- 1Facultad de Ciencias en Física y Matemáticas, Universidad Autónoma de Chiapas, Tuxtla Gutiérrez, Mexico
- 2División de Ciencias e Ingenierías, Universidad de Guanajuato, León, Mexico
- 3Departamento de Ingeniería Física, División de Ciencias e Ingenierías, Universidad de Guanajuato, León, Mexico
Brownian motion is a universal characteristic of colloidal particles embedded in a host medium, and it is the fingerprint of molecular transport or diffusion, a generic feature of relevance not only in physics but also in several branches of science and engineering. Since its discovery, Brownian motion, also known as colloidal dynamics, has been important in elucidating the connection between the molecular details of the diffusing macromolecule and the macroscopic information on the host medium. However, colloidal dynamics is far from being completely understood. For instance, the diffusion of non-spherical colloids and the effects of the underlying geometry of the host medium on the dynamics of either passive or active particles are a few representative cases that are part of the current challenges in soft matter physics. In this contribution, we take a step forward to introduce a covariant description of the colloidal dynamics in curved spaces. Without the loss of generality, we consider the case where hydrodynamic interactions are neglected. This formalism will allow us to understand several phenomena, for instance, the curvature effects on the kinetics during spinodal decomposition and the thermodynamic properties of colloidal dispersion, to mention a few examples. This theoretical framework will also serve as the starting point to highlight the role of geometry on colloidal dynamics, an aspect that is of paramount importance to understanding more complex transport phenomena, such as the diffusive mechanisms of proteins embedded in cell membranes.
1 Introduction
Since the pioneering work of Einstein [1], Brownian motion has become the paradigm for the description and understanding of a large variety of diffusion processes that are present in numerous physical, biological, and chemical systems. In recent years, the dynamics of macromolecules and nanoparticles on surfaces or curved spaces has been the subject of intensive investigations, especially because particle diffusion shows a richer dynamical behavior at different time scales [2, 3] than its counterpart in open and flat geometries. In particular, diffusion plays a key role in the dynamics of molecular motors moving along heterogeneous substrates [4], in the transport of biomacromolecules in the cell due to crowding [5, 6], and in the lateral diffusion of proteins on fluctuating membranes [7, 8]. Membranes can be simplified as two-dimensional differentiable manifolds based on their typical length scales. This approach allows us to describe many of its characteristics, such as the shape and dynamics, as if they were continuous pieces of material [9–11]. As a result, particles that move along the membrane due to certain underlying interactions can be explained as the random motion of particles restricted to the surface. Although this perspective might ignore relevant molecular details, most of the diffusion properties of tiny particles confined1 on surfaces depend strongly on the generic features of the surface or, strictly speaking, on the surface geometry [12]. Typically, particle dynamics is not only influenced by geometrical features but also by local and thermodynamic properties that experience the effects of the geometry of the manifold where the particles are embedded [13–15]. In a more general setting, the relevance of studying Brownian motion in differential manifolds is due to its ability to describe the system of Brownian particles with holonomic constraints that can model their interactions with the medium that houses them, as explicitly discussed in [16, 17] and recently in [18].
A great effort for understanding Brownian motion on surfaces can be found in colloidal soft matter, where the dynamics of colloidal particles on quasi-two-dimensional geometries have been both experimentally and theoretically investigated by using optical techniques such as digital videomicroscopy, computer simulations, and theoretical approximations [19, 20]. Nonetheless, such investigations deal basically with (almost) flat surfaces, i.e., without including curvature effects. The interest in the use of colloids resides in the fact that they are small (nanometer to micrometer-sized) particles and typically are considered model systems because of other interesting features [21]. Their characteristic time and length scales are experimentally accessible, which allows us to follow the colloidal dynamics and transport processes in real-time [21]. Furthermore, since the colloidal interactions are relatively weak, colloids are highly susceptible to external forces, and hence, their static and dynamical properties can be controlled through the application of external fields or by imposing geometrical restrictions. In other words, it is assumed that the molecular forces holding particles to the manifold are stronger than the forces between colloids (see, e.g., Ref. [21] and references therein). Then, colloids represent an ideal model system to account for the effects of geometry on the nature and dynamics of many-body systems.
In particular, it has already been demonstrated and experimentally corroborated that the curvature dependence of a fluctuating membrane affects the diffusion processes of molecules on the membrane surface [22–24]. These geometrical effects, although important, are still difficult to interpret. The lack of a precise interpretation resides in the fact that, unfortunately, there is not a unique way to define diffusion observables on a curved surface (see, for instance [22, 23, 25]). The description of colloidal dynamics in curved spaces is a non-trivial task; it represents a formidable physical and mathematical challenge. Recently, one of us proposed the generalization of the Smoluchowski equation on curved spaces [26]. Furthermore, Castro–Villarreal also put forward different geometrical observables to quantify the displacement of a single colloidal particle [25]. Within this approach, it was shown that the geodesic mean-squared displacement captures the intrinsic elements of the manifold, whereas the Euclidean displacement provides extrinsic information from the surface. An interesting extension of the theory now provides the description of the motion of active Brownian particles [27], where the mean-squared geodesic displacement captures the relationship between the curvature and the activity of the active colloid. This theoretical framework provided evidence that an active Brownian particle experiences a dynamical transition in any compact surface from a monotonic to an oscillating behavior observed in the mean-squared geodesic displacement [27]; a theoretical prediction of a dynamic transition of this type can be established using a run-and-tumble active particle confined on circle S1. This prediction was recently corroborated in experiments using a non-vibrating magnetic granular system (see, e.g., Ref. [28]). However, we still face challenges in colloidal dynamics on curved manifolds, for example, the generalization of this approach to the situation where the colloids interact not only with other macromolecules, i.e., direct forces, but also with the inclusion of all those geometrical mechanisms originating from the curvature and to situations when passive and active colloids are not just restricted geometrically but also are immersed in a viscoelastic fluid [29]. Furthermore, the transitions occurring in compact manifolds previously are examples of a non-trivial recurrence. It would be interesting to approach the problem from the perspective of dynamical system theory [30].
The aforementioned theoretical formalism has also allowed us to determine the equation of motion of interacting colloids in curved spaces; a generalized Ermack–McCammon algorithm has been developed to study a broader class of transport phenomena in curved manifolds [31]. Interestingly, the predictions of the particle transport in non-Euclidean spaces have been partially corroborated in a series of experiments [3, 24]; superparamagnetic colloids embedded in a circle and subjected to external magnetic fields [3] and polystyrene nanoparticles diffusing on highly curved water–silicone oil interfaces [24]. However, further experimental, computational, and theoretical studies are needed to better understand the rich diffusion mechanisms, particle distribution, and thermodynamic properties that emerge in colloidal dispersions when the curvature of the space plays an important role.
From a rigorous mathematical perspective, the study of stochastic processes on differential manifolds has a long history, dating back almost a century ago. A. Kolmogoroff made a seminal contribution to the covariant formulation of the Fokker–Planck (FP) equation (32), which was further developed in [33, 34] by proving the existence of solutions to the covariant FP equation and stochastic differential equations in manifolds, respectively. In addition, through different limiting cases reported in [35], when t → 0, asymptotic expressions were observed for the probability density function ρ(ξ, ξ′, t) in the cases of nearby and distant points [35]. These covariant versions of the Fokker–Planck equation are also introduced in [36], where one can identify a Riemannian geometry defined by a metric tensor given by the diffusion tensor [37], while a recent formulation starting from a coordinate covariance and gauge invariance in [38] was able to show the existence of a Langevin equation with a space-dependent diffusion matrix, whose corresponding Fokker–Planck equation is also given in a covariant form. The covariant formulation of stochastic equations has been further developed intensely during the last decade (e.g., see, [18, 39]). In this work, we take a little step forward to develop a covariant approach for a system of interacting colloidal particles in a sub-manifold of the Euclidean space
The article is structured as follows. Section 2 introduces the covariant approach of the colloidal dynamics from a stochastic differential equation in the Stratonovich sense. In particular, we can simplify the equation for the N-particle system moving on
2 Covariant approach of the colloidal dynamics
As discussed previously, one of the main challenges in understanding the effects of geometry on the dynamics of colloids embedded in a curved space is to develop experimental tools and theoretical frameworks that account for the transport properties that occur on the manifold. In the following section, we then provide the first preliminary steps to build a covariant theoretical formulation of the dynamics of an interacting colloidal system based on the many-body Langevin equation in the so-called overdamped limit [31], which allows us to deduce a Smoluchowski equation [40] for the interacting system on the manifold. We should mention that, without loss of generality, within this approach, hydrodynamic interactions are disregarded, but their explicit inclusion will be reported elsewhere. Before starting with the covariant formulation, let us introduce our notation. Let us consider the colloidal system confined on a d − dimensional manifold
As we have pointed out previously, our starting point to describe the dynamics of colloids confined in a curved manifold is based on a previous contribution [31], where the many-body Langevin stochastic equations are posed in the overdamped regime, i.e., the diffusive time scale, in local coordinates, is written as
where ζ is the friction coefficient and with
In the present exposition, we adopt the consideration that Eq. 1 is a set of N stochastic differential equations in Stratonovich’s sense [41],
where
It is a well-known fact that given a set of differential stochastic equation in the Stratonovich form, such as Eq. 2, one can find its associate Chapman–Kolmogorov differential equation for the joint probability density function
In this equation, let us note that the partial derivation
where the covariant derivative acting on a vector field vα is
Equation 4 represents the covariant formulation of the Smoluchowski equation of a colloidal system of interacting particles constrained to a curved space
Furthermore, one can write down Eq. 4 in a more compact form that allows us to prove that both systems shown in Figure 1, that is, the system of N interacting particles confined to a d-dimensional manifold
in terms of the metric tensor gαβ of the coordinates of each particle. Thus, the metric tensor associated with the line element (5) for the manifold
and the corresponding Laplace–Beltrami operator acting on scalars is simply the sum of each Laplace–Beltrami operator:
FIGURE 1. Left: Schematic representation of a set of particles embedded in a manifold
Now, defining
By expressing the Smoluchowski equation in this compact manner, it is now clear in what sense one can interpret the problem of the interacting colloidal system as the Brownian motion of a single particle in an external field
where β = 1/(ζD0) = 1/(kBT). Let us note that the expression of this partition function can also be obtained by integrating the momentum
Consequently, Eqs (4) and (8) represent the starting point of a covariant description, without the inclusion of hydrodynamic forces that allows us to study in detail the colloid dynamics in curved spaces. In the following paragraphs, we will discuss some applications of this formulation and highlight some challenges and future perspectives that can be tackled within this approach.
3 Application of the covariant approach: general behavior of the short-time dynamics in a dilute colloidal system
In this section, we study an application of the advantage of deriving the Smoluchowski equation in curved spaces in a covariant formulation (8). This consists in providing a general behavior of the joint probability density function at the short-time regime, or equivalently, in a neighborhood around a point of the manifold
where
In the following section, we use the De Witt procedure [46], that is, we first separate the points to write the term
where
Next, we take the advantage of the fact that the Smoluchowski equation is invariant under a general coordinate transformation. Thus, we choose Riemann normal coordinates (RNC) yA in a local neighborhood
In the subsequent section, we have all the pieces to split the operator (12) as
is a free “Hamiltonian” and
an interacting “Hamiltonian,” where we have defined a “momentum operator” as
where
where the terms τ(0),
Equation 16 represents the probability distribution function of the interacting particle system at the short-time regime3; it can be appreciated that the leading term,
The expectation values of the observables can be calculated using the standard definition
We are interested in the calculation of the mean-squared geodesic displacement
One can notice in Eq. 20 that in the absence of the interaction term, that is, when
One can notice that in the absence of the curvature,
For practical purposes, we will provide an explicit expression for the mean-squared displacement of a tagged particle in the interacting system. We can utilize the fact that the big metric GAB can be viewed as a block diagonal matrix, which enables us to express the Ricci curvature
It can be observed that unlike the Brownian motion of a single particle [26], where the curvature effects are solely intrinsic, in an interacting particle system where the interaction happens in Euclidean space, extrinsic curvature terms are introduced by the mean curvature K in the geodesic mean-squared displacement. As an additional observation, we should comment that the aforementioned result (21) is consistent with the standard result for flat surfaces with Rg = K = 0 [47].
In the following paragraphs, we present two illustrative examples that will highlight the importance of the geometry on the particle diffusion.
Example 1: We now provide the estimation of the order of the curvature effect for a dilute hard-sphere gas on a spherical surface. According to [40], the equation of motion for the density of hard-sphere-like Brownian particles (that might include smooth inhomogeneities) at low concentrations, ϕ, is given by the effective diffusion equation ∂ρ/∂t = D∇2ρ, where the self-diffusion coefficient of gas of hard-spheres, D gas (at low concentrations), is given by D = D0(1 + α∇ϕ), where α∇ ≃ 1.55 is a number obtained by considering hydrodynamic interactions (e.g., see chapter 6 of [40]). If one constrains the particles to diffusive exclusively on a curved surface, one expects, at least at the short-time regime, that one just needs to replace the Laplacian ∇2 by the Laplace–Beltrami Δg operator associated with the surface. Thus, the geodesic-mean squared displacement is of the form
Let us take particularly the values R/σ = 10, α∇ = 1.55 [40] and a packing fraction of ϕ ≃ 0.1; thus, one has
Example 2: We now provide an illustrative example for interacting particles confined on a sphere S2. Let us first consider a generic force Fij = f(rij)rij, where rij is the distance between the particles in the Euclidean space
Notably,
where
In addition, note that the value of the ratio σe/R depends on the strength of the interaction potential. Now, we carry out a numerical bound of the finite sum (25) as follows: note that first
Now, using the value N = 10, weak interaction
4 Concluding remarks, challenges, and perspectives
The covariant form of Smoluchowski Eq. 4 opens up the possibility of developing a theoretical framework to study different interesting phenomena that cannot be understood with the standard statistical mechanics approximations based on a Euclidean formulation. For example, one of the topics that can be tackled with this approach is the initiation of the spinodal separation of particles interacting with short-ranged attractive forces and constrained to curved space in an analogy with the procedure presented by Jan Dhont in the case of Euclidean space [40]. Following these ideas, we need to convert Eq. 4 into an expression for the probability density of one particle instead of the joint probability of all the particles. To this end, it is necessary to perform a hierarchy of equations that allows us to marginalize the joint probability density function. Once the reduced Smoluchowski equation is obtained, it is necessary to take advantage of the short-range interactions to relate out-of-equilibrium phenomena with their counterparts in equilibrium. The connection between both cases, as usual, is made through approximations concerning the equilibrium values; at this point, there exists a wide range of ways to proceed. For instance, a perturbation approach can be combined using Riemann normal coordinate formalism, Monge’s parameterization, or covariant Fourier series to calculate all the relevant observables. On the other hand, a covariant Taylor expansion [48] approach can also be performed to compare the results with their flat counterparts [40].
In addition, the covariant formalism provided by Eq. 4 can be straightforwardly used to highlight the role of the geometry on the equilibrium equation of the state of colloidal dispersions embedded in a curved space, to elucidate the geometrical contributions during the onset of non-equilibrium states, such as gels and glasses, to study the dynamics of either passive or active colloidal particles on manifolds, and to investigate the curvature effects on the structural, kinetic, and phase transitions of attractive colloids, to mention a few examples of interest in the colloidal soft matter domain. As mere speculation and motivated by the recent contribution presented in [28], the formalism here presented can also be considered to study the dynamics of granular matter in curved manifolds.
Two aspects that need to be considered toward the extension of the covariant description of the Smoluchowski equation are the generalization of the fluctuation–dissipation theorem and the inclusion of the hydrodynamic interactions in the manifold. Here, we considered that the former is satisfied in Euclidean space and that the latter are completely neglected. However, both aspects are definitely crucial to account for the dynamical properties on the manifold since they are also intimately related to the onset of non-equilibrium states. Furthermore, this theoretical framework can be extended to include rotational motion in the manifold to consider cases where the particles are anisotropic in nature, i.e., anistropic particles, such as rods, and not all the orientations are independent in the case of board-like colloids or even in those physical situations where anisotropy is associated with the particle interaction, like in patchy colloids.
Finally, we should mention that the covariant compact form of Smoluchowski Eq. 8 allowed us to obtain an expression for the joint probability density function for the full system in the short-time regime. The method implemented can be extended to capture corrections of the order of
Data availability statement
The original contributions presented in the study are included in the article/Supplementary Material; further inquiries can be directed to the corresponding author.
Author contributions
All authors listed have made a substantial, direct, and intellectual contribution to the work and approved it for publication.
Funding
Authors acknowledge the financial support from CONAHCYT (grant nos 237425, 287067, and A1-S-9098), PRODEP (grant no. 511-6/17-11852), and the University of Guanajuato (grant no. 103/2023).
Acknowledgments
The authors would like to acknowledge interesting and stimulating scientific discussions with Dr. Alejandro Villada-Balbuena and Prof. José M. Méndez-Alcaraz.
Conflict of interest
The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.
Publisher’s note
All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.
Supplementary material
The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fphy.2023.1204751/full#supplementary-material
Footnotes
1For example, the size of the particles is significantly smaller than the standard size of the membrane.
2Indeed, starting from the geodesic equation of Riemannian geometry
3Notably, one can show that ρ(ξ, 0, t) is a normalized order by order in the perturbation theory of powers of
References
1. Einstein A. Zur theorie der brownschen bewegung. Annalen der Physik (1906) 324:371–81. doi:10.1002/andp.19063240208
2. Apaza L, Sandoval M. Active matter on riemannian manifolds. Soft Matter (2018) 14:9928–36. doi:10.1039/C8SM01034J
3. Villada-Balbuena A, Ortiz-Ambriz A, Castro-Villarreal P, Tierno P, Castañeda Priego R, Méndez-Alcaraz JM. Single-file dynamics of colloids in circular channels: time scales, scaling laws and their universality. Phys Rev Res (2021) 3:033246. doi:10.1103/PhysRevResearch.3.033246
4. Kafri Y, Lubensky DK, Nelson DR. Dynamics of molecular motors with finite processivity on heterogeneous tracks. Phys Rev E (2005) 71:041906. doi:10.1103/PhysRevE.71.041906
5. Basak S, Sengupta S, Chattopadhyay K. Understanding biochemical processes in the presence of sub-diffusive behavior of biomolecules in solution and living cells. Biophysical Rev (2019) 11:851–72. doi:10.1007/s12551-019-00580-9
6. Ando T, Skolnick J. Crowding and hydrodynamic interactions likely dominate in vivo macromolecular motion. Proc Natl Acad Sci (2010) 107:18457–62. doi:10.1073/pnas.1011354107
7. Ramadurai S, Holt A, Krasnikov V, van den Bogaart G, Killian JA, Poolman B. Lateral diffusion of membrane proteins. J Am Chem Soc (2009) 131:12650–6. doi:10.1021/ja902853g
8. Alenghat FJ, Golan DE. Membrane protein dynamics and functional implications in mammalian cells. Curr Top Membranes (2013) 72:89–120. doi:10.1016/B978-0-12-417027-8.00003-9
9. Seifert U. Configurations of fluid membranes and vesicles. Adv Phys (1997) 46:13–137. doi:10.1080/00018739700101488
10. Powers TR. Dynamics of filaments and membranes in a viscous fluid. Rev Mod Phys (2010) 82:1607–31. doi:10.1103/RevModPhys.82.1607
11. Guven J, Vázquez-Montejo P. The geometry of fluid membranes: variational principles, symmetries and conservation laws. Cham: Springer International Publishing (2018). p. 167–219. doi:10.1007/978-3-319-56348-0_4
12. Tarjus G, Sausset F, Viot P. Statistical mechanics of liquids and fluids in curved space. New Jersey, United States: John Wiley & Sons, Ltd (2011). p. 251–310. doi:10.1002/9781118158715.ch4
13. Ramírez-Garza OA, Méndez-Alcaraz JM, González-Mozuelos P. Structural and dynamic inhomogeneities induced by curvature gradients in elliptic colloidal halos of paramagnetic particles. J Chem Phys (2017) 146:194903. doi:10.1063/1.4983496
14. Quintana C, González-Mozuelos P. Nanoparticles confined to a spherical surface in the presence of an external field: interaction forces and induced microstructure. J Chem Phys (2018) 148:234901. doi:10.1063/1.5014991
15. Ramírez-Garza OA, Méndez-Alcaraz JM, González-Mozuelos P. Effects of the curvature gradient on the distribution and diffusion of colloids confined to surfaces. Phys Chem Chem Phys (2021) 23:8661–72. doi:10.1039/D0CP06474B
16. Graham R. Covariant formulation of non-equilibrium statistical thermodynamics. Z Physik B Condensed Matter (1977) 26:397–405. doi:10.1007/BF01570750
17. van Kampen NG. Brownian motion on a manifold. J Stat Phys (1986) 44:1–24. doi:10.1007/BF01010902
18. Giordano S. Stochastic thermodynamics of holonomic systems. The Eur Phys J B (2019) 92:174. doi:10.1140/epjb/e2019-100162-6
19. Sarmiento-Gómez E, Villanueva-Valencia JR, Herrera-Velarde S, Ruiz-Santoyo JA, Santana-Solano J, Arauz-Lara JL, et al. Short-time dynamics of monomers and dimers in quasi-two-dimensional colloidal mixtures. Phys Rev E (2016) 94:012608. doi:10.1103/PhysRevE.94.012608
20. Villanueva-Valencia JR, Santana-Solano J, Sarmiento-Gómez E, Herrera-Velarde S, Arauz-Lara JL, Castañeda Priego R. Long-time dynamics and hydrodynamic correlations in quasi-two-dimensional anisotropic colloidal mixtures. Phys Rev E (2018) 98:062605. doi:10.1103/PhysRevE.98.062605
21. Castañeda-Priego R. Colloidal soft matter physics. Revista Mexicana de Física (2021) 67:050101. doi:10.31349/RevMexFis.67.0501011
22. Hołyst R, Plewczyński D, Aksimentiev A, Burdzy K. Diffusion on curved, periodic surfaces. Phys Rev E (1999) 60:302–7. doi:10.1103/PhysRevE.60.302
23. Faraudo J. Diffusion equation on curved surfaces. i. theory and application to biological membranes. J Chem Phys (2002) 116:5831–41. doi:10.1063/1.1456024
24. Zhong Y, Zhao L, Tyrlik PM, Wang G. Investigating diffusing on highly curved water–oil interface using three-dimensional single particle tracking. The J Phys Chem C (2017) 121:8023–32. doi:10.1021/acs.jpcc.7b01721
25. Castro-Villarreal P. Intrinsic and extrinsic measurement for brownian motion. J Stat Mech Theor Exp (2014) 2014:P05017. doi:10.1088/1742-5468/2014/05/P05017
26. Castro-Villarreal P. Brownian motion meets riemann curvature. J Stat Mech Theor Exp (2010) 2010:P08006. doi:10.1088/1742-5468/2010/08/P08006
27. Castro-Villarreal P, Sevilla FJ. Active motion on curved surfaces. Phys Rev E (2018) 97:052605. doi:10.1103/PhysRevE.97.052605
28. Ledesma-Motolinía M, Carrillo-Estrada JL, Escobar A, Donado F, Castro-Villarreal P. Magnetized granular particles running and tumbling on the circle S1. Phys Rev E (2023) 107:024902. doi:10.1103/PhysRevE.107.024902
29. Narinder N, Gomez-Solano JR, Bechinger C. Active particles in geometrically confined viscoelastic fluids. New J Phys (2019) 21:093058. doi:10.1088/1367-2630/ab40e0
30. Katok A, Hasselblatt B. Introduction to the modern theory of dynamical systems. In: Encyclopedia of mathematics and its applications. Cambridge: Cambridge University Press (1995). doi:10.1017/CBO9780511809187
31. Castro-Villarreal P, Villada-Balbuena A, Méndez-Alcaraz JM, Castañeda-Priego R, Estrada-Jiménez S. A brownian dynamics algorithm for colloids in curved manifolds. J Chem Phys (2014) 140:214115. doi:10.1063/1.4881060
32. Kolmogoroff A. Zur umkehrbarkeit der statistischen naturgesetze. Mathematische Annalen (1937) 113:766–72. doi:10.1007/bf01571664
33. Yosida K. Integration of Fokker-Planck’s equation in a compact Riemannian space. Arkiv för Matematik (1949) 1:71–5. doi:10.1007/BF02590470
34. Itô K. Stochastic differential equations in a differentiable manifold. Nagoya Math J (1950) 1:35–47. doi:10.1017/S0027763000022819
35. Molchanov SA. Diffusion processes and riemannian geometry. Russ Math Surv (1975) 30:1–63. doi:10.1070/RM1975v030n01ABEH001400
36. Stratonovich RL. Auxiliary information concerning probability theory and equilibrium thermodynamics. Berlin, Heidelberg: Springer Berlin Heidelberg (1992). p. 25–58. doi:10.1007/978-3-642-77343-3_2
37. Risken H. Fokker-planck equation. Berlin, Heidelberg: Springer Berlin Heidelberg (1996). p. 63–95. doi:10.1007/978-3-642-61544-3_4
38. Polettini M. Generally covariant state-dependent diffusion. J Stat Mech Theor Exp (2013) 2013:P07005. doi:10.1088/1742-5468/2013/07/P07005
39. Ding M, Tu Z, Xing X. Covariant formulation of nonlinear Langevin theory with multiplicative Gaussian white noises. Phys Rev Res (2020) 2:033381. doi:10.1103/PhysRevResearch.2.033381
40. Dhont JK. An introduction to dynamics of colloids. Amsterdam, Netherlands: ISSN Elsevier Science (1996).
41. Gardiner C. Stochastic methods: a handbook for the natural and social Sciences. Berlin, Heidelberg: Springer Berlin Heidelberg (2009).
42. Nakahara M. Geometry, topology, and physics. In: Graduate student series in physics. 2nd ed. Bristol, United Kingdom: Institute of Physics Publishing (2003).
43. Castro-Villarreal P, Contreras-Aburto C, Estrada-Jiménez S, Huet-Hernández I, Vázquez-Rodríguez O. Single file diffusion meets feynman path integral. J Stat Mech Theor Exp (2021) 2021:093208. doi:10.1088/1742-5468/ac21d8
44. Eisenhart L. Riemannian geometry. In: Princeton landmarks in mathematics and physics. New Jersey, United States: Princeton University Press (2016).
45. Hatzinikitas A. A note on riemann normal coordinates (2000)arXiv:hep-th/0001078.
46. DeWitt BS. Dynamical theory of groups and fields. Documents on modern physics. Pennsylvania, United States: Gordon & Breach (1965).
47. Pusey PN, Tough RJA. Langevin approach to the dynamics of interacting brownian particles. J Phys A: Math Gen (1982) 15:1291–308. doi:10.1088/0305-4470/15/4/030
Keywords: diffusion, Brownian motion, colloids, Smoluchowski equation, curved manifold
Citation: Castro-Villarreal P, Solano-Cabrera CO and Castañeda-Priego R (2023) Covariant description of the colloidal dynamics on curved manifolds. Front. Phys. 11:1204751. doi: 10.3389/fphy.2023.1204751
Received: 12 April 2023; Accepted: 10 October 2023;
Published: 15 November 2023.
Edited by:
Jasper Van Der Gucht, Wageningen University and Research, NetherlandsReviewed by:
Jan Karel George Dhont, Helmholtz Association of German Research Centres (HZ), GermanyAdriano Valdes-Gomez, National Autonomous University of Mexico, Mexico
Copyright © 2023 Castro-Villarreal, Solano-Cabrera and Castañeda-Priego. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.
*Correspondence: Pavel Castro-Villarreal, cGNhc3Ryb3ZAdW5hY2gubXg=; Ramón Castañeda-Priego, cmFtb25jcEBmaXNpY2EudWd0by5teA==