int _{{\mathbb {T}}^{d}} \nabla _{y}\cdot \left( \sqrt{\left|g \right|(y)}g^{-1}(y)\right) \cdot \nabla _{x}\left( \chi \cdot \nabla _{x}v_{0}(x,t)\right) \, \hbox {d}y \\&\qquad + \frac{2}{Z}\int _{{\mathbb {T}}^{d}}\sqrt{\left|g \right|(y)}g^{-1}(y):\nabla _{x}\nabla _{y}\left( \chi \cdot \nabla _{x}v_{0}(x,t)\right) \,\hbox {d}y \\&\qquad +\frac{1}{Z}\int _{{\mathbb {T}}^{d}}\sqrt{\left|g \right|(y)}g^{-1}(y):\nabla _{x}\nabla _{x}v_{0}(x,t)\,\hbox {d}y. \end{aligned}$$

Integrating the second term by parts with respect to \(y\) and simplifying, we obtain

$$\begin{aligned} \frac{\partial v_{0}(x,t)}{\partial t} = \Bigl (\frac{1}{Z}\int _{{\mathbb {T}}^{d}} g^{-1}(y)\bigl (I + \nabla _{y}\chi (y)\bigl ) \sqrt{\left|g \right|(y)}\, \hbox {d}y\Bigl ):\nabla _{x}\nabla _{x}v_{0}(x,t), \end{aligned}$$

where we have used the symmetry of \(g^{-1}\). Thus, the homogenized diffusion equation for \(v_0\) is

$$\begin{aligned} \frac{\partial v_{0}(x,t)}{\partial t} = D:\nabla _{x}\nabla _{x}v_{0}(x,t), \end{aligned}$$
(113)

where

$$\begin{aligned} D = \frac{1}{Z}\int _{{\mathbb {T}}^{d}} g^{-1}(y)\bigl (I + \nabla _{y}\chi (y)\bigl ) \sqrt{\left|g \right|(y)}\, \hbox {d}y. \end{aligned}$$

Multiplying (39) by \(\chi (y)\rho (y)\) and integrating by parts gives

$$\begin{aligned} \int _{{\mathbb {T}}^{d}}\left( I + \nabla _{y}\chi (y)\right) ^{\top } g^{-1}(y)\nabla _{y}\chi (y)\,\sqrt{\left|g \right|(y)}\,\hbox {d}y = 0, \end{aligned}$$

so that the effective diffusion matrix can be written in the following symmetric form

$$\begin{aligned} D = \frac{1}{Z}\int _{{\mathbb {T}}^{d}} \bigl (I + \nabla _{y}\chi (y)\bigl )^\top g^{-1}(y)\bigl (I + \nabla _y \chi (y)\bigl ) \sqrt{\left|g \right|(y)}\, \hbox {d}y. \end{aligned}$$

From the limiting backward Kolmogorov equation (113), we can read off the limiting SDE \(\hbox {d}X^0(t) = \sqrt{2D}\,\hbox {d}B(t)\). A rigorous proof of this result can found in Duncan (2013).

1.2 Case II

Analogous to the previous case, we look for solutions \(v\) of the form

$$\begin{aligned} v^\varepsilon (x ,\eta , t) = v_{0}(x,\eta ,t) + \varepsilon v_{1}(x,\eta ,t) + \cdots , \end{aligned}$$

for some smooth functions \(v_i:\mathbb {R}^d \times \mathbb {R}^K\times [0,T] \rightarrow \mathbb {R}^d\). Substituting this ansatz in (72a) and identifying equal powers of \(\varepsilon \), we obtain the following pair of equations

$$\begin{aligned}&\begin{array}{ll} O\left( \frac{1}{\varepsilon }\right) : &{}\quad {\fancyscript{L}}_0 v_0(x, \eta ,t) = 0, \\ \end{array}\end{aligned}$$
(114)
$$\begin{aligned}&\begin{array}{ll} O(1): &{}\quad \frac{\partial v(x, \eta , t)}{\partial t} = {\fancyscript{L}}_{0}v_1(x,\eta , t) + {\fancyscript{L}}_{1}v_0(x, \eta , t),\\ \end{array} \end{aligned}$$
(115)

where \((x, \eta , t) \in \mathbb {R}^d \times \mathbb {R}^K \times (0,T].\)

The \(O(\frac{1}{\varepsilon })\) equation immediately implies that \(v_0\) is independent of the fast-scale fluctuations. The second equation then becomes

$$\begin{aligned} {\fancyscript{L}}_0 v_1(x, \eta , t) = \frac{\partial v(x, \eta , t)}{\partial t} - {\fancyscript{L}}_1 v_0(x, \eta , t). \end{aligned}$$

Applying the Fredholm alternative, a necessary condition for the existence of a solution \(v_1\) is that the RHS is orthogonal to the invariant measure \(\rho _\eta \), that is,

$$\begin{aligned} \frac{\partial v_0}{\partial t}(x,t) = \left[ \int _{\mathbb {R}^K}F(x,\eta )\rho _\eta (\eta )\right] \cdot \nabla v_0(x,t) + \left[ \int _{\mathbb {R}^K}\varSigma (x,\eta )\rho _\eta (\eta )\right] :\nabla \nabla v_0(x,t), \end{aligned}$$

which is the backward equation for SDE (74).

1.3 Case III

We make the ansatz that

$$\begin{aligned} v^\varepsilon = v_{0} + \varepsilon v_{1} + \varepsilon ^{2} v_{2} + \cdots , \end{aligned}$$

for some smooth functions \(v_i:\mathbb {R}^d \times {\mathbb {T}}^d \times \mathbb {R}^K \times [0,T] \rightarrow \mathbb {R}\). Substituting \(v^\varepsilon \) in (82a) and identifying equal powers of \(\varepsilon \), we obtain the following equations

$$\begin{aligned}&\begin{array}{ll} O\left( \frac{1}{\varepsilon ^{2}}\right) : &{}\qquad {\fancyscript{L}}_{0}v_{0} = 0, \\ \end{array}\end{aligned}$$
(116)
$$\begin{aligned}&\begin{array}{ll} O\left( \frac{1}{\varepsilon }\right) : &{}\qquad {\fancyscript{L}}_{0}v_{1} = -{\fancyscript{L}}_{\eta }v_{0} -{\fancyscript{L}}_{1}v_{0} ,\\ \end{array}\end{aligned}$$
(117)
$$\begin{aligned}&\begin{array}{ll} O(1): \qquad \qquad &{}{\fancyscript{L}}_{0}v_{2} = -\left( \frac{\partial v_{0} }{\partial t} - {\fancyscript{L}}_{\eta }v_{1} - {\fancyscript{L}}_{1}v_{1} - {\fancyscript{L}}_{2}v_{0}\right) .\\ \end{array} \end{aligned}$$
(118)

The first equation implies that \(v_{0} \in {\fancyscript{N}}[{\fancyscript{L}}_{0}]\) so that \(v_{0}\) is a constant in \(y\). The second equation thus becomes

$$\begin{aligned} {\fancyscript{L}}_{0}v_1(x,y, \eta , t) = \left( {\fancyscript{L}}_{\eta }v_{0}(x, \eta , t) + F(y,\eta )\cdot \nabla _{x}v_{0}(x, \eta , t)\right) . \end{aligned}$$

By the Fredholm alternative applied to \({\fancyscript{L}}_{0}\), we require that the RHS is centered with respect to \(\sqrt{\left|g \right|}\), for each fixed \(x\) and \(\eta \) that is,

$$\begin{aligned} \int _{{\mathbb {T}}^{d}}\left( F(y,\eta )\cdot \nabla _{x}v_{0}(x,\eta , t) + {\fancyscript{L}}_{\eta }v_{0}(x, \eta , t) \right) \sqrt{\left|g \right|(y,\eta )}\,\hbox {d}y = 0. \end{aligned}$$

The first term in the above integral is clearly \(0\). Since \({\fancyscript{L}}_{\eta }v_{0}\) is independent of \(y\), the centering condition becomes

$$\begin{aligned} Z(\eta ){\fancyscript{L}}_{\eta }v_{0}(x, \eta , t) = 0. \end{aligned}$$

Since \(Z > 1\), it follows that \(v_{0} \in {\fancyscript{N}}[{\fancyscript{L}}_{\eta }]\) is a sufficient condition for the centering condition to hold, which we therefore will assume. By ergodicity of the Ornstein–Uhlenbeck process \(\eta (t)\) over \(\mathbb {R}^{K}\), it follows that \(v_{0}\) is also independent of \(\eta \) so that \(v_{0}\) is a function of \(x\) only. The second equation thus becomes

$$\begin{aligned} {\fancyscript{L}}_{0}v_1 = F(y,\eta )\cdot \nabla _{x}v_{0}. \end{aligned}$$

Let \(\chi (\cdot , \eta )\) be the unique, mean-zero solution of the cell equation solution by the Fredholm alternative, since the centering condition holds. Choosing \(v_{1} = \chi \cdot \nabla _{x}v_{0}\), it is clear that \(v_1\) solves the \(O(\frac{1}{\varepsilon })\) equation.

We now consider the \(O(1)\) equation. By the Fredholm alternative, a necessary condition for the existence of a unique solution \(v_{2}\) is that the RHS is centered with respect to the invariant measure of \({\fancyscript{L}}_{0}\), that is,

$$\begin{aligned} \frac{\partial v_{0}}{\partial t} =\int _{{\mathbb {T}}^{d}}\left( {\fancyscript{L}}_{\eta }v_{1} + {\fancyscript{L}}_{1}v_{1} + {\fancyscript{L}}_{2}v_{0}\right) \rho _{y}\, \hbox {d}y, \end{aligned}$$

which, substituting the definitions of the \({\fancyscript{L}}_{i}\)’s and \(v_{j}\)’s, can be written as follows:

$$\begin{aligned} \frac{\partial v_{0}}{\partial t}&= \int _{{\mathbb {T}}^{d}} F\otimes \chi \, \rho _{y}\,\hbox {d}y: \nabla _{x}\nabla _{x}v_{0} \end{aligned}$$
(119a)
$$\begin{aligned}&\quad + \int _{{\mathbb {T}}^{d}} g^{-1}\nabla _{y}\chi \rho _{y} + \nabla _{y}\chi \, g^{-1}\rho _{y}\,\hbox {d}y: \nabla _{x}\nabla _{x}v_{0} \end{aligned}$$
(119b)
$$\begin{aligned}&\quad + \int _{{\mathbb {T}}^{d}} g^{-1}\rho _{y}\,\hbox {d}y:\nabla _{x}\nabla _{x}v_{0} \end{aligned}$$
(119c)
$$\begin{aligned}&\quad + \int _{{\mathbb {T}}^{d}} {\fancyscript{L}}_{\eta }\chi \rho _{y}\,\hbox {d}y\, \nabla _{x}v_{0}. \end{aligned}$$
(119d)

First, we note that

$$\begin{aligned}&\int _{{\mathbb {T}}^{d}} F(y,\eta )\otimes \chi (y, \eta ) \sqrt{\left|g \right|(y, \eta )}\,\hbox {d}y = -\int _{{\mathbb {T}}^{d}} {\fancyscript{L}}_{0}\chi (y, \eta )\otimes \chi (y, \eta ) \sqrt{\left|g \right|(y, \eta )}\,\hbox {d}y \\&\quad = \int _{{\mathbb {T}}^{d}}\nabla _{y}\chi (y, \eta ) g^{-1}(y, \eta ) \nabla _{y}\chi (y, \eta )^\top \sqrt{\left|g \right|(y,\eta )}\, \hbox {d}y, \end{aligned}$$

so that we can write (119) as

$$\begin{aligned} \frac{\partial v_{0}}{\partial t}&= \int _{{\mathbb {T}}^{d}}\left( I + \nabla _{y}\chi \right) g^{-1}\left( I + \nabla _{y}\chi \right) ^\top \rho _{y}\, \hbox {d}y: \nabla _{x}\nabla _{x}v_{0}\\&+ \int _{{\mathbb {T}}^{d}} {\fancyscript{L}}_{\eta }\chi \rho _{y}\,\hbox {d}y \cdot \nabla _{x}v_{0}. \end{aligned}$$

Averaging with respect to the invariant measure \(\rho _{\eta }\) of \({\fancyscript{L}}_{1}\), we derive the effective diffusion equation

$$\begin{aligned} \frac{\partial v_{0}}{\partial t}&= \int \int _{{\mathbb {T}}^{d}}\left( I + \nabla _{y}\chi \right) g^{-1}\left( I + \nabla _{y}\chi \right) ^\top \rho _{y}\, \rho _{\eta }\, \hbox {d}y\,\hbox {d}\eta : \nabla _{x}\nabla _{x}v_{0}\\&+ \int \int _{{\mathbb {T}}^{d}} {\fancyscript{L}}_{\eta }\chi \rho _{y}\rho _{\eta }\,\hbox {d}y\, \hbox {d}\eta \cdot \nabla _{x}v_{0}, \end{aligned}$$

or more compactly

$$\begin{aligned} \frac{\partial v_{0}}{\partial t} = D: \nabla _{x}\nabla _{x}v_{0} + V\cdot \nabla _{x}v_{0}, \end{aligned}$$

where \(D\) and \(V\) are given by (86) and (87), respectively. From the limiting backward Kolmogorov equation, we can read off the limiting SDE (85) for the process \(X^{\varepsilon }(t)\).

1.4 Case IV

We look for solutions \(v\) of the form

$$\begin{aligned} v^\varepsilon = v_0 + \varepsilon v_1 + \varepsilon ^2 v_2 + \cdots \end{aligned}$$

of (103a) for some smooth functions \(v_i:\mathbb {R}^d \times {\mathbb {T}}^d \times \mathbb {R}^K \times [0,T]\rightarrow \mathbb {R}\). Substituting this ansatz in (103a) and equating equal powers of \(\varepsilon \), we obtain the following three equations

$$\begin{aligned}&\begin{array}{ll} O\left( \frac{1}{\varepsilon ^{2}}\right) : &{}\qquad \fancyscript{G}v_{0} = 0, \\ \end{array}\end{aligned}$$
(120)
$$\begin{aligned}&\begin{array}{ll} O\left( \frac{1}{\varepsilon }\right) :&{}\qquad \fancyscript{G}v_{1} = -{\fancyscript{L}}_{1}v_{0},\\ \end{array}\end{aligned}$$
(121)
$$\begin{aligned}&\begin{array}{ll} O(1):&\qquad \fancyscript{G}v_{2} = -\left( \frac{\partial v_{0}}{\partial t} - {\fancyscript{L}}_{1}v_{1} - {\fancyscript{L}}_{2}v_{0}\right) . \end{array} \end{aligned}$$
(122)

As the fast process is ergodic, the first equation implies that \(v_0\) is independent of \(y\) and \(\eta \). The second equation thus becomes

$$\begin{aligned} \fancyscript{G}v_1 = -F(y, \eta )\cdot \nabla _x v_0. \end{aligned}$$

Since we are assuming Assumption (101), there exists a unique solution of the Poisson problem (100), by Proposition 8. By choosing \(v_1 = \chi \cdot \nabla _x v_0\), we see that the second equation is satisfied.

By Proposition 8, a sufficient condition for the final equation to have a solution is that the RHS is orthogonal to the measure \(\rho (\hbox {d}y, \hbox {d}\eta )\), (assuming that the RHS grows at most polynomially), that is,

$$\begin{aligned} \frac{\partial v_0}{\partial t}(y, \eta )&= \int F(y, \eta )\cdot \nabla _x v_1 \, \rho (\hbox {d}y, \hbox {d}\eta )+ \int 2\varSigma (y, \eta ):\nabla _x \nabla _y v_1 \, \rho (\hbox {d}y, \hbox {d}\eta ) \\&\quad + \int \varSigma (y,\eta ):\nabla _x\nabla _x v_0 \, \rho (\hbox {d}y, \hbox {d}\eta ), \end{aligned}$$

which we can rewrite as

$$\begin{aligned} \frac{\partial v_0}{\partial t} = D:\nabla _x\nabla _x v_0, \end{aligned}$$

where the effective diffusion tensor \(D\) is given by

$$\begin{aligned} D = \int \left[ \frac{1}{\sqrt{\left|g \right|}}\nabla _y\cdot \left( \sqrt{\left|g \right|}g^{-1}\right) \otimes \chi + g^{-1}\nabla _y \chi ^\top + \nabla _y \chi g^{-1} + g^{-1}\, \right] \rho (\hbox {d}y, \hbox {d}\eta ) \end{aligned}$$

Note that the first term on the RHS

$$\begin{aligned} \displaystyle \int \frac{1}{\sqrt{\left|g \right|}}\nabla _y\cdot \left( \sqrt{\left|g \right|}g^{-1}\right) \otimes \chi \, \rho (\hbox {d}y, \hbox {d}\eta ):\nabla _x \nabla _x v_0, \end{aligned}$$

can be rewritten as \(\fancyscript{K}:\nabla _x \nabla _x v_0,\) where

$$\begin{aligned} \fancyscript{K} = \text{ Sym }\left[ \displaystyle \int \frac{1}{\sqrt{\left|g \right|}}\nabla _y\cdot \left( \sqrt{\left|g \right|}g^{-1}\right) \otimes \chi \, \rho (\hbox {d}y, \hbox {d}\eta )\right] , \end{aligned}$$

where \(\text{ Sym }\left[ \cdot \right] \) denotes the symmetric part of the matrix. Let \(e \in \mathbb {R}^d\) be a unit vector and consider

$$\begin{aligned} \fancyscript{K}^e := e \cdot \fancyscript{K} e = \displaystyle \int \frac{1}{\sqrt{\left|g \right|}}\nabla _y\cdot \left( \sqrt{\left|g \right|}g^{-1}e\right) \chi ^e\, \rho (\hbox {d}y, \hbox {d}\eta ), \end{aligned}$$

where \(\chi ^e = \chi \cdot e\). Noting that

$$\begin{aligned} -\fancyscript{G}\chi ^e = \frac{1}{\sqrt{\left|g \right|}}\nabla _y\cdot \left( \sqrt{\left|g \right|}g^{-1}e\right) , \end{aligned}$$

it follows that

$$\begin{aligned} \fancyscript{K}^e = \int \chi ^e\left( -\fancyscript{G}\chi ^e\right) \, \rho (\hbox {d}y, \hbox {d}\eta ), \end{aligned}$$

which, by (102), can be written as

$$\begin{aligned} \fancyscript{K}^e = \int \nabla _y \chi ^e\cdot g^{-1} \nabla _y \chi ^e \,\rho (\hbox {d}y, \hbox {d}\eta ) + \int \nabla _\eta \chi ^e \cdot \varGamma \Pi \nabla _\eta \chi ^e \, \rho (\hbox {d}y, \hbox {d}\eta ), \end{aligned}$$

so that

$$\begin{aligned} e\cdot D e \!=\! \int \left( e + \nabla _y\chi ^e\right) \cdot g^{-1}\left( e \!+\! \nabla _y\chi ^e\right) \, \rho (\hbox {d}y, \hbox {d}\eta ) \!+\! \int \nabla _\eta \chi ^e \cdot \varGamma \Pi \nabla _\eta \chi ^e \,\rho (\hbox {d}y, \hbox {d}\eta ), \end{aligned}$$

or in matrix notation

$$\begin{aligned} D = \int \left( I + \nabla _y\chi \right) g^{-1}\left( I + \nabla _y\chi \right) ^\top \, \rho (\hbox {d}y, \hbox {d}\eta ) + \int \nabla _\eta \chi \, \varGamma \Pi \nabla _\eta \chi ^\top \,\rho (\hbox {d}y, \hbox {d}\eta ), \end{aligned}$$

as required. A rigorous proof of this result can be found in Duncan (2013).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Duncan, A.B., Elliott, C.M., Pavliotis, G.A. et al. A Multiscale Analysis of Diffusions on Rapidly Varying Surfaces. J Nonlinear Sci 25, 389–449 (2015). https://doi.org/10.1007/s00332-015-9237-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00332-015-9237-x

Keywords

Mathematics Subject Classification

Navigation