Skip to main content
Log in

A quasi-continuum multi-scale theory for self-diffusion and fluid ordering in nanochannel flows

  • Research Paper
  • Published:
Microfluidics and Nanofluidics Aims and scope Submit manuscript

Abstract

We present a quasi-continuum self-diffusion theory that can capture the ordering effects and the density variations that are predicted by non-equilibrium molecular dynamics (NEMD) in nanochannel flows. A number of properties that affect fluid ordering in NEMD simulations are extracted and compared with the quasi-continuum predictions. The proposed diffusion equation requires the classic diffusion coefficient D and a micro structural internal length g that relates directly to the shape of the molecular potential of the NEMD calculations. The quasi-continuum self-diffusion theory comes as an alternative to atomistic simulation, bridging the gap between continuum and atomistic behavior with classical hydrodynamic relations and reduces the computational burden as compared with fully atomistic simulations.

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

Similar content being viewed by others

Notes

  1. If a gravitational field \(\phi\) is present, we replace \(\mu\) with \(\mu + m\phi\) (\(m\) is the mass per mole). For \(\phi = {\text{const}}\), the present results will still hold true.

Abbreviations

A 1–4 :

Constants determined by BC

A, B :

Constants for inhomogeneous diffusion

A i , B i :

Airy functions

c :

Concentration

c 1–3 :

Real constants

D :

Bulk diffusion coefficient

D ap :

Apparent diffusion coefficient

F :

Diffusion functional

F ap :

Apparent diffusion functional

1 F 1 :

Hypergeometric function

F ext :

Magnitude of external driving force

g :

Wavelength

G :

Gibbs free energy

G 0 :

Gibbs free energy at equilibrium

h :

Boundary value for concentration

H:

Hermitian polyomial

h ch :

Channel height

J :

Diffusional flux

K :

Gradient energy coefficient

K * :

Spring constant

L x :

Length of the computational domain in the x-direction

L y :

Length of the computational domain in the y-direction

L z :

Length of the computational domain in the z-direction

M :

Diffusional mobility

n :

Integer number, n = 0, 1, 2…

p :

Pressure

q :

Boundary value for the non-classic flux term

r eq :

Position of a wall atom on fcc lattice site

r i :

Position vector of atom i

r ij :

Distance vector between ith and jth atom

T :

Temperature

S :

Area

u(r ij ):

LJ potential of atom i with atom j

V :

Volume

w :

Boundary condition for the normal flux

z * :

Normalized distances in the z-direction

ε :

Energy parameter in the LJ potential

μ :

Local chemical potential

ρ :

Fluid density

σ :

Length parameter in the LJ potential

References

  • Allen MP, Tildesley TJ (1987) Computer simulation of liquids. Clarendon Press, Oxford

    MATH  Google Scholar 

  • Asproulis N, Drikakis D (2011) Wall-mass effects on hydrodynamic boundary slip. Phys Rev E 84:031504

    Article  Google Scholar 

  • Binder K, Horbach J, Kob W, Paul W, Varnik F (2004) Molecular dynamics simulations. J Phys: Condens Matter 16:429–453

    Google Scholar 

  • Cahn JW (1961) On spinodal decomposition. Act Meta 9:765–801

    Google Scholar 

  • Cahn JW, Hilliard JE (1958) Free energy of a non uniform system. I. Interfacial free energy. J Chem Phys 28:258–267

    Article  Google Scholar 

  • Cahn JW, Hilliard JE (1959) Free energy of a non uniform system. III. Nucleation in a two-component incompressible fluid. J Chem Phys 31:688–699

    Article  Google Scholar 

  • Cook HE (1970) Brownian motion in spinodal decomposition. Act Meta 18:297–306

    Article  Google Scholar 

  • Courant R, Hilbert D (1953) Methods of mathematical physics, vol I. Wiley Interscience, New York

    Google Scholar 

  • Crank J (1975) The mathematics of diffusion. Oxford University Press, New York

    Google Scholar 

  • Darken LS (1948) Diffusion, mobility and their interrelation through free energy in binary metallic systems. Trans AIME 175:184–201

    Google Scholar 

  • Evans DJ, Holian BL (1985) The Nosé–Hoover thermostat. J Chem Phys 83:4069–4074

    Article  Google Scholar 

  • Fouladi F, Hosseinzadeh E, Barari A, Dmairry G (2010) Highly nonlinear temperature-dependent fin: analysis by variational iteration method. Heat Tranfer Res 41:1–11

    Article  Google Scholar 

  • Gini-Castagnoli G, Ricci FP (1960) Self-diffusion in liquid argon. J Chem Phys 32:19–20

    Article  Google Scholar 

  • Hartkamp R, Ghosh A, Weinhart T, Luding S (2012) A study of the anisotropy of stress in a fluid confined in a nanochannel. J Chem Phys 137:044711

    Article  Google Scholar 

  • Holian BL, Voter AF (1995) Thermostatted molecular dynamics: How to avoid the Toda demon hidden in Nosé–Hoover dynamics. Phys Rev E 52:2338–2347

    Article  Google Scholar 

  • Iwai Y, Uchida H, Koga Y, Arai Y, Movi Y (1996) Monte Carlo simulation of solubilities of aromatic compounds in supercritical carbon dioxide by a group contribution site model. Ind Eng Chem Res 35:3782–3787

    Article  Google Scholar 

  • Li Q, Liu C (2012) Molecular dynamics simulation of heat transfer with effects of fluid–lattice interactions. Int J Heat Mass Trans 55:8088–8092

    Article  Google Scholar 

  • Lipowsky R, Fisher ME (1986) Wetting in random systems. Phys Rev Lett 56:472–475

    Article  MathSciNet  Google Scholar 

  • Lipowsky R, Huse DA (1986) Diffusion-limited growth of wetting layers. Phys Rev Lett 57:353–356

    Article  Google Scholar 

  • Marko JF (1993) Influence of surface interactions on spinodal decomposition. Phys Rev E 48:2861–2879

    Article  Google Scholar 

  • Nagayama G, Cheng P (2004) Effects of interface wettability on microscale flow by molecular dynamics simulation. Int J Heat Mass Trans 47:501–513

    Article  MATH  Google Scholar 

  • Priezjev NV (2007) Effect of surface roughness on rate-dependent slip in simple fluids. J Chem Phys 127:144708

    Article  Google Scholar 

  • Prigogine I, Glansdorff P (1965) Variational properties and fluctuation theory. Physica 31:1242–1256

    Article  MathSciNet  Google Scholar 

  • Sofos F, Karakasidis TE, Liakopoulos A (2009a) Non-equilibrium molecular dynamics investigation of parameters affecting planar nanochannel flows. Cont Eng Sci 2:283–298

    Google Scholar 

  • Sofos F, Karakasidis TE, Liakopoulos A (2009b) Transport properties of liquid argon in krypton nanochannels: anisotropy and non-homogeneity introduced by the solid walls. Int J Heat Mass Trans 52:735–743

    Article  MATH  Google Scholar 

  • Sofos F, Karakasidis TE, Liakopoulos A (2012) Surface wettability effects on flow in rough wall nanochannels. Microfluid Nanofluid 12:25–31

    Article  Google Scholar 

  • Somers SA, Davis HT (1991) Microscopic dynamics of fluids confined between smooth and atomically structured solid surfaces. J Chem Phys 96:5389–5407

    Article  Google Scholar 

  • Stanley HE (1971) Introduction to phase transitions and critical phenomena. Clarendon, Oxford

    Google Scholar 

  • Thomas JA, McGaughey AJH (2007) Effect of surface wettability on liquid density, structure, and diffusion near a solid surface. J Chem Phys 126:034707

    Article  Google Scholar 

  • Voronov RS, Papavassiliou DV, Lee LL (2006) Boundary slip and wetting properties of interfaces: correlation of the contact angle with the slip length. J Chem Phys 124:204701

    Article  Google Scholar 

Download references

Acknowledgments

This project was implemented under the “ARISTEIA II” Action of the “OPERATIONAL PROGRAMME EDUCATION AND LIFELONG LEARNING” and is co-founded by the European Social Fund (ESF) and National Resources.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Theodoros E. Karakasidis.

Appendices

Appendix 1: ODE solution for the 1D state problem

The general solution of Eq. (30) is

$$c(z) = A_{1} + A_{2} z + A_{3} \sin \frac{z}{g} + A_{4} \cos \frac{z}{g}$$
(46)

where A 1, …A 4 are constants to be determined by BC. The first two terms are similar to the classic solution. The other two terms are the non-classic part of the solution. Taking the region 0 ≤ z ≤ h ch, (h ch is the channel width), we impose the following BCs that are compatible with the Poiseuille flow in the normal to z-direction.

$$J = - D\left( {\frac{{{\text{d}}c}}{{{\text{d}}z}} + g^{2} \frac{{{\text{d}}^{3} c}}{{{\text{d}}z^{3} }}} \right)$$
(47)

At z = 0 and h ch, J = 0, that is no flux at the walls, which gives

$$A_{2} = 0$$
(48)

At z = h ch/2, c = c o a datum point at the middle of the channel (can be thought of as a constant that increases with temperature), which gives

$$A_{1} = c_{o} - A_{3} \sin \frac{{h_{\text{ch}} }}{2g} - A_{4} \cos \frac{{h_{\text{ch}} }}{2g},c_{o} \ge 0$$
(49)

Symmetry of the concentration implies dc/dz = 0 at z = h ch/2, which gives

$$A_{3} \cos \frac{{h_{\text{ch}} }}{2g} = A_{4} \sin \frac{{h_{\text{ch}} }}{2g}$$
(50)

This condition implies that J = 0 at z = h ch, as expected. It can be shown that J = 0 for all z, 0 ≤ z ≤ h ch.

Mass conservation implies

$$\int\limits_{0}^{{h_{\text{ch}} /2}} {c(z){\text{d}}z} = \bar{c}\frac{{h_{\text{ch}} }}{2},\quad\bar{c} > 0$$
(51)

which gives

$$A_{1} + A_{4} \frac{2g}{{h_{\text{ch}} }}\sin \frac{{h_{\text{ch}} }}{2g} + A_{3} \frac{2g}{{h_{\text{ch}} }}\left( {1 - \cos \frac{{h_{\text{ch}} }}{2g}} \right) = \bar{c}$$
(52)

Solving Eqs. (49), (50) and (52), we obtain

$$A_{4} = (c_{o} - A_{1} )\cos \frac{{h_{\text{ch}} }}{2g}$$
(53)
$$A_{3} = (c_{o} - A_{1} )\sin \frac{{h_{\text{ch}} }}{2g}$$
(54)
$$A_{1} = c_{o} \frac{{\frac{{\bar{c}}}{{c_{o} }} - \frac{2g}{{h_{\text{ch}} }}\sin \frac{{h_{\text{ch}} }}{2g}}}{{1 - \frac{2g}{{h_{\text{ch}} }}\sin \frac{{h_{\text{ch}} }}{2g}}}$$
(55)

For \(\bar{c} = c_{o}\), we obtain the classic relation c = c o .

Appendix 2: Asymptotic analysis

The solution of Eq. (37) can be written as combinations of hypergeometric function 1 F 1 and the Hermitian polynomial H:

$$\frac{{{\text{d}}^{2} c}}{{{\text{d}}z^{2} }} = e^{{ - \frac{{Az^{2} }}{{2g_{0} }}}} \left\{ {c_{1} H\left( {\frac{{1 - Ag_{o} }}{{2Ag_{o} }},\frac{\sqrt A z}{{\sqrt {g_{o} } }}} \right) + c_{2} \,\,{}_{1}F_{1} \left( {\frac{{ - 1 + Ag_{o} }}{{4Ag_{o} }};\frac{1}{2};\frac{{Az^{2} }}{{g_{o} }}} \right)} \right\}$$
(56)

where c 1 and c 2 are real constants. Note that only the hypergeometric function 1 F 1 is symmetric with respect to z. Therefore, symmetric cases are accepted only if c 1 = 0. An interesting expansion of the solution around z = 0 (middle of the channel) is given below. In some cases in the field of heat transfer, the variational iteration method is utilized as an approximate analytical method to overcome some inherent limitations arising as uncontrollability to the nonzero endpoint boundary conditions (see, for example, Fouladi et al. 2010).

It is of interest to expand the symmetric solution (c 1 = 0) around point z = 0 (the middle of the channel width).

$$\begin{aligned} \frac{{{\text{d}}^{2} c}}{{{\text{d}}z^{2} }} & = e^{{ - \frac{{Az^{2} }}{{2g_{o} }}}} c_{2} {}_{1}F_{1} \left( {\frac{{ - 1 + Ag_{o} }}{{24g_{o} }};\frac{ 1}{ 2};\frac{{Az^{2} }}{{g_{o} }}} \right) \\ & \quad \approx c_{2} \left\{ {1 - \frac{{z^{2} }}{{2g_{o}^{2} }} + \left( {\frac{1}{{24g_{o}^{4} }} + \frac{{A^{2} }}{{12g_{o}^{2} }}} \right)z^{4} - \left( {\frac{1}{{720g_{o}^{6} }} + \frac{{7A^{2} }}{{360g_{o}^{2} }}} \right)z^{6} + 0\left( {z^{8} } \right)} \right\} \\ \end{aligned}$$
(57)

Note that the expansion (57) involves only even powers of z. Taking the conditions dc/dz = 0 and c = c o at z = 0, and retaining the first two terms of (57), we obtain

$$\frac{{{\text{d}}c}}{{{\text{d}}x}} \approx c_{2} \left( {z - \frac{{z^{3} }}{{6g_{o}^{2} }}} \right) + 0\left( {z^{5} } \right)$$
(58)

and

$$c \approx c_{o} + c_{2} \left( {\frac{{z^{2} }}{2} - \frac{{z^{4} }}{{24g_{o}^{2} }}} \right) + 0\left( {z^{6} } \right)$$
(59)

Assuming that (59) should provide c ≥ 0 for all values of c o  ≥ 0, we conclude that c 2 ≥ 0. In other words, dc 2/dz 2 ≥ 0 at z = 0, a result that is verified by the NEMD calculations!

Appendix 3: Diffusion for asymmetric wall conditions

The case of asymmetric wall conditions can be studied by a diffusion inhomogeneous distribution of the type

$$D = D_{o} (1 - Bz/g_{o} )$$
(60)

The diffusion equation becomes

$$\frac{{d^{4} c}}{{{\text{d}}z^{4} }} + \frac{{(1 - Bz/g_{o} )}}{{g_{o}^{2} }}\frac{{d^{2} c}}{{{\text{d}}z^{2} }} = 0$$
(61)

The solution of Eq. (61) can be put in closed form.

$$\frac{{d^{2} c}}{{{\text{d}}z^{2} }} = c_{2} A_{i} \left[ {\frac{{ - 1 + Bz/g_{o} }}{{B^{2/3} }}} \right] + c_{3} B_{i} \left[ {\frac{{ - 1 + Bz/g_{o} }}{{B^{2/3} }}} \right]$$
(62)

where A i and B i are Airy functions and c 2, c 3 are constants. The full solution of Eq. (62) is given by

$$c(z) = c_{o} + c_{1} z + c_{2} C_{\rm A} (z) + c_{3} C_{\rm B} (z)$$
(63)

where c o , c 1, c 2, c 3 are constants.

The general form of the functions C A(z) and C B(z)are

$$\begin{aligned} C_{A} (z) & = \frac{1}{{3^{11/3} \varGamma \left( \frac{4}{3} \right)\varGamma \left( \frac{5}{3} \right)}}\left\{ {3\left( { - \frac{{g_{o} }}{B} + z} \right)^{2} \varGamma \left( \frac{1}{3} \right){}_{p}F_{q} \left[ {\left\{ \frac{1}{3} \right\},\left\{ {\frac{4}{3},\frac{5}{3}} \right\},\frac{{B\left( { - \frac{{g_{o} }}{B} + z} \right)^{2} }}{{9g_{o}^{2} }}} \right]} \right. \\ & \quad - \frac{1}{B}\left[ {3^{1/3} \left( {\frac{B}{{g_{o}^{3} }}} \right)^{1/3} \varGamma \left( \frac{2}{3} \right)} \right.\left[ {6{\text{g}}_{\text{o}}^{ 2} - 3^{1/3} 2g_{\text{o}}^{ 2} \left( {\frac{{\sqrt B \left( { - \frac{2}{B} + z} \right)^{3/2} }}{{g_{o}^{3/2} }}} \right)^{2/3} I\left[ { - \frac{2}{3},\frac{{2\sqrt B \left( { - \frac{{g_{o} }}{B} + z} \right)^{3/2} }}{{3g_{o}^{3/2} }}} \right]\varGamma \left( \frac{1}{3} \right)} \right. \\ & \quad + 3B\left( { - \frac{{g_{o} }}{B} + z} \right)^{3} {}_{p}F_{q} \left. {\left. {\left. {\left[ {\left\{ \frac{2}{3} \right\},\left\{ {\frac{4}{3},\frac{5}{3}} \right\},\frac{{B\left( { - \frac{{g_{o} }}{B} + z} \right)^{2} }}{{9g_{o}^{2} }}} \right]} \right]} \right]} \right\} \\ \end{aligned}$$
(64)
$$\begin{aligned} C_{B} (z) & = \frac{1}{{3^{17/6} \varGamma \left( \frac{4}{3} \right)\varGamma \left( \frac{5}{3} \right)}}\left\{ {3^{2/3} \left( { - \frac{{g_{o} }}{B} + z} \right)^{2} \varGamma \left( \frac{1}{3} \right){}_{p}F_{q} \left[ {\left\{ \frac{1}{3} \right\},\left\{ {\frac{4}{3},\frac{5}{3}} \right\},\frac{{B\left( { - \frac{{g_{o} }}{B} + z} \right)^{2} }}{{9g_{o}^{3} }}} \right]} \right. \\ & \quad + \frac{1}{B}\left[ {\left( {\frac{B}{{g_{o}^{3} }}} \right)^{1/3} \varGamma \left( \frac{2}{3} \right)} \right.\left[ {6{\text{g}}_{o}^{2} - 3^{1/3} 2g_{o}^{2} \left( {\frac{{\sqrt B \left( { - \frac{{g_{o} }}{B} + z} \right)^{3/2} }}{{g_{o}^{3/2} }}} \right)^{2/3} I\left[ { - \frac{2}{3},\frac{{2\sqrt B \left( { - \frac{{g_{o} }}{B} + z} \right)^{3/2} }}{{3g_{o}^{3/2} }}} \right]\varGamma \left( \frac{1}{3} \right)} \right. \\ & \quad + 3B\left( { - \frac{{g_{o} }}{B} + z} \right)^{3} {}_{p}F_{q} \left. {\left. {\left. {\left[ {\left\{ \frac{2}{3} \right\},\left\{ {\frac{4}{3},\frac{5}{3}} \right\},\frac{{B\left( { - \frac{{g_{o} }}{B} + z} \right)^{3} }}{{9g_{o}^{3} }}} \right]} \right]} \right]} \right\} \\ \end{aligned}$$
(65)

p F q is the generalized hypergeometric function, I is the modified Bessel function of the first kind, Γ is the Gamma function, and B is a parameter of the problem.

We believe that the chemical affinity between the walls and the fluid, as well as the roughness of the walls, influence the diffusion coefficient distribution and this, in turn, affects the steady state concentration profiles.

Appendix 4: Temperature profile

Temperature is calculated in each bin across the channel using

$$T_{\text{bin}} = \frac{{m_{Ar} }}{{3N_{\text{bin}} k_{B} }}\sum\limits_{n - 1}^{{N_{\text{bin}} }} {\sum\limits_{i - 1}^{3} {\left( {v_{n ,i} - {\bar{\text{v}}}_{i} } \right)^{2} } }$$
(66)

where N bin is the number of fluid atoms in the bin examined, i = 1, 2, 3, … denotes the xyz component of the atomic velocity \(v_{n - 1}\), \({\bar{\text{v}}}_{\text{i}}\) is the ith component of the mean macroflow velocity, k B is the Boltzman constant, and m Ar is argon atom mass.

In Fig. 6a, the calculated fluid temperature distribution for h ch = 6.3 nm is presented. We can see that for each system temperature (the temperature value imposed at the wall’s thermostats, 100, 120, or 150 K), the temperature profile is approximately flat at this temperature value. We obtain similar flat behavior when we vary system parameters such as channel width h ch, the wall atoms spring constant K *, the wall–fluid interaction ratio ɛ wall /ɛ fluid an the average fluid density ρ *. However, in Fig. 6b, for h ch = 17.1 nm and external forces above 0.036 pN, temperature profiles are not uniform and show a rise in the middle of the channel and near the solid walls. Similar behavior is reported on Nagayama and Cheng (2004) when the external force that drives the flow becomes large. We attribute this behavior to the fact that at large channel widths, in combination with external forces of large magnitude, strain rates inside the nanochannel increase, and the system has non-linear response.

Fig. 6
figure 6

Temperature profiles a for system temperature T = 100, 120 and 150 K, and h ch = 6.3 nm b for various magnitudes of the F ext (normalized), T = 120 K and \(h_{\text{ch}} = 17.1{\text{ nm}}\)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Giannakopoulos, A.E., Sofos, F., Karakasidis, T.E. et al. A quasi-continuum multi-scale theory for self-diffusion and fluid ordering in nanochannel flows. Microfluid Nanofluid 17, 1011–1023 (2014). https://doi.org/10.1007/s10404-014-1390-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10404-014-1390-2

Keywords

Navigation