Chapter 2. Fundamentals of mass transport in the microscale

Chapter Outline

2.1. Transport Phenomena10
2.1.1. Molecular level10
2.1.2. Continuum level14
2.1.2.1. Conservation of mass14
2.1.2.2. Conservation of momentum15
2.1.2.3. Conservation of energy21
2.1.2.4. Conservation of species22
2.2. Molecular Diffusion22
2.2.1. Random walk and Brownian motion22
2.2.2. Stokes–Einstein model of diffusion23
2.2.3. Diffusion coefficient24
2.2.3.1. Diffusion coefficient in gases24
2.2.3.2. Diffusion coefficient in liquids25
2.2.3.3. Diffusion coefficient of electrolytes26
2.3. Taylor Dispersion28
2.3.1. Two-dimensional analysis28
2.3.2. Three-dimensional analysis34
2.4. Chaotic Advection36
2.4.1. Basic terminologies36
2.4.2. Examples of chaotic advection40
2.4.2.1. Lorentz's convection flow40
2.4.2.2. Dean flow in curved pipes40
2.4.2.3. Flow in helical pipes46
2.4.2.4. Flow in twisted pipes47
2.4.2.5. Flow in a droplet50
2.5. Viscoelastic effects53
2.6. Electrokinetic Effects55
2.6.1. Electroosmosis55
2.6.1.1. The Debye layer55
2.6.1.2. Electroosmotic transport effect57
2.6.1.3. Electrokinetic flow between two parallel plates58
2.6.1.4. Electrokinetic flow in a cylindrical capillary60
2.6.1.5. Electrokinetic flow in a rectangular microchannel61
2.6.1.6. Ohmic model for electrolyte solutions63
2.6.2. Electrophoresis64
2.6.3. Dielectrophoresis65
2.7. Magnetic and Electromagnetic Effects66
2.7.1. Magnetic effects66
2.7.1.1. Electromagnetic effects68
2.8. Scaling Law and Fluid Flow in Microscale68
References71
Transport phenomena in micromixers can be described theoretically at two basic levels: molecular level and continuum level. The two different levels of description correspond to the typical length scale involved. Continuum model can describe most transport phenomena in micromixers with a length scale ranging from micrometers to centimeters. Most micromixers for practical applications are in this range of length scale. Molecular models involve transport phenomena in the range of one nanometer to one micrometer. Mixers with length scale in this range should be called “nanomixer.” The term “micromixer” in this book will cover devices with submillimeter length dimension.

2.1. Transport Phenomena

Transport phenomena in micromixers can be described theoretically at two basic levels: molecular level and continuum level. The two different levels of description correspond to the typical length scale involved. Continuum model can describe most transport phenomena in micromixers with a length scale ranging from micrometers to centimeters. Most micromixers for practical applications are in this range of length scale. Molecular models involve transport phenomena in the range of one nanometer to one micrometer. Mixers with length scale in this range should be called “nanomixer.” The term “micromixer” in this book will cover devices with submillimeter length dimension.
At the continuum level, the fluid is considered as a continuum. Fluid properties are defined continuously throughout the space. At this level, fluid properties, such as viscosity, density, and conductivity, are considered as material properties. Transport phenomena can be described by a set of conservation equations for mass, momentum, and energy. These equations of change are partial differential equations, which can be solved for physical fields in a micromixer, such as concentration, velocity, and temperature.
Miniaturization technologies have pushed the length scale of microdevices further. In the event of nanotechnology, scientists and engineers will encounter more phenomena at molecular level. At this level, transport phenomena can be described through molecular structure and intermolecular forces. Because many micromixers are used as microreactors, a fundamental understanding of molecular processes is important for designing devices with length scale in the micrometer to centimeter range.

2.1.1. Molecular level

At molecular level, the simplest description of transport phenomena is based on the kinetic theory of diluted monatomic gases, which is also called the Chapman–Enskog theory. The interaction between nonpolar molecules is represented by the Lennard–Jones potential, which has an empirical form of:
(2.1)
B9781437735208000024/si7.gif is missing
where σ is the characteristic diameter of the molecule, r is the distance between the two molecules, and ɛ is the characteristic energy, which is the maximum energy of attraction between the molecules. In (2.1), the term (σ/r)12 represents the repulsion potential, while the term (σ/r)6 represents the attraction potential between the pair of molecules. The coefficients cij and dij are determined by molecule types and often assumed to be 1. Table 2.1 lists the parameters of some common gases. With the Lennard–Jones potential, the force between the molecules can be derived as:
Table 2.1 Lennard–Jones characteristic energies and characteristic diameters of common gases [1]
Boltzmann constant: kB=1.38×10−23J/K, dij=cij=1.
GasCharacteristic Energy (ɛ/kB)Characteristic Diameter σ (nm)
Air97.00.362
N291.50.368
CO2190.00.400
O2113.00.343
Ar124.00.342
(2.2)
B9781437735208000024/si8.gif is missing
The Lennard–Jones model results in the characteristic time:
(2.3)
B9781437735208000024/si9.gif is missing
where M is the molecular mass. This characteristic time corresponds to the oscillation period between repulsion and attraction. Furthermore, the model allows the determination of the dynamic viscosity of a pure monatomic gas [2]:
(2.4)
B9781437735208000024/si10.gif is missing
where the collision integral Ω is a function of the dimensionless temperature kBT/ɛ describing the deviation from rigid sphere behavior, with kB being the Boltzmann constant. Fig. 2.1 depicts the function of Ω. The value of the collision integral Ω is on the order of 1. The above equation allows the determination of Lennard–Jones parameters σ and ɛ from the measurement of viscosity μ, a macroscopic continuum property.
B9781437735208000024/f02-01-9781437735208.jpg is missing
FIGURE 2.1
Collision integral Ω as a function of dimensionless temperature kBT/ɛ.
Example 2.1 (Estimation of gas viscosity using kinetic theory). Estimate the viscosity of pure nitrogen at 25 °C.
Solution. Using the Lennard–Jones parameters of nitrogen listed in Table 2.1, the dimensionless temperature is:
B9781437735208000024/si11.gif is missing
According to the diagram in Fig. 2.1, the collision integral of N2 at 25°C is 0.95. The estimated viscosity is then:
B9781437735208000024/si12.gif is missing
The most important characteristic length scale in gas dynamics is the mean free path, which is the average distance traveled by a molecule between successive collisions:
(2.5)
B9781437735208000024/si13.gif is missing
where n is the number density (number of molecules per unit volume):
(2.6)
B9781437735208000024/si14.gif is missing
The ratio between the mean free path and the characteristic length of the device, for instance, the channel diameter, is called the Knudsen number:
(2.7)
B9781437735208000024/si15.gif is missing
Because the Knudsen number represents a link between the length scale of a device and the interaction between fluid molecules, it can be used to estimate the right model for describing the transport phenomena. For Kn<10−3, the fluid is a continuum. For 10−3<Kn<10−1, a continuum model with modified boundary conditions is appropriate. For Kn>10, the fluid can only be described by a free molecular flow model.
Kinetic theory can be applied to liquids as well. In this model, the motion of liquid molecules is confined within a space limited by its neighboring molecules. Based on this theory, the viscosity of a liquid can be estimated as:
(2.8)
B9781437735208000024/si16.gif is missing
where NA=6.023×1023 is the Avogadro number or the number of molecules per mole; ħ=6.626068×10−34m2kg/s is the Planck constant; B9781437735208000024/si17.gif is missing is the molar volume; Tb is the boiling temperature; and T is the temperature of the liquid.
The models of viscosity for gas (2.4) and for liquid (2.8) show opposite temperature dependency. While the viscosity of gases increases with higher temperature, the viscosity of liquids decreases.
Example 2.2 (Dynamic viscosity of water). If the molar volume of water at 25°C is 18×10−6m3/mol, determine the viscosity of water at this temperature.
Solution. The boiling temperature of water under atmospheric pressure is assumed to be 100°C. According to (2.8), the viscosity of water can be estimated as:
B9781437735208000024/si18.gif is missing
The equation overestimates the viscosity of water.
In the previous discussion, continuum properties are derived from the molecular model using statistic methods. If there are not enough molecules for good statistics, numerical tools are used for modeling transport phenomena at the molecular level. There are two numerical methods: molecular dynamics (MD) and direct simulation Monte Carlo (DSMC). While MD is a deterministic method, DSMC is a statistical method.
Molecular dynamics is a numerical method for modeling the motion of single molecules. The interactions between the molecules can be described by the classical second law of Newton. The simplest model of a molecule is a hard sphere of a mass m. The binary interaction between two molecules is determined by the Lennard–Jones force (2.2):
(2.9)
B9781437735208000024/si19.gif is missing
where r=|rirj| is the distance between the molecules. The bold letter indicates a vector variable. The dynamics of molecule i can be described by Newton’s second law:
(2.10)
B9781437735208000024/si20.gif is missing
where N is the total number of molecules in the modeled system. The basic steps of an MD-simulation are:
• Determination of initial conditions and geometry parameters,
• Determination of interaction force (2.2), and
• Integration of equation of motion (2.10).
Because of its deterministic nature, MD is extremely expensive in terms of computational resources. Less resources would be needed if the system is modeled with a statistic method.
Direct simulation Monte Carlo is a statistic method for modeling at the molecular level. In DSMC, many molecules are modeled as a single particle. The interactions between the molecules of each particle are determined statistically, while the motion of the particle is modeled deterministically. The basic steps of DSMC are [3]:
• Determination of particle motion,
• Indication and cross-referencing of particles,
• Simulation of particle collision, and
• Sampling of macroscopic properties.

2.1.2. Continuum level

At continuum level, transport phenomena are described with a set of conservation equation. Because flow in micromixers is laminar, we do not need to deal with turbulent flow, which is impossible to solve analytically. The three basic conservation equations are:
• Conservation of mass: continuity equation,
• Conservation of momentum: Newton’s second law or Navier–Stokes equation, and
• Conservation of energy: first law of thermodynamics or energy equation.
Solving these three equations will result in three basic variables: the velocity field v, the pressure field p, and the temperature field T. Fluid properties, depending on the thermodynamic state (pressure p and temperature T), such as density, viscosity, thermal conductivity, and enthalpy, can be derived from these variables and fed back into the conservation equations. The above three equations are formulated for a single phase of homogenous composition. In micromixers, most fluids carry one or more species other than the carrying fluids. If the mixers are used as microreactors, chemical reactions also need to be considered. Thus, in addition to the above three conservation equations, two further equations are needed to describe the transport of species in micromixers:
• Conservation of species: convective/diffusive equation and
• Laws of chemical reactions.

2.1.2.1. Conservation of mass

The continuity equation has the general form:
(2.11)
B9781437735208000024/si21.gif is missing
where B9781437735208000024/si22.gif is missing is the nabla operator and B9781437735208000024/si23.gif is missing is the total derivative operator, which is defined as:
(2.12)
B9781437735208000024/si24.gif is missing
where v=(u, υ, w) is the velocity vector.

2.1.2.2. Conservation of momentum

Conservation of momentum is described by Newton’s second law:
(2.13)
B9781437735208000024/si25.gif is missing
The left-hand side of the above equation represents the acceleration force, while the right-hand side consists of forces per unit volume acting on the fluid particle. The force f may consist of body force and surface forces. In microscale, surface forces such as viscous force, electrostatic force, or surface stress are dominant over body forces such as gravity. If the only body force is gravity and surface forces are caused by a pressure gradient and viscous force and both the density and viscosity are constant, the Navier–Stokes equation can be derived from the general conservation Eqn (2.13):
(2.14)
B9781437735208000024/si26.gif is missing
where ρ and μ are the density and the dynamic viscosity of the fluid, respectively. In micromixers, we often encounter a pressure-driven flow in a straight microchannel. If the flow in the axial x-direction is fully developed, the continuity equation is automatically satisfied with v=w=0 and du/dx=0. The Navier–Stokes Eqn (2.14) can then be further simplified to the two-dimensional form:
(2.15)
B9781437735208000024/si27.gif is missing
where the right-hand side is a constant. Applying the no-slip boundary condition at the channel wall, an analytical solution can be obtained for channels with simple cross-section geometry, such as circle, ellipse, concentric annulus, rectangle, and equilateral triangle (Fig. 2.2). Table 2.2 and Fig. 2.3 show the typical solution of the distribution of the dimensionless velocity B9781437735208000024/si28.gif is missing and the mean velocity B9781437735208000024/si29.gif is missing in the channel.
B9781437735208000024/f02-02ad-9781437735208.jpg is missing
FIGURE 2.2
Cross section of channels considered in Table 2.2: (a) circle; (b) ellipse; (c) concentric annulus; and (d) rectangle.
Table 2.2 Analytical solution for velocity field inside a straight channel
Channel TypeSolution
CircleB9781437735208000024/si1.gif is missing
EllipseB9781437735208000024/si2.gif is missing
Concentric annulusB9781437735208000024/si3.gif is missing
RectangleB9781437735208000024/si4.gif is missing
B9781437735208000024/f02-03ad-9781437735208.jpg is missing
FIGURE 2.3
Distribution of dimensionless velocity according to Table 2.2 (y=y/a, z=z/a): (a) circle; (b) ellipse (a=2b); (c) concentric annulus (a=2b); and (d) rectangle (a=2b).
Many micromachining technologies result in microchannels with a rectangular cross section. The following examples investigate the flows of a single phase as well as multiple phases in rectangular microchannels.
Example 2.3 (Single-phase flow in a rectangular microchannel). A liquid flow in a rectangular microchannel of width W and height H. The viscosity of the liquid is μ. Determine the velocity distribution in this microchannel and the flow rate if a pressure difference of Δp is applied across the channel length of L.
Solution. The fully developed flow in the microchannel is governed by Navier–Stokes equations:
(2.16)
B9781437735208000024/si27.gif is missing
where μ is the dynamic viscosity of the fluid and dp/dx is the pressure gradient along the x-axis. Normalizing the coordinate system by the channel width W (y=y/W, z=z/W) and the velocity by a reference velocity u0 (u=u/u0), the dimensionless Navier–Stokes equation for the hatched regions in Fig. 2.4 is:
B9781437735208000024/f02-04-9781437735208.jpg is missing
FIGURE 2.4
Dimensionless model of a single-phase flow in a rectangular microchannel; only half of the channel cross section (hatched areas) is considered in the analytical model.
(2.17)
B9781437735208000024/si30.gif is missing
where P′=W2/(μ1u0)dp/dx, representing the constant pressure gradient along the channel with the reference velocity u0. The no-slip conditions at the wall result in:
B9781437735208000024/si31.gif is missing
B9781437735208000024/si32.gif is missing
B9781437735208000024/si33.gif is missing
where h=H/W is the dimensionless height of the channel. Using Fourier series analysis, the dimensionless velocity has the form (0<y<1, 0<z<h/2):
B9781437735208000024/si34.gif is missing
If a pressure difference of Δp is applied across the channel length of L, the flow rate is:
B9781437735208000024/si35.gif is missing
For a microchannel with low aspect ratio (h=H/W0), the flow rate can be estimated as:
B9781437735208000024/si36.gif is missing
Example 2.4 (Velocity distribution of streams with different viscosities in a rectangular microchannel). Two immiscible fluids flow side by side in a rectangular microchannel of width W and height H. The viscosity ratio and flow rate ratio of the two fluids are β=μ2/μ1 and B9781437735208000024/si37.gif is missing, respectively. Determine the velocity distribution in this microchannel [4].
Solution. The fully developed flow in the microchannel is governed by Navier–Stokes equations:
B9781437735208000024/si38.gif is missing
where μ1 and μ2 are the viscosities of the two streams and dp/dx is the pressure gradient along the x-axis. Normalizing the coordinate system by the channel W (y=y/W, z=z/W) and the velocity by a reference velocity u0 (u=u/u0), the dimensionless Navier–Stokes equations for the regions 1 and 2 in Fig. 2.5 are:
B9781437735208000024/f02-05-9781437735208.jpg is missing
FIGURE 2.5
Dimensionless model of two immiscible streams for estimating the velocity distribution inside the mixing channel; only half of the channel cross section (hatched areas) is considered in the analytical model.
B9781437735208000024/si39.gif is missing
B9781437735208000024/si40.gif is missing
where P=W2/(μ1u0)dp/dx, representing the constant pressure gradient along the channel with the reference velocity u0, and the ratio of viscosities β=μ2/μ1. The no-slip conditions at the wall result in:
B9781437735208000024/si41.gif is missing
B9781437735208000024/si42.gif is missing
B9781437735208000024/si43.gif is missing
where h=H/W is the dimensionless height of the channel. At the interface position r between the two streams, the velocity and the shear stress are continuous:
B9781437735208000024/si44.gif is missing
For a flat channel (h<<1) and a constant fluid density, the interface position of the two streams can be estimated, based on the mass conservation, as:
B9781437735208000024/si45.gif is missing
The solutions have the Fourier forms (0<y<1, 0<z<h/2):
B9781437735208000024/si46.gif is missing
where θ=(2n1)π/h. The coefficients A1, A2, B1, B2 can be obtained by solving the Fourier series with the previous boundary conditions:
B9781437735208000024/si47.gif is missing
B9781437735208000024/si48.gif is missing
B9781437735208000024/si49.gif is missing
Figure 2.6 shows the typical velocity distribution in a rectangular channel for streams with different flow rates. For streams with the same viscosity, the velocity distribution is flat.
B9781437735208000024/f02-06-9781437735208.jpg is missing
FIGURE 2.6
Velocity distribution in a rectangular mixing channel (h=0.14) with: (a) the same viscosity β=1 and (b) different viscosities β=2.

2.1.2.3. Conservation of energy

The conservation of energy is governed by the first law of thermodynamics:
(2.18)
B9781437735208000024/si50.gif is missing
which means, the change of the total energy in a system is equal to the sum of the heat and work added to the system. The energy equation can then be formulated for the absolute temperature T as
(2.19)
B9781437735208000024/si51.gif is missing
where cp is the specific heat at constant pressure, B9781437735208000024/si52.gif is missing is the thermal expansion coefficient, k is the thermal conductivity, and Φ is the dissipation function. For Newtonian fluid, the dissipation function caused by viscous stress is:
(2.20)
B9781437735208000024/si53.gif is missing
Assuming an incompressible flow, a constant thermal conductivity, and ignoring the kinetic energy change, the energy equation can be simplified to heat-convection equation:
(2.21)
B9781437735208000024/si54.gif is missing

2.1.2.4. Conservation of species

The conservation of species leads to the diffusion/convection equation:
(2.22)
B9781437735208000024/si55.gif is missing
where c is the concentration of the species, D is the diffusion coefficient of the species (solute) in the carrier fluid (solvent), and rg is the generation rate of the species per volume. The above equation assumes a constant, isotropic diffusion coefficient. The left-hand side of (2.22) represents the accumulation and convection of species. The first term on the right-hand side represents molecular diffusion, while the last term is the generation of species. The above conservation equations can also be formulated for the cylindrical and the spherical coordinate systems.

2.2. Molecular Diffusion

2.2.1. Random walk and Brownian motion

A random walk is the path traced by a particle taking successive steps, each in a random direction. The construction of a simple random walk follows the three basic rules:
• The particle starts at a predefined point,
• The distance done by each step is equal, and
• The direction from one point to the next is random.
Following these rules, random walk of a particle can be realized with a simple program. Considering a one-dimensional random walk on a line, the particle has a random choice of two directions for each of its steps. The distance done by each step is assumed to be s. The position x(n) at a step (n) can be described as
(2.23)
B9781437735208000024/si56.gif is missing
where si is a random step, which can take a value of either +s or –s. The squared value of the position x(n) is:
(2.24)
B9781437735208000024/si57.gif is missing
The proportionality comes from the assumption that the same amount of time is taken to make a step B9781437735208000024/si58.gif is missing. The proportionality factor determines how fast the particle moves, which in reality is the diffusion coefficient. Brownian motion is the random walk with very small steps. Figure 2.7 shows the random walk of a particle calculated for 10,000 steps in two and three dimensions.
B9781437735208000024/f02-07ab-9781437735208.jpg is missing
FIGURE 2.7
Random movement of a particle: (a) two-dimensional and (b) three-dimensional.
Jan Ingenhousz, a Dutch doctor, was the first to observe the irregular motion of coal dust particles on the surface of alcohol in 1785 [5]. However, the botanist Robert Brown, who observed, in 1827, pollen particles floating in water under the microscope [6], is credited for the discovery of this motion. The Brownian motion of particles in a liquid is due to the instantaneous imbalance in the force exerted by the small liquid molecules on the particle. Thus, the diffusion coefficient of this particle can be derived from the force balance equation.

2.2.2. Stokes–Einstein model of diffusion

The time evolution of the position of the Brownian particle itself can be described approximately by the force balance equation where the random force of the liquid molecules represents one term in the balance. This equation is called the Langevin equation:
(2.25)
B9781437735208000024/si59.gif is missing
where m is the mass of the particle, β is the friction coefficient, and F(t) is the random force of the liquid molecules. On small time scales, inertial effects are dominant in the Langevin equation. The friction coefficient can be calculated using Stoke’s drag on a spherical particle with a radius of σp:
(2.26)
B9781437735208000024/si60.gif is missing
where μ is the viscosity of the surrounding liquid. The force F(t) is random in time; thus, its auto-correlation function should represent the delta function:
(2.27)
B9781437735208000024/si61.gif is missing
where δ is the Dirac function. Solving (2.25) for the one-dimensional case leads to the particle velocity:
(2.28)
B9781437735208000024/si62.gif is missing
The variance of the displacement x(t) can subsequently be determined as [8]:
(2.29)
B9781437735208000024/si63.gif is missing
where B9781437735208000024/si64.gif is missing is the variance of the particle velocity. For time scale much larger than m/β,
(2.30)
B9781437735208000024/si65.gif is missing
Thus, the diffusion coefficient of the particle can be determined as:
(2.31)
B9781437735208000024/si66.gif is missing
The kinetic energy of the particle is related to the temperature as:
(2.32)
B9781437735208000024/si67.gif is missing
Substituting <u2>=kBT/m in (2.31) results in the Stokes–Einstein equation of the diffusion coefficient [9]:
(2.33)
B9781437735208000024/si68.gif is missing

2.2.3. Diffusion coefficient

2.2.3.1. Diffusion coefficient in gases

Using the kinetic theory discussed in Section 2.1.1, diffusion coefficient in meters squared per second between two gases i and j can be formulated as [7]:
(2.34)
B9781437735208000024/si69.gif is missing
where Mi and Mj are the molecular weights of the gases, T is the absolute temperature, and p is the pressure. The collision diameter σij is the arithmetic average of the characteristic diameter of the gas molecules σi and σj:
(2.35)
B9781437735208000024/si70.gif is missing
The collision integral Ω can be taken from the diagram depicted in Fig. 2.1. In this diagram, the dimensionless temperature is calculated based on the interaction energy, which is the geometric average of the individual characteristic energies:
(2.36)
B9781437735208000024/si71.gif is missing
Example 2.5 (Estimation of diffusion coefficient of gases). Estimate the diffusion coefficient of hydrogen in air at 282K. Lennard–Jones potential parameters of air and hydrogen are (σ1=0.3711nm, B9781437735208000024/si72.gif is missing and (σ2=0.2827nm, B9781437735208000024/si73.gif is missing, respectively. The molecular weights of air and hydrogen are 29 and 2 respectively. The experimental value is 0.710×10−4m2/sec.
Solution. Although air consists mainly of oxygen and nitrogen, we assume air as a gas molecule. According to (2.35), the collision diameter between air and hydrogen is:
B9781437735208000024/si74.gif is missing
According to (2.36), the dimensionless temperature is:
B9781437735208000024/si75.gif is missing
According to Fig. 2.1, the collision potential is approximately 0.88. Thus, the diffusion coefficient of hydrogen in air at 282K is:
B9781437735208000024/si76.gif is missing
The estimated diffusion coefficient is 33% lower than the measured data.

2.2.3.2. Diffusion coefficient in liquids

While diffusion coefficients of gases are on the order of 10−5m2/sec, diffusion coefficients of liquids are on the order of 10−9m2/sec. The diffusion coefficients of large molecules can be on the order of 10−11m2/sec. Diffusion coefficient of a molecule i in a solute j with a viscosity μj can be estimated by the Stokes–Einstein equation:
(2.37)
B9781437735208000024/si77.gif is missing
where σi is the diameter of the molecule i. In the above equation, the term in the numerator represents the kinetic energy of the molecule, while the denominator represents the friction force acting on the molecule. Equation (2.37) breaks down if the size of the solute i is 5 times less than that of the solvent. For a small solute, the factor 3π in (2.37) can be replaced by the factor 2π because there is less friction due to slip on the surface of the solute molecule. If the solute is not spherical but ellipsoid, with dimensions a and b, the characteristic diameter can be estimated as:
(2.38)
B9781437735208000024/si78.gif is missing
Example 2.6 (Diffusion coefficient of a large molecule in water). Fibrinogen is a protein that plays a key role in blood clotting. Fibrinogen significantly increases the risk of stroke. This protein molecule has a rod shape of 67nm in length and 2.2nm in diameter. Estimate the diffusion coefficient of fibrinogen in water at 25°C. Dynamic viscosity of water at this temperature is μ=0.903×10−3Pa·s.
Solution. The large molecule can be assumed to have the ellipsoid shape with a=67×10−9m and b=2.2×10−9nm. According to (2.38), the characteristic diameter of a fibrinogen molecule is:
(2.39)
B9781437735208000024/si79.gif is missing
Thus, based on the Stokes–Einstein equation (2.37), the diffusion coefficient of fibrinogen is:
B9781437735208000024/si80.gif is missing
The diffusion of a large molecule in water is about six orders slower than that of gases.

2.2.3.3. Diffusion coefficient of electrolytes

Salts and many other molecules dissolve in water to form cations and anions. However, they do not diffuse as a single molecule, because different ions have different diffusion coefficients. Table 2.3 and Table 2.4 show the diffusion coefficients of typical cations and anions in water at 25°C. However, both anion and cation together should have the same diffusion coefficient to maintain the charge neutrality. It is interesting to observe from Table 2.3 that the diffusion coefficient of proton H+ is about five times larger than other ions. Since the size difference between H+ and the other ion is not significant, H+ should have a different diffusion mechanism in water. H+ actually does not move through water but reacts with a water molecule and releases a proton on the other side. This new proton can cause another reaction. This chain reaction speeds up the transport process and leads to a much higher diffusion coefficient of H+ in water [7].
Table 2.3 Diffusion coefficients of typical cations in water at 25°C (in 10−9m2/sec) [7]
H+Li+Na+K+Rb+Cs+Ag+NH4+Ca2+Mg2+La3+
9.311.031.331.962.072.061.651.960.790.710.62
Table 2.4 Diffusion coefficients of typical anions in water at 25°C (in 10−9m2/sec) [7]
OHFClBrINO3+CH3COOB(C6H5)4B9781437735208000024/si5.gif is missingB9781437735208000024/si6.gif is missing
5.281.472.032.082.051.901.090.531.060.92
The diffusion coefficient of a molecule consisting of a cation of charge z1 and an anion of charge z2 is:
(2.40)
B9781437735208000024/si81.gif is missing
where D1 and D2 are the diffusion coefficients of the cation and the anion, respectively. According to (2.40), the overall diffusion coefficient is determined by the slower ion. Since the diffusion coefficients are weighted by the charge, the faster ion with a much smaller charge can dominate the overall diffusion coefficient.
Example 2.7 (Diffusion coefficient of sodium chloride). Determine the diffusion coefficient of sodium chloride at 25°C.
Solution. From Table 2.3 and Table 2.4, the diffusion coefficients of sodium cation and chloride anion are D1=1.33×10−9m2/sec and D2=2.03×10−9m2/sec, respectively. Applying z1=+1 and z2=−1 in (2.40) results in the diffusion coefficient of sodium chloride:
B9781437735208000024/si82.gif is missing
The diffusion coefficients of electrolytes in water are about four orders smaller than those of gases and about two orders larger than those of large molecules.
In many lab-on-a-chip applications, micromixers are needed for mixing proteins. The behavior of proteins is very complex. A protein molecule consists of chains of amino acids. The molecular weight is very large and can be on the order of 105. A protein molecule has a large number of side chains that end in amino (–NH2) or carboxylic acid (–COOH) groups. Depending on the pH concentration, amino groups can be positively charged to become (B9781437735208000024/si83.gif is missing) and carboxylic acid groups become negatively charged (–COO). Therefore, the net charge of a protein can be either positive or negative. Further, depending on the pH, the protein chains can be folded differently, resulting in various sizes and shapes. The different charges and shapes make the diffusion coefficient of proteins depend on the pH concentration. Furthermore, the diffusion coefficient of protein also depends strongly on its concentration and the concentration of electrolytes, such as NaCl.
Many solute molecules, such as surfactant, also have diffusion coefficients that depend on the solute concentration. At high concentrations, the molecules can aggregate to form micelles. The aggregation and the electrostatic interaction cause the strong concentration dependence of the diffusion coefficient.

2.3. Taylor Dispersion

Taylor dispersion is an effective mechanism for mixing a solute in a distributed velocity field, such as a pressure-driven flow in a microchannel. This axial effect arises from a coupling between molecular diffusion in the transverse direction and transverse distribution of the flow velocity. Figure 2.8 illustrates the difference between molecular diffusion in a plug-like flow and Taylor dispersion in a distributed flow. In a uniform flow field, such as the plug-like electroosmotic flow, advection and diffusion are independent. Axial diffusion is the same as molecular diffusion (Fig. 2.8 (a)). In a distributed flow field, such as the pressure-driven flow with a parabolic velocity distribution, the solvent is stretched more in the middle of the channel than near the wall, due to axial convective transport. The resulting concentration gradient between the different fluid layers is then blurred by diffusion in the transverse direction (see Fig. 2.8 (b)). As a result, the solute appears to be “diffusing” in the axial direction at a rate that is much faster than what would be predicted by ordinary molecular diffusion.
B9781437735208000024/f02-08ab-9781437735208.jpg is missing
FIGURE 2.8
Particle distribution in a microchannel with: (a) a uniform velocity profile and (b) a parabolic velocity profile.

2.3.1. Two-dimensional analysis

The dispersion coefficients are derived by a two-dimensional model, involving one axial and one transversal spatial dimension. G. I. Taylor was the first to present a working model for the transverse average that managed to capture the influences of both transverse diffusion and the transverse variations of the fluid velocity field. The analysis of Taylor [10] is based on the model of a long cylindrical capillary with a radius r0 (Fig. 2.9). The derivation of the dispersion coefficient given below follows Brenner and Edwards [11]. According to Table 2.2, the velocity distribution in the capillary cross section is:
B9781437735208000024/f02-09-9781437735208.jpg is missing
FIGURE 2.9
Model for determination of Taylor dispersion in a circular channel.
(2.41)
B9781437735208000024/si84.gif is missing
where B9781437735208000024/si85.gif is missing is the mean velocity as listed in Table 2.2. For this capillary, the general equation for conservation of species (2.22), also called the convective/diffusive equation, can be formulated in the cylindrical coordinate system with no species generation (rg=0) as:
(2.42)
B9781437735208000024/si86.gif is missing
where
(2.43)
B9781437735208000024/si87.gif is missing
The boundary and initial conditions of (2.42) are:
(2.44)
B9781437735208000024/si88.gif is missing
Equation (2.42) can be solved numerically with the above boundary and initial condition for c(r, θ, x, t)). Taylor derived an analytical asymptotic solution for (2.42) as follows.
If the observer moves along the flow with the mean velocity B9781437735208000024/si29.gif is missing, we can consider a new axial coordinate x:
(2.45)
B9781437735208000024/si89.gif is missing
Substituting B9781437735208000024/si90.gif is missing into (2.42) results in:
(2.46)
B9781437735208000024/si91.gif is missing
For (2.46), the previous boundary and initial conditions with the new spatial variable x apply. In order to solve (2.46) analytically, the following asymptotic assumptions are made:
• Radial diffusion is complete B9781437735208000024/si92.gif is missing thus, B9781437735208000024/si93.gif is missing
• Axial diffusion is negligible compared to axial convection
B9781437735208000024/si94.gif is missing thus, B9781437735208000024/si95.gif is missing
• Any inhomogeneity in θ is ignored; thus, B9781437735208000024/si96.gif is missing and
• The solute has the same velocity as the solvent.
Next, an average specie concentration is introduced:
(2.47)
B9781437735208000024/si97.gif is missing
With the assumption of the long-time behavior, the axial concentration gradient is independent of B9781437735208000024/si98.gif is missing. Thus, (2.46) can be reduced to:
(2.48)
B9781437735208000024/si99.gif is missing
The above equation can be solved for c by integration with respect to r:
(2.49)
B9781437735208000024/si100.gif is missing
where C(x, t) is the function of integration. Substituting (2.49) into (2.47) results in the r-independent function of the integration constant:
(2.50)
B9781437735208000024/si101.gif is missing
Now, the axial concentration (2.49) can be expressed in terms of the average concentration:
(2.51)
B9781437735208000024/si102.gif is missing
In order to introduce the dispersion coefficient or the so-called effective diffusion coefficient D, the conservation of species (2.46) is written in the flux form as:
(2.52)
B9781437735208000024/si103.gif is missing
where J is the area-averaged axial flux, which consists of both diffusive and convective components:
(2.53)
B9781437735208000024/si104.gif is missing
The convective flux Jconv can be evaluated as:
(2.54)
B9781437735208000024/si105.gif is missing
Substituting (2.51) in (2.54) results in:
(2.55)
B9781437735208000024/si106.gif is missing
The area-averaged axial flux now can be expressed as:
(2.56)
B9781437735208000024/si107.gif is missing
According to Fick’s law, the term in the square bracket of the above equation can be called the effective diffusion coefficient or, more accurately, the dispersion coefficient:
(2.57)
B9781437735208000024/si108.gif is missing
The specie conservation equation can be now formulated as:
(2.58)
B9781437735208000024/si109.gif is missing
The above partial differential equation can be solved analytically. For instance, if the initial condition of the specie concentration is a pulse B9781437735208000024/si110.gif is missing, where δ(x) is the Dirac function, the transient one-dimensional solution of the average concentration is:
(2.59)
B9781437735208000024/si111.gif is missing
Figure 2.10 shows the typical concentration distribution of the one-dimensional dispersion at different time instances.
B9781437735208000024/f02-10-9781437735208.jpg is missing
FIGURE 2.10
One-dimensional dispersion, typical concentration distribution at different time instances.
A few years after Taylor’s publication, Aris provided a firmer theoretical framework for this theory by using a moment analysis [12]. He also generalized the problem to handle time-periodic flows [13]. Following the works of Taylor and Aris, there are several other contributions to the theory of deriving effective transport models for the transverse averages of solutes flowing through channels with more general cross-sectional geometries and flow properties. Common techniques are the use of asymptotic analysis [14], [15], [16] and [17], the theory of projection operators [18], the center manifold theory [19]. All the above works only dealt with nonreactive problems. Johns and DeGance considered the influences of a system of linear reactions upon Taylor dispersion [20]. Yamanaka and Inui used their projection operator theory to solve problems involving a single irreversible reaction [21]. The following example by Bloechle [22] demonstrates an intuitive approach similar to that of Taylor [10]. The approach is called the mean-fluctuation method commonly used in turbulent flow.
Example 2.8 (Taylor dispersion in Poiseuille flow between two parallel plates[22]). Determine the dispersion coefficient of a Poiseuille flow between two parallel plates with a gap of 2h as depicted in Fig. 2.11.
B9781437735208000024/f02-11-9781437735208.jpg is missing
FIGURE 2.11
Model for determination of Taylor dispersion in the Poiseuille flow between two parallel plates.
Solution. Because of the symmetric geometry, only half of the model is considered (−∞<x<∞, 0<y<h). The governing Eqn (2.22) reduces to the two-dimensional form of the parallel plates model (t>0):
(2.60)
B9781437735208000024/si112.gif is missing
where u(y) is the velocity distribution [2]:
(2.61)
B9781437735208000024/si113.gif is missing
The boundary and initial conditions for (2.60) are:
B9781437735208000024/si114.gif is missing
B9781437735208000024/si115.gif is missing
B9781437735208000024/si116.gif is missing
The concentration and velocity can be formulated as the sum of an average component and a fluctuating component:
(2.62)
B9781437735208000024/si117.gif is missing
where
B9781437735208000024/si118.gif is missing
Substituting (2.62) into (2.60) results in:
(2.63)
B9781437735208000024/si119.gif is missing
with
B9781437735208000024/si120.gif is missing
B9781437735208000024/si121.gif is missing
B9781437735208000024/si122.gif is missing
Averaging (2.63) from 0 to h leads to:
(2.64)
B9781437735208000024/si123.gif is missing
The initial condition of the above equation is:
(2.65)
B9781437735208000024/si124.gif is missing
Similar to Taylor’s original approach, the dispersion coefficient can be derived from the above equation if the fluctuation concentration c′ is known. Subtracting (2.64) from (2.63) leads to the partial differential equation for c′:
(2.66)
B9781437735208000024/si125.gif is missing
with
(2.67)
B9781437735208000024/si126.gif is missing
Assuming that
(2.68)
B9781437735208000024/si127.gif is missing
Eqn (2.66) reduces to:
(2.69)
B9781437735208000024/si128.gif is missing
with
(2.70)
B9781437735208000024/si129.gif is missing
Next, (2.69) can be integrated with respect to y:
(2.71)
B9781437735208000024/si130.gif is missing
The symmetry and wall conditions imply that C0(x, t)=0 and
(2.72)
B9781437735208000024/si131.gif is missing
respectively. Integrating (2.71) with respect to y results in:
(2.73)
B9781437735208000024/si132.gif is missing
C1(x, t) can be determined by solving the condition:
(2.74)
B9781437735208000024/si133.gif is missing
The final expression for the fluctuating component of the concentration is:
(2.75)
B9781437735208000024/si134.gif is missing
Substituting (2.75) into (2.64) leads to:
(2.76)
B9781437735208000024/si135.gif is missing
Thus, the dispersion coefficient in a Poiseuille flow between two parallel plates is:
(2.77)
B9781437735208000024/si136.gif is missing
where 2h is the gap between the two parallel plates.

2.3.2. Three-dimensional analysis

For axisymmetric channel geometry, such as a cylindrical capillary, the two-dimensional analysis described in the above subsection is appropriate. For real channel geometry, the cross-sectional velocity profile and, consequently, the dispersion coefficient also depend on the channel shape. In other words, the second transversal spatial dimension needs to be considered in the analysis.
Dutta et al. [23] introduce a factor f into (2.77) to consider the three-dimensional effect of Taylor dispersion:
(2.78)
B9781437735208000024/si137.gif is missing
where d=2h and f=1 for the case of the parallel plate model. The factor f is a function of the aspect ratio d/W, where d is the characteristic length of the shallow channel height and W is the channel width. Based on this definition, the longer cross-sectional dimension is considered as channel width; thus, d/W1. Using the Aris approach [12] and numerical simulation, the factor f can be determined for different geometries. Figure 2.12 shows the dispersion factors of typical channel geometries as a function of the aspect ratio d/W.
B9781437735208000024/f02-12-9781437735208.jpg is missing
FIGURE 2.12
Dispersion factor f for typical channel geometries versus aspect ratio d/W.
(after [23])
Because of the velocity gradient at the sharp corners of a rectangular channel, factor f increases from f=1.76 in the case of a square channel cross section (d/W=1) to f=7.95 in the case of a shallow channel (Wd). The factor of an elliptical channel cross section can be calculated explicitly as:
(2.79)
B9781437735208000024/si138.gif is missing
where B9781437735208000024/si139.gif is missing is the eccentricity of the geometry.
Ajdari et al. [24] argued that for a shallow channel Wd with a continually varying height, the dispersion coefficient is not determined by the channel height and aspect ratio. The channel width W is the only geometric parameter that determines the dispersion coefficient. For instance, the dispersion coefficients for channels with triangle, parabolic, and elliptical cross sections are:
(2.80)
B9781437735208000024/si140.gif is missing
(2.81)
B9781437735208000024/si141.gif is missing
(2.82)
B9781437735208000024/si142.gif is missing
respectively.
In the case of an elliptical cross section, a low aspect ratio Wd refers to an eccentricity of unity ɛ=1. Substituting ɛ=1 in (2.79) results in a factor:
(2.83)
B9781437735208000024/si143.gif is missing
Substituting (2.83) into (2.78) leads to the equation:
(2.84)
B9781437735208000024/si144.gif is missing
Because 5/(192×12)0.0022, the approaches of Dutta [23] and Ajdari [24] agree in the case of a shallow elliptical cross section.

2.4. Chaotic Advection

2.4.1. Basic terminologies

The term chaotic advection refers to the phenomenon where a simple Eulerian velocity field leads to a chaotic response in the distribution of a Lagrangian marker, such as a tracing particle [25]. Advection refers to species transport by the flow. A flow field can be chaotic even in the laminar flow regime. Chaotic advection can be created in a simple two-dimensional flow with time-dependent disturbance or in a three-dimensional flow even without time-dependent disturbance. It is to be noted that chaotic advection is not turbulent. For a flow system without disturbance, the velocity components of chaotic advection at any point in space remain constant over time, while the velocity components of turbulence are random. The streamlines of the steady chaotic advection flow across each other, causing the particles to change their paths. Under chaotic advection, the particles diverge exponentially and enhance the mixing between the solvent and solute flows. In a time-periodic system, the condition for chaos is that streamlines cross at two consecutive time instants.
There are few terminologies related to visualization of an Eulerian velocity field. The first and most common terminology is the pathline, also called trajectory of a fluid particle in the flow field. In experiments, pathlines, orbits, or trajectories can be obtained by an image with a long-time exposure of a fluorescent fluid particle.
If the particles are idealized so that they are small enough not to disturb the flow and large enough so that molecular diffusion is neglected, they can move passively with the flow. The particle transport mechanism can simply be described by the advection equations:
(2.85)
B9781437735208000024/si145.gif is missing
Mathematically, pathlines or trajectories can be obtained by solving (2.85) with the initial condition at t=0 (x=x0, y=y0, z=z0). Numerical integration methods, such as the Runge–Kutta method, can be used for determining the positions of the particles.
In a two-dimensional flow, streamlines are given by the solution of:
(2.86)
B9781437735208000024/si146.gif is missing
where t is treated as a constant and s is the independent variable. Streamlines build the image of the flow field at a time instant. In experiments, streamlines can be constructed from the two images recorded with particle image velocimetry (PIV). The flow is traced with fluorescent particles. Particle images are recorded at two successive time instances t and t+Δt. The particle velocities are determined by the recorded particle displacement and the time delay Δt. The streamlines are tangential to the velocity at each point. The streamlines can be depicted as the level sets of the stream function B9781437735208000024/si147.gif is missing, which is defined as:
(2.87)
B9781437735208000024/si148.gif is missing
A streakline through a point (x, y, z) at a time instance τ is the curve formed by all particles, which (t<τ) passed through this point previously. In experiments, the streakline is the tracing curve of a nondiffusive tracer injected into the flow at the given point.
Example 2.9 (Trajectories and streamlines). An Eulerian velocity field is given as:
B9781437735208000024/si149.gif is missing
All variables and constants are dimensionless. Determine the trajectories and streamlines of particles initially at (x0=1, y0=1), (x0=1, y0=2), and (x0=1, y0=3).
Solution. Solving the differential equations with the initial condition (x0, y0) results in the position of the fluid particle as a function of time:
B9781437735208000024/si150.gif is missing
Figure 2.13 shows graphically the above pathlines, with the three initial particle positions, a=1, b=2, and c=π/2.
B9781437735208000024/f02-13-9781437735208.jpg is missing
FIGURE 2.13
Example 2.9: Pathlines of three particles with initial position at (x0=1, y0=1), (x0=1, y0=2), and (x0=1, y0=3) (a=1, b=2, and c=π/2).
The stream function B9781437735208000024/si147.gif is missing can be solved for the explicit form using the equation system (2.87). Taking the time t as constant, u(x, y, t)=x sin(at), and v(x, y, t)=y sin(bt+c), we have:
B9781437735208000024/si151.gif is missing
It is apparent that streamlines are time dependent. Figure 2.14 shows the streamlines as the level sets of stream function B9781437735208000024/si152.gif is missing.
B9781437735208000024/f02-14-9781437735208.jpg is missing
FIGURE 2.14
Example 2.9: Velocity field and streamlines at t=2 (a=1, b=2, and c=π/2).
Equation (2.85) represents a system of coupled ordinary differential equations (ODEs). Similar dynamical systems in engineering and physics have shown a strong chaotic behavior. Poiseuille flow in a straight microchannel is considered as a one-dimensional incompressible flow at low Reynolds number. The dynamics of this flow is simple and nonchaotic. In the case of a two-dimensional flow, the dynamic behavior of the flow is more interesting. The two-dimensional continuity equation:
(2.88)
B9781437735208000024/si153.gif is missing
is fulfilled by the stream function B9781437735208000024/si147.gif is missing(2.87). Equation system (2.87) has the same form of the Hamilton equation of motion, where the stream function B9781437735208000024/si147.gif is missing plays the role of the Hamiltonian. Thus, steady two-dimensional incompressible flow and time-independent Hamiltonians with one degree of freedom are integrable and deterministic dynamics. Adding one more dimension to the systems, such as unsteady two-dimensional incompressible flow and independent Hamiltonians, makes the equations nonintegrable and causes chaotic dynamics.
The terminologies for chaotic advection can be borrowed from the more established field of dynamical systems theory. The advection equations (2.85) can be solved explicitly for the fluid particle position (x, y, z) at a given time t. The solution of (x, y, z) is then used for describing the motion of fluid particles in a region R, such as a channel cross section, and thus can be called a mapping function S[26]. From an initial condition (t=0), R can be transformed into a new region using S. This transformation or mapping can mathematically be described as S(R). Each transformation is called an advection cycle, which corresponds to a mixing element in different micromixer designs such as sequential lamination, discussed later in this book. Repeating these mixing elements n times refers to repeated application of S to R, or Sn(R). With the discrete number of advection cycles, the transformation S is understood as a discrete, not continuous, operation as in the case of the time function.
If the fluid is assumed to be incompressible, then the volume (in a three-dimensional case) or the area (in a two-dimensional case) are preserved after each transformation. Thus, the above mapping function S is called a volume-preserving transformation or a area-preserving transformation.
A trajectory of a point after applying many discrete transformations is called an orbit. If a point p returns to the same place after N transformations, the orbit is periodic with a period of N. The two typical periodic orbits are the stable elliptic orbit and the unstable hyperbolic orbit. Elliptic orbits lead to a region that does not mix with the surrounding fluid and thus are bad for mixing. Hyperbolic orbits lead to squeezing and stretching of fluid regions and thus are good for mixing. In an unstable orbit, if the fluid particle changes its path at the intersection of the same orbit, the orbit is called homoclinic. If the fluid particle changes its path at the intersection with another orbit, the orbit is called heteroclinic.
The trajectories of a chaotic three-dimensional flow are complicated. Three-dimensional positions of fluid particles can be reduced into a two-dimensional map called the Poincaré section. In a time-periodic system, Poincaré section is a collection of intersections of trajectories with a plane. The continuous trajectories become discrete points of the transformations PnPn+1. The time needed between the two points Pn and Pn+1 does not need to be the period of the system. In three-dimensional space-periodic systems, the plane is taken at the same position of the repeated spatial structure. A trajectory intersects all these periodic planes at several points. The collection of these points forms the Poincaré section. In this case, the transformation PnPn+1 is the advection cycle.

2.4.2. Examples of chaotic advection

2.4.2.1. Lorentz’s convection flow

For a three-dimensional system, the equations in (2.85) are more than enough to create a nonintegrable or chaotic dynamics. Lorenz [27] derived a simplified system of equations for convection rolls in the atmosphere:
(2.89)
B9781437735208000024/si154.gif is missing
where the variable x is proportional to convective intensity, y is proportional to the temperature difference between descending and ascending currents, and z is proportional to the difference in vertical temperature profile from linearity in this system of equations. Pr, Ra, Rac, and β are the Prandtl number, Rayleigh number, critical Rayleigh number, and the geometric factor. Figure 2.15 shows the solution of (2.89) for different Rayleigh numbers. A small change in Rayleigh number leads to a large change in the solution.
B9781437735208000024/f02-15ad-9781437735208.jpg is missing
FIGURE 2.15
Solution of Lorenz’s equations for different Rayleigh numbers (Pr=10 and β=8/3): (a) Ra/Rac=10; (b) Ra/Rac=20; (c) Ra/Rac=30; and (d) Ra/Rac=40.

2.4.2.2. Dean flow in curved pipes

The flow field inside a curved pipe was first derived by Dean [28]. For a more detailed review on flow in curved pipes, see [29]. The following detailed derivation was given by Gratton [30]. The model for the flow in a toroidal pipe is depicted in Fig. 2.16. The pipe has the form of a toroid of a radius of R. The pipe diameter is a. The coordinate system for this model is based on the cylindrical coordinate, where s is the coordinate of the toroid’s center line q. The metric of this coordinate system is:
B9781437735208000024/f02-16-9781437735208.jpg is missing
FIGURE 2.16
Model of Dean flow in a toroidal pipe.
(2.90)
B9781437735208000024/si155.gif is missing
With the assumption of a laminar flow, the change in s is zero. With u, v, and w are velocity components in s, r, and θ. Continuity Eqn (2.11) and Navier–Stokes Eqns (2.14) have the following forms in the new coordinate system:
(2.91)
B9781437735208000024/si156.gif is missing
(2.92)
B9781437735208000024/si157.gif is missing
(2.93)
B9781437735208000024/si158.gif is missing
(2.94)
B9781437735208000024/si159.gif is missing
where p is the pressure and v is the kinematic viscosity. Assuming that the radius of curvature is much larger than the pipe diameter (Ra). The solution of the above four equations can be derived based on the Poiseuille flow:
(2.95)
B9781437735208000024/si160.gif is missing
where A and C are the constants for velocity and pressure gradient, respectively. The tilde-marked variables are the small perturbations. Rewriting the equations and ignoring the small products of the tilde-marked variables lead to:
(2.96)
B9781437735208000024/si161.gif is missing
(2.97)
B9781437735208000024/si162.gif is missing
(2.98)
B9781437735208000024/si163.gif is missing
(2.99)
B9781437735208000024/si164.gif is missing
where ɛ=a/R is the ratio between the pipe diameter and the radius of curvature, and all the higher order of the tilde-marked terms as well as of ɛ are neglected. Setting the small terms in (2.97) to zero
(2.100)
B9781437735208000024/si165.gif is missing
and rearranging for C:
(2.101)
B9781437735208000024/si166.gif is missing
Equation (2.97) then has the form:
(2.102)
B9781437735208000024/si167.gif is missing
Separating variables according to r and θ and substituting the separated variables:
(2.103)
B9781437735208000024/si168.gif is missing
in (2.96), (2.98), (2.99) and (2.102), normalizing the velocities by B9781437735208000024/si169.gif is missing and spatial variables by a and applying the no-slip boundary condition at r=0 result in the dimensionless velocity components:
(2.104)
B9781437735208000024/si170.gif is missing
where B9781437735208000024/si171.gif is missing is the Reynolds number. Using the Cartesian coordinates (x=r sinθ, y=r cosθ), the three velocity components have the form:
(2.105)
B9781437735208000024/si172.gif is missing
For the s component, all the ɛ terms are canceled from the expression of u in (2.104). The functions of h(r) and h′(r) are:
(2.106)
B9781437735208000024/si173.gif is missing
Further normalization of the time and s by Reɛ/144 and Reɛ/288 results in the dimensionless velocity components:
(2.107)
B9781437735208000024/si174.gif is missing
Figure 2.17 shows the dimensionless three-dimensional velocity field (2.107) of a pipe cross section graphically. The effect of the centrifugal force can be observed clearly. The stream function can be determined in the same way as shown in Example 2.9:
B9781437735208000024/f02-17-9781437735208.jpg is missing
FIGURE 2.17
The dimensionless velocity field of a pipe cross section.
(2.108)
B9781437735208000024/si175.gif is missing
Figure 2.18 (a) shows the secondary velocity field of the pipe cross section. The inner side of the torus is on the left. Centrifugal force causes the fluid to move outward. The streamlines depicted in Fig. 2.18 (b) show the two vortices on the lower and upper half of the pipe.
B9781437735208000024/f02-18-9781437735208.jpg is missing
FIGURE 2.18
Dean vortices: (a) secondary velocity field and (b) stream line of the secondary velocity field.
The trajectories of the fluid particles are calculated using the velocity solutions (2.107) and numerical integration with the Runge–Kutta method. Projecting the particle position on a single two-dimensional cross section results in the Poincaré section. Figure 2.19 shows the trajectories and Poincaré sections of the Dean flow where the s-axis is straightened for clarity. The results clearly show that independent of the orientation seeding lines, the Poincaré sections follow the streamlines as depicted in Fig. 2.18 (b). If this flow is used in a micromixer, the solvent and solute should be introduced on the left and right of the cross section or the outer and inner side of the curved channel (Fig. 2.18), so that the trajectories of the fluid particle can sample both sides of the channel. If the solvent and solute are introduced in the upper and lower halves of the channel, the trajectories will keep them in their respective channel section and advective mixing will not work. Even if the solvent and solute enter at the outer and inner side of the curved channel, the trajectories are stable, elliptic, and homoclinic. Transversal transport is advective but not chaotic. This means, they do not cross each other. Therefore, chaotic advection cannot be realized with the original Dean flow.
B9781437735208000024/f02-19ab-9781437735208.jpg is missing
FIGURE 2.19
Trajectories and Poincaré sections for fluid particles in a Dean flow: (a) seeding line parallel to x-axis and (b) seeding lines parallel to y-axis.
The above analysis assumes a small ratio between the pipe diameter and radius of curvature, a/R1. For realistic channel designs, this ratio can be approximately unity, and the secondary flows are more obvious. In this case, the flow is characterized by the Dean number:
(2.109)
B9781437735208000024/si176.gif is missing
where Re is the Reynolds number. The Dean number represents the ratio between centrifugal force and the inertial force. There exists a critical Dean number Decr=150 where the secondary flow pattern changes. For De<150, there are only a pair of counter-rotating vortices as analyzed above. At higher Dean numbers De>150, the centrifugal force is dominant, leading to the formation of two additional vortices at the outer channel wall. This effect and its application for mixing at a high Dean number or high Reynolds number will be discussed later in Section 5.1.

2.4.2.3. Flow in helical pipes [32]

In helical pipes, the cross section rotates around the pipe center line q (Fig. 2.20). Assuming a constant torsion of τ, the rotation at s is τs. Further, a constant curvature k=dΦ/ds is assumed. Based on this assumption, Germano [31] derived the metric:
B9781437735208000024/f02-20-9781437735208.jpg is missing
FIGURE 2.20
Model of a helical pipe.
(2.110)
B9781437735208000024/si177.gif is missing
The flow in this coordinate system has only a second-order dependence on τ. Thus, the solution (2.107) can be used for helical pipes by changing the basis to the new coordinate system (2.110). The solution for the velocity field is then:
(2.111)
B9781437735208000024/si178.gif is missing
where the curvature and torsion are combined in the geometry parameter:
(2.112)
B9781437735208000024/si179.gif is missing
where B9781437735208000024/si29.gif is missing is the mean velocity in s direction. The stream function of (2.111) has the form:
(2.113)
B9781437735208000024/si180.gif is missing
Figure 2.21 shows the streamlines calculated using (2.113). The initial positions of the particles in the depicted trajectories are on a seeding line parallel to the y-axis. The results show that at increasing torsion, the secondary flow transforms from two counter-rotating vortices into a single vortex.
B9781437735208000024/f02-21-9781437735208.jpg is missing
FIGURE 2.21
Streamlines of the flow inside a helical pipe of different geometry parameters: (a) λ=1; (b) λ=2; and (c) λ=3.
The Poincaré sections of flow helical pipes with different geometry parameters are shown in Fig. 2.22. The flow is initially sampled with a seeding line parallel to the y-axis. At λ=0, there is no torsion and the pipe is a torus. The flow is clearly not chaotic; the particles follow the streamlines. At λ>0, chaotic advection is apparent.
B9781437735208000024/f02-22-9781437735208.jpg is missing
FIGURE 2.22
Poincaré sections of different geometry parameters: (a) λ=0; (b) λ=1; (c) λ=2; (d) λ=3; (e) λ=4; and (f) λ=5.

2.4.2.4. Flow in twisted pipes [32]

While a straight channel is one dimensional, a three-dimensional flow (transverse cross-sectional plane and longitudinal axis) can be created in curved channels. In such curved channels, secondary vortices in the transverse cross-sectional plane can move fluid particles between the center of the channel and its wall. A unit of the simplest configuration for chaotic advection is depicted in Fig. 2.23. The secondary flow in a twisted pipe causes the so-called Dean vortices. Fluid particles rotate with an angle of χ between successive units. The following analysis was reported by Jones et al. [32].
B9781437735208000024/f02-23-9781437735208.jpg is missing
FIGURE 2.23
The basic unit of a twisted pipe consisting of two C-shaped sections.
Using the polar coordinate system (θ, r) in the transverse x – y plane, the stream function B9781437735208000024/si147.gif is missing and the axial velocity u are determined through the following dimensionless equation system:
(2.114)
B9781437735208000024/si181.gif is missing
where De is the Dean number, a is the pipe radius, and R is the radius of curvature of the C-shape bend. The Dean number describes the ratio between the centrifugal force and the viscous force. From (2.109), the Dean number is proportional to the Reynolds number. The dimensionless pressure gradient is defined as:
(2.115)
B9781437735208000024/si182.gif is missing
where B9781437735208000024/si29.gif is missing is the average axial velocity. In (2.114), the lengths and velocities are normalized by the pipe radius a and the average axial velocity B9781437735208000024/si29.gif is missing, respectively.
The first-order perturbation solution results in the following equations of the particle motion [32]:
(2.116)
B9781437735208000024/si183.gif is missing
where α=DeC2, and β=DeC/Re.
Using the angle θ to describe the three-dimensional motion of fluid particles, the velocity components in x – y plane can be formulated as [32]:
(2.117)
B9781437735208000024/si184.gif is missing
This solution results in chaotic advection for a combination of α/β and X. For instance, the most chaotic pattern is achieved with α/β=100 and χ=90°. The condition for chaotic advection to occur in this configuration is:
(2.118)
B9781437735208000024/si185.gif is missing
FIGURE 2.24, FIGURE 2.25 and FIGURE 2.26 show the Poincaré sections with the different model parameters.
B9781437735208000024/f02-24-9781437735208.jpg is missing
FIGURE 2.24
Poincaré sections of twisted pipes with different twisting angles χ (α/β=100): (a) χ=0; (b) χ=π/16; (c) χ=π/8; (d) χ=π/4; (e) χ=3π/8; (f) χ=π/2; (g) χ=5π/8; (h) χ=3π/4; (i) χ=7π/8.
(Reproduced from [32] by permission of Cambridge University Press.)
B9781437735208000024/f02-25-9781437735208.jpg is missing
FIGURE 2.25
Poincaré sections of twisted pipes with different twisting angles χ (α/β=200): (a) χ=π/4; (b) χ=π/2; (c) χ=3π/4.
(Reproduced from [32] by permission of Cambridge University Press.)
B9781437735208000024/f02-26-9781437735208.jpg is missing
FIGURE 2.26
Poincaré sections of twisted pipes with different ratios α/β (χ=π/2): (a) α/β=50; (b) α/β=150; (c) α/β=250.
(Reproduced from [32] by permission of Cambridge University Press.)

2.4.2.5. Flow in a droplet

With the increasing popularity of droplet-based microfluidics, mixing in droplets becomes a crucial task in designing a droplet-based lab-on-a-chip. The analytical solution for the internal flow inside a droplet was first reported by Hadamard [33]. Consider a spherical microdroplet with a radius a. The droplet experiences a uniform shear flow of a velocity B9781437735208000024/si29.gif is missing in the z-axis. We now consider the viscosity ratio β=μ1/μ2, where μ1 is the viscosity of the droplet fluid and μ2 is the viscosity of the surrounding fluid. Normalizing the spatial variables by the droplet radius a, the velocity by B9781437735208000024/si29.gif is missing and the time by B9781437735208000024/si186.gif is missing result in the dimensionless equations of particle motion inside the droplet [34]:
(2.119)
B9781437735208000024/si187.gif is missing
The flow described by (2.119) is actually one dimensional and not chaotic because it has two invariants [35]:
(2.120)
B9781437735208000024/si188.gif is missing
In order to create chaotic advection, an external shear flow u=Gy is superimposed on the uniform flow B9781437735208000024/si29.gif is missing. Defining the dimensionless parameter B9781437735208000024/si189.gif is missing results in the following equations of particle motion inside the droplet under only the shear flow:
(2.121)
B9781437735208000024/si190.gif is missing
where B9781437735208000024/si191.gif is missing is the radial variable in the spherical coordinate system. Superposition of (2.119) and (2.121) results in the equations of motions of a fluid particle inside a droplet immersed in a combined shear and uniform flow as shown in Fig. 2.27.
B9781437735208000024/f02-27-9781437735208.jpg is missing
FIGURE 2.27
Streamlines of flow inside a spherical droplet; the flow direction of the surrounding fluid is in z-axis: (a) external uniform flow; (b) external shear flow; (c) superposition flow with α=0, β=1; (d) superposition flow with α=0.2, β=1; (e) superposition flow with α=0.4, β=1; and (f) superposition flow with α=0.6, β=1.
(Reprinted with permission from [34].)

2.5. Viscoelastic effects

In most analysis and design considerations of micromixers, the solute and solvent are assumed to be Newtonian fluids. In these fluids, the viscosities do not depend on the velocity gradient or the shear stress. This means at a given temperature and pressure, the viscosity is a constant and the velocity gradient is linearly proportional to the shear stress. The Newtonian assumption is true for solvents and solutes with small molecules. However, if they contain large molecules such as long polymers, the viscosity is also a function of the shear stress. Since the shear stress and viscosity gradient in microscale increase with miniaturization, nonlinear effects can be expected and exploited for mixing applications.
At the molecular level, viscoelastic fluids can be described by two models: network model and single-molecule model [2]. The network model is based on the formation and rupture of junctions between polymer molecules. The network model is suitable for solutions with high polymer concentration. Dilute solutions are better described with a single-molecule model where interactions between the molecules are not frequent. The polymer molecule is represented by a “dumb-bell” or “bead-string” model where two spheres are connected by a spring. Kinetic theory with Stokes drag theory and Brownian motion can be used with this model for deriving macroscopic properties.
Fluids with large molecules display elastic behavior due to the stretching and coiling of the polymer chain. Here, these fluids and their behaviors are called viscoelastic fluids and viscoelastic effects respectively. The most apparent viscoelastic effect is the change of velocity profile in a channel. The dimensionless velocity profile of a viscoelastic fluid in a circular capillary can be approximated as [2]:
(2.122)
B9781437735208000024/si192.gif is missing
where 0<n<1 is a parameter unique for the fluid and r0 is the radius of the capillary. If n=1, the fluid becomes Newtonian and the velocity profile is parabolic.
The next viscoelastic effect, which is relevant to mixing in microscale, is the entry flow at a contraction. The operation point of a viscoelastic flow is represented by the Wi–Re diagram, where Wi is the Weissenberg number and Re is the Reynolds number. With a characteristic length scale Lc, the mean velocity B9781437735208000024/si29.gif is missing, the density ρ, and the zero-stress viscosity μ0, the Reynolds number is defined here as:
(2.123)
B9781437735208000024/si193.gif is missing
The Weissenberg number represents the elastic character of the fluid by using the ratio between the relaxation time λ of the fluid and the characteristic residence time τflow:
(2.124)
B9781437735208000024/si194.gif is missing
The characteristic residence time is the inverse value of the characteristic shear rate B9781437735208000024/si195.gif is missing and is defined as:
(2.125)
B9781437735208000024/si196.gif is missing
Because both Reynolds number and Weissenberg number are proportional to the average velocity B9781437735208000024/si29.gif is missing, it is useful to define the elasticity number, which is independent of the flow velocity:
(2.126)
B9781437735208000024/si197.gif is missing
Elasticity number represents the importance of elastic effect over the inertial effect. For the small Reynolds number in microfluidics, the inertial effect is negligible. However, if the elasticity number of the fluid is large enough, elastic effect may also be large enough to compete with the dominant viscous effect. Figure 2.28 shows the typical operation region of a possible micromixer based on viscoelastic instabilities of a 4-to-1 contraction. Shear-thinning viscoelastic fluids are, for instance, concentrated polymer solutions. Boger fluids are dilute solutions of a polymer and a solvent. The slopes of the curves represent the elasticity numbers [36].
B9781437735208000024/f02-28-9781437735208.jpg is missing
FIGURE 2.28
The Wi–Re diagram (based on a planar 4:1 contraction flow); the gray area represents the operation region of micromixers based on viscoelastic instabilities.

2.6. Electrokinetic Effects

2.6.1. Electroosmosis

Electrokinetic effects are based on an electric double layer at the interface between a solid and a liquid or between two liquids. This double layer is also called the Debye layer. This section focuses on the interface between a solid and a liquid. In general, there are four basic electrokinetic effects:
Electroosmosis is the flow of a liquid in an electric field relative to a stationary charged surface.
Electrophoresis is the motion of a charged particle in an electric field relative to the surrounding liquid.
Streaming potential is the opposite effect of electroosmosis. An electric potential is created when a liquid is forced to flow relative to a charged surface.
Sedimentation potential is the opposite effect of electrophoresis. An electric potential is created when charged particles are forced to flow relative to a surrounding liquid.

2.6.1.1. The Debye layer

The Debye layer is an electric double layer, which occurs due to interaction between an electrolyte and a charged solid surface. Ions in an electrolyte solution are attracted to the charged surface and form a thin charge layer, which is called the Stern layer. The Stern layer is attracted to the surface due to the electrostatic force. The layer leads to the formation of a thicker charge layer in the solution. This diffuse and mobile layer is called the Gouy–Chapman layer. The Stern layer and the Gouy–Chapman layer together form the Debye layer (Fig. 2.29 (a)). Because the Gouy–Chapman layer is mobile, it can move if an electric field is applied. The interface between the Stern layer and the Gouy–Chapman layer is called the shear surface. The potential of the charged solid surface is called the wall potential Ψwall. The potential of the shear surface is called the zeta potential (Fig. 2.29 (b)). The potential distribution in the electrolyte solution can be described by the one-dimensional Poisson equation:
B9781437735208000024/f02-29-9781437735208.jpg is missing
FIGURE 2.29
The Debye layer: (a) the Stern layer and the Gouy–Chapman layer and (b) the potential distribution near the wall.
(2.127)
B9781437735208000024/si198.gif is missing
where ρel and ɛ=ɛ0ɛr are the electric charge density and the dielectric constant of the electrolyte, respectively. Assuming the Boltzmann distribution for the charge density, the ion concentration in the electrolyte solution can be determined as:
(2.128)
B9781437735208000024/si199.gif is missing
where ni is the ion concentration of the electrolyte with a unit of 1/m3, zi is the ionic valence, and e=1,602×10−19 is the elementary charge. Thus, the total charge in the double layer is:
(2.129)
B9781437735208000024/si200.gif is missing
The charge density ρel in a symmetric electrolyte is proportional to the concentration difference between cations and anions:
(2.130)
B9781437735208000024/si201.gif is missing
Combining (2.127) and (2.130) results in the Poisson–Boltzmann equation:
(2.131)
B9781437735208000024/si202.gif is missing
Under conditions such as a large characteristic length compared to the double layer thickness or a high ion concentration in the electrolyte and small zeta potential relative to 25 mV, the right-hand side of (2.131) can be linearized by the relation sinh(x)=x:
(2.132)
B9781437735208000024/si203.gif is missing
where λD is the double layer thickness, which is called the Debye length:
(2.133)
B9781437735208000024/si204.gif is missing
Solving (2.132) results in the potential distribution:
(2.134)
B9781437735208000024/si205.gif is missing

2.6.1.2. Electroosmotic transport effect

The continuum models using mass and energy conservation equations can be used for describing the electroosmotic transport effects. The conservation of momentum needs to consider the electrostatic force created by the electric field:
(2.135)
B9781437735208000024/si206.gif is missing
If there is no pressure gradient applied to the flow, (2.135) has the one-dimensional form:
(2.136)
B9781437735208000024/si207.gif is missing
With the assumption of a thin Debye length compared to the channel diameter, the electrokinetic velocity is:
(2.137)
B9781437735208000024/si208.gif is missing
The velocity ueo is also called the Smoluchowski velocity. If the Debye length is negligible compared to other channel dimensions, the electrokinetic flow can be modeled with slip boundary condition, where the slip velocity is the Smoluchowski velocity (Fig. 2.30).
B9781437735208000024/f02-30-9781437735208.jpg is missing
FIGURE 2.30
Electroosmotic flow in a capillary with a negatively charged wall.
With a constant viscosity μ and a constant zeta potential B9781437735208000024/si209.gif is missing, the electroosmotic velocity is proportional to the electric field strength Eel. The negative sign shows that the flow direction is opposite to the field direction. The proportional factor is called the electroosmotic mobility:
(2.138)
B9781437735208000024/si210.gif is missing
Equation (2.137) shows that the analysis of electrokinetic flows in a microchannel network can be replaced by the analysis of a resistance network. Electric currents and potentials can be calculated based on the basic Kirchhoff law:
• The sum of all currents at a node is zero,
• The sum of all voltages in a closed loop is zero.
After determining the potentials at the nodes of the network, the field strengths in each microchannel can be calculated. The velocity can then be determined by the given electroosmotic mobility.

2.6.1.3. Electrokinetic flow between two parallel plates

Figure 2.31 shows the model of electrokinetic flow between two parallel plates. The velocity distribution U(y) of an electrokinetic flow between two parallel plates can be derived from the Navier–Stokes equation. For further simplicity, the variable U(y) is introduced:
B9781437735208000024/f02-31-9781437735208.jpg is missing
FIGURE 2.31
Model for electrokinetic flow between two parallel plates.
(2.139)
B9781437735208000024/si211.gif is missing
Equation (2.136) has then the homogenous form:
(2.140)
B9781437735208000024/si212.gif is missing
With the boundary conditions:
B9781437735208000024/si213.gif is missing
the solution of (2.140) is:
(2.141)
B9781437735208000024/si214.gif is missing
Introducing the dimensionless velocity u, dimensionless potential Ψ, dimensionless zeta potential ζ, and the dimensionless spatial variable y:
(2.142)
B9781437735208000024/si215.gif is missing
The solution (2.141) and the Poisson–Boltzmann Eqn (2.132) have the dimensionless forms:
(2.143)
B9781437735208000024/si216.gif is missing
(2.144)
B9781437735208000024/si217.gif is missing
The boundary conditions for (2.144) are:
B9781437735208000024/si218.gif is missing
The solution of (2.144) is:
(2.145)
B9781437735208000024/si219.gif is missing
Substituting (2.145) into (2.143) results in the dimensionless velocity distribution of a electrokinetic flow between two parallel plates:
(2.146)
B9781437735208000024/si220.gif is missing

2.6.1.4. Electrokinetic flow in a cylindrical capillary

Figure 2.32 shows the model of electrokinetic flow in a cylindrical capillary. The Navier–Stokes equation and the Poisson–Boltzmann equation are formulated for the cylindrical coordinate system:
B9781437735208000024/f02-32-9781437735208.jpg is missing
FIGURE 2.32
Model of electrokinetic flow in a cylindrical capillary.
(2.147)
B9781437735208000024/si221.gif is missing
(2.148)
B9781437735208000024/si222.gif is missing
For simplicity, the following dimensionless variables are introduced:
(2.149)
B9781437735208000024/si223.gif is missing
where R is the capillary radius. Equations (2.147) and (2.148) then have the linearized (Section 2.6.1.1) dimensionless forms:
(2.150)
B9781437735208000024/si224.gif is missing
(2.151)
B9781437735208000024/si225.gif is missing
The dimensionless boundary conditions are:
(2.152)
B9781437735208000024/si226.gif is missing
Solving (2.150) results in:
(2.153)
B9781437735208000024/si227.gif is missing
where J0 and J1 are the Bessel functions (of the first kind) of order 0 and 1, respectively. λn is the nth positive zero value of the Bessel function J0(λn)=0. Cn is a function of the dimensionless potential Ψ:
(2.154)
B9781437735208000024/si228.gif is missing
The boundary conditions for (2.151) are:
(2.155)
B9781437735208000024/si229.gif is missing
The solution for the dimensionless potential is:
(2.156)
B9781437735208000024/si230.gif is missing
where I0(x)=inJ0(ix) are the modified Bessel functions of the first kind and zero order. Substituting (2.156) into (2.154) results in:
(2.157)
B9781437735208000024/si231.gif is missing
The dimensionless velocity distribution in a cylindrical capillary is then:
(2.158)
B9781437735208000024/si232.gif is missing

2.6.1.5. Electrokinetic flow in a rectangular microchannel

Due to the characteristics of microtechnology, many micromixers have a rectangular cross section. Figure 2.33 shows the model of electrokinetic flow in a rectangular microchannel. The Navier–Stokes equation and the Poisson–Boltzmann equation are formulated in the Cartesian coordinate system as:
B9781437735208000024/f02-33-9781437735208.jpg is missing
FIGURE 2.33
Model for electrokinetic flow in a rectangular microchannel.
(2.159)
B9781437735208000024/si233.gif is missing
Following dimensionless variables are introduced:
(2.160)
B9781437735208000024/si234.gif is missing
where Dh=4WH/(W+H) is the hydraulic diameter of the microchannel. For simplicity, the approximation of sinh (x)=x is used. As mentioned above, this assumption is correct if the hydraulic diameter Dh is much larger than the Debye length λD or the ion concentration is dilute. At the molecular scale, this assumption means that the electric energy of the ions is much smaller than their thermal energy. Using the above dimensionless variables, the Navier–Stokes equation and the Poisson–Boltzmann equation have their dimensionless forms:
(2.161)
B9781437735208000024/si235.gif is missing
(2.162)
B9781437735208000024/si236.gif is missing
The dimensionless number
(2.163)
B9781437735208000024/si237.gif is missing
describes the interplay between the electrokinetic force and the friction force. Using the dimensionless boundary conditions
B9781437735208000024/si238.gif is missing
and solving (2.161) lead to the dimensionless velocity distribution of the electrokinetic flow:
(2.164)
B9781437735208000024/si239.gif is missing
where
B9781437735208000024/si240.gif is missing
The potential distribution:
B9781437735208000024/si241.gif is missing
is obtained by solving the Poisson–Boltzmann Eqn (2.162), with the boundary conditions
B9781437735208000024/si242.gif is missing

2.6.1.6. Ohmic model for electrolyte solutions

In this section, a model for electrolyte solutions is derived. This model is useful for formulating mixing problems in an electrokinetic system. The model was formulated by Chen et al. [37], who followed the approach of Levich [38]. We consider here a monovalent binary electrolyte (|z+|=|z|=1), where the subscripts + and − denote the cation and anion, respectively. The local charge density and conductivity σel are determined as:
B9781437735208000024/si243.gif is missing
B9781437735208000024/si244.gif is missing
where F is the Faraday constant, m is the ionic mobility, and c is the concentration. Electro-neutrality can be evaluated based on the ratio between the concentration difference of cations and anions and the total concentration of ions.
(2.165)
B9781437735208000024/si245.gif is missing
While the concentration difference contributes to the charge density, the total ion concentration contributes to the electrical conductivity. Thus, electroneutrality can be assumed if the above ratio is very small, Θ1. Under electroneutrality, the concentration of both ion types is the same, c+=c=c, which is called the reduced concentration. The conductivity is then:
(2.166)
B9781437735208000024/si246.gif is missing
The conservation of species can be formulated for the ions as:
(2.167)
B9781437735208000024/si247.gif is missing
(2.168)
B9781437735208000024/si248.gif is missing
where D is the effective diffusion coefficient. According to (2.40), the diffusion coefficient of the ion with |z+|=|z|=1 is:
(2.169)
B9781437735208000024/si249.gif is missing
where D+ and D are the diffusion coefficients of the cations and anions, respectively. Since the convection current is zero (iC=ρeEel) due to electroneutrality, the electrical current in the electrolyte is caused by electro-migration and diffusion (iO>>iD):
(2.170)
B9781437735208000024/si250.gif is missing
where E is the electric field strength. Diffusive current iD is usually much smaller than the electro-migration current iO; thus, B9781437735208000024/si251.gif is missing. The conservation of ionic species and currents can be formulated for the conductivity as:
(2.171)
B9781437735208000024/si252.gif is missing
(2.172)
B9781437735208000024/si253.gif is missing
With Gauss’s law B9781437735208000024/si254.gif is missing the current continuity equation can be written as:
(2.173)
B9781437735208000024/si255.gif is missing
where ɛ is the permittivity of the liquid and is assumed to be uniform.

2.6.2. Electrophoresis

Electrophoresis is the motion of a charged particle relative to the surrounding liquid in an electric field (Fig. 2.34). Because of the small size and low Reynolds number involved, the Stokes model can be assumed for the motion of the particle:
B9781437735208000024/f02-34-9781437735208.jpg is missing
FIGURE 2.34
Electrophoretic motion of a positively charged sphere.
(2.174)
B9781437735208000024/si256.gif is missing
where qsurf, B9781437735208000024/si257.gif is missing, and rp are the surface charge, the particle velocity and the radius of the particle. The charge density on the particle surface is:
(2.175)
B9781437735208000024/si258.gif is missing
If the radius of the particle is much larger than the Debye length (rp/λD1), the surface charge can be calculated as follows:
(2.176)
B9781437735208000024/si259.gif is missing
Combining (2.174) and (2.176) results in the velocity of the particle:
(2.177)
B9781437735208000024/si260.gif is missing
At a constant dynamic viscosity μ and a constant zeta potential ζ, the electrophoretic velocity is proportional to the field strength Eel. The proportional factor is called the electrophoretic mobility:
(2.178)
B9781437735208000024/si261.gif is missing
If the particle radius is much smaller than the Debye length (rp/λD<<1), the electrophoretic mobility approaches the electroosmotic mobility on a flat wall:
(2.179)
B9781437735208000024/si262.gif is missing
The electrophoretic mobility of a particle can generally be formulated as follows:
(2.180)
B9781437735208000024/si263.gif is missing
where the correction factor C is a function of the ratio rp/λD :
(2.181)
B9781437735208000024/si264.gif is missing
B9781437735208000024/si265.gif is missing

2.6.3. Dielectrophoresis

Dielectrophoresis (DEP) is the motion of a dielectric particle in a dielectric fluid. Because the particle is charge neutral, dielectric force is caused by the inhomogeneity of the electric field. Assuming a homogenous linear dielectric fluid with a susceptibility of χ, the polarization field P of the fluid is given as:
(2.182)
B9781437735208000024/si266.gif is missing
The displacement field D of the fluid is:
(2.183)
B9781437735208000024/si267.gif is missing
where ɛf=ɛ0(1+χ) is the permittivity of the fluid. The relation between the displacement field and the charge density is:
(2.184)
B9781437735208000024/si268.gif is missing
The dielectric force acting on a dipole moment in an inhomogeneous electric field is:
(2.185)
B9781437735208000024/si269.gif is missing
For a spherical particle with the permittivity ɛp, the polarization leads to a dipole moment m:
(2.186)
B9781437735208000024/si270.gif is missing

2.7. Magnetic and Electromagnetic Effects

2.7.1. Magnetic effects

Although magnetic forces are body forces and therefore do not scale favorably in micromixers, high field gradients can be achieved with integrated microcoils. The use of liquids with magnetic properties and an external actuating magnetic field promises to be a niche for inducing transversal transport and chaotic advection in micromixers. The best candidate for this concept is ferrofluid. Pure substances such as liquid oxygen also behave like a magnetic liquid or ferrofluid. However, the term ferrofluid is commonly referred to as colloidal ferrofluid. The magnetic property of this fluid is credited to ferromagnetic nanoparticles, usually magnetite, hematite, or some other compounds containing iron 2+ or 3+. These nanoparticles are solid, single-domain magnetic particles that are suspended in a carrier fluid. The particles are coated with a monolayer of surfactant molecules to avoid them to stick to each other. Because the size of the particles is on the order of nanometers, Brownian motion, which represents the kinetic or thermal energy of the particles, is able to disperse them homogenously in the carrier fluid. The dispersion is strong enough that the solid particles do not agglomerate or separate even under strong magnetic fields. A typical ferrofluid is opaque to visible light. It is to be noted that the term magnetorheological fluid (MRF) refers to liquids similar in structure to ferrofluids but differing in behavior. MRF particle sizes are on the order of micrometers that are one to three orders of magnitude larger than those of ferrofluids. MRF also has a higher volume fraction on the order of 20–40%. Exposing MRF to a magnetic field can transform it from a light viscous fluid to a thick solid-like material [39].
Ferromagnetic nanoparticles are fabricated based on size reduction through ball milling, chemical precipitation, and thermophilic iron reducing bacteria. In ball milling, magnetic material of several micrometers in size such as magnetite powder is mixed with carrier liquid and surfactant. The ball milling process takes approximately 1000h. Subsequently, the product mixture undergoes centrifuge separation to filter out oversize particles. The purified mixture can be concentrated or diluted in the final ferrofluid.
Synthesis by chemical precipitation is a more common approach in which the particles precipitate out of solution during chemical processes. A typical reaction for magnetite precipitation is:
(2.187)
B9781437735208000024/si271.gif is missing
The reaction product is subsequently coprecipitated with concentrated ammonium hydroxide NH4OH. Next, a peptization process transfers the particles from water-based phase to an organic phase, such as kerosene with a surfactant, for example, oleic acid. The oil-based ferrofluid can then be separated by a magnetic field.
Another approach for fabrication of ferrofluid is based upon thermophilic bacteria that reduce amorphous iron oxyhydroxides to nanometer-sized iron oxides. The thermophilic bacteria are able to reduce a number of different metal ions. Thus, this approach allows incorporating other compounds, such as Mn(II), Co(II), Ni(III), Cr(III)), into magnetite. Varying the composition of the nanoparticles can adjust magnetic, electrical, and physical properties of the substituted magnetite and consequently of the ferrofluid. Ferromagnetic particle with extremely low Curie temperature can be designed with this method. Most of the particle materials commonly used in ferrofluid have much higher Curie temperatures. The temperature dependence of magnetic properties can be used for micromixing applications.
At the typical channel size of microfluidics (about 100μm), ferrofluid flow in a microchannel can be described as a continuum flow. The governing equations are based on the conservation of mass and conservation of momentum. In the case of temperature-dependent magnetic properties, the conservation of energy may be needed for calculating the temperature field. The Navier–Stokes equation for a ferrofluid has the following form:
(2.188)
B9781437735208000024/si272.gif is missing
Compared to other types of fluid, ferrofluid flow in microchannel has an additional term for magnetic force [39]:
(2.189)
B9781437735208000024/si273.gif is missing
where μ0=4π×10−7 Hm−1 is the permeability of space, M is the intensity of magnetization, v is the specific volume, and H is the magnetic field strength in A/m. The magnetic term can be grouped together with static pressure to form an apparent pressure. Thus, the conservation of momentum can be reduced to the conventional Navier–Stokes equation. Because magnetic force is a body force, ferrofluid flow in microchannel should have the same velocity distribution as a pressure-driven flow.
The first term in (2.189) shows that the magnetic force is a body force, which is proportional to the volume. According to the scaling law, or the so-called cube-square law, magnetic force will be dominated by viscous force in microscale. However, the second term in (2.189) may have advantages in microscale due to the high-magnetic-field gradient that is achievable with integrated microcoils. As mentioned previously, ferrofluid with low Curie temperature is readily available. Magnetization can be controlled by adjusting the temperature from room temperature to an acceptably low Curie temperature. The temperature dependence of magnetization can be implemented in the first term of (2.189); the magnetic force then has the form [40]:
(2.190)
B9781437735208000024/si274.gif is missing
It is clear from (2.190) that the high-temperature gradient B9781437735208000024/si275.gif is missing in microscale can be another advantage for driving ferrofluid in microchannels. Thus besides microcoils, microheaters can be another tool for controlling ferrofluid-based micromixers.

2.7.1.1. Electromagnetic effects

Electromagnetic effect or magnetohydrodynamics (MHD) deals with behavior of electrically conducting fluids in a magnetic field. A magnetic field induces currents in a moving conductive fluid. A current passing through a conductive fluid can create forces on the fluid and affect the magnetic field. Similar to electrokinetics, MHD effects represent multiphysics problems, which require the coupling of the different fields. MHD effects can be described by the Navier–Stokes equations of fluid dynamics and Maxwell’s equations of electromagnetism.
The Navier–Stokes equation of an MHD flow has the form:
(2.191)
B9781437735208000024/si276.gif is missing
where v is the velocity vector, B is the magnetic field of flux density, and J is the current density. The term J×B represents the Lorentz force. The relation between the current density field, the electric field, the velocity field, and the magnetic field is:
(2.192)
B9781437735208000024/si277.gif is missing
where Eel is the electric field and Ψ is the electric potential.

2.8. Scaling Law and Fluid Flow in Microscale

The diffusion coefficient D, kinematic viscosity v, and the thermal diffusivity α=kpc – where k, p, and c are thermal conductivity, density, and specific heat, respectively – are transport properties and all have the same unit of m2/s. The ratios between these properties represent a group of nondimensional numbers that are characteristic for the interplay between the competing transport processes. These nondimensional numbers help to compare molecular diffusion with other transport processes in microfluidics.
The Schmidt number is the ratio between momentum transport and diffusive mass transport:
(2.193)
B9781437735208000024/si278.gif is missing
For most liquids and gases, the Schmidt number is larger than unity Sc1. This means in most cases spreading fluid motion is easier than molecules of the species.
Lewis number is the ratio between heat transport and diffusive mass transport:
(2.194)
B9781437735208000024/si279.gif is missing
The ratio between advective transport and momentum transport is called the Reynolds number:
(2.195)
B9781437735208000024/si280.gif is missing
where B9781437735208000024/si29.gif is missing is the mean velocity in the flow direction and Lch is the characteristic length of the considered channel. In many cases, the hydraulic diameter of the channel is taken as the characteristic length Lch. For typical values of Lch=100μm, B9781437735208000024/si29.gif is missing=1mm/s, and v=10−5cm2/s, the typical Reynolds number is Re=0.01. This small number means that laminar flow exists in almost all microfluidic applications.
Peclet number is the ratio between advective mass transport and diffusive mass transport:
(2.196)
B9781437735208000024/si281.gif is missing
The ratio between the Peclet number and the Reynolds number is actually the ratio between momentum transport and diffusive mass transport or the Schmidt number above. For diffusion coefficients ranging from 10−5m2s−1 to 10−7cm2s−1, the Peclet numbers for the typical values in the above example are 100<Pe<10,000. This means, advective mass transport dominates over diffusive transport in almost all microfluidic applications.
The average diffusion time t over the characteristic mixing length Lmixing, also called the striation thickness, is represented by the Fourier number [7]:
(2.197)
B9781437735208000024/si282.gif is missing
The Fourier number is usually in the range between 0.1 and 1. For a simple T-mixer with two streams in a microchannel of a length Lmixer and a width of W, the residence time should be the same as the average diffusion time:
(2.198)
B9781437735208000024/si283.gif is missing
Thus, the ratio between the channel length and channel width is:
(2.199)
B9781437735208000024/si284.gif is missing
For the above typical values of Fourier number 0.1<Fo<1 and Peclet number based on the channel width 10<PeW<10,000, the range of this ratio is 10<Lmixer/W<10,000. For some applications, the required mixing channel is unacceptably long.
If the inlets are split and rejoined as n pairs of solute/solvent streams, the mixing length is reduced to Lmixing=W/n. The ratio of the required channel length and channel width then becomes:
(2.200)
B9781437735208000024/si285.gif is missing
This concept is called parallel lamination where the channel length can be reduced by a factor of n2.
If the inlets are stretched and folded in n cycles, the mixing length is reduced to Lmixing=W/bn. The base b depends on the type of mixer. In the case of sequential lamination as discussed in the next section, the base is, for instance, b=2. The base could have a different value in the case of mixing based on chaotic advection. The ratio between the required channel length and the channel width is:
(2.201)
B9781437735208000024/si286.gif is missing
The above equation reveals that a very compact micromixer can be designed using sequential lamination or chaotic advection.
In general, fast mixing can be achieved with smaller mixing path and larger interfacial area. If the channel geometry is very small, the fluid molecules collide most often with the channel wall and not with other molecules. In this case, the diffusion process is called Knudsen diffusion [7]. The ratio between the distance of molecules and the channel size is characterized by the dimensionless Knudsen number:
(2.202)
B9781437735208000024/si287.gif is missing
where λ is the mean free path and Dh is the hydraulic diameter of the channel structure. The mean free path for gases is given by (see Section 2.1.1):
(2.203)
B9781437735208000024/si288.gif is missing
where kB=1.38066×10−23J/K is the Boltzmann constant, T is the absolute temperature, p is the pressure, and σm is the molecular diameter of the diffusing species. The Knudsen number for liquid is small, because the mean free path of liquid is on the order of a few angstroms. Thus, Knudsen diffusion may occur only in pores with nanometer sizes. In gases, the mean free path is on the order of a hundred nanometers to several micrometers. For example, at room condition, the mean free path of hydrogen is 0.2μm. Knudsen diffusion may occur in microchannels with diameters on the order of a few micrometers.
Among the above dimensionless numbers, Reynolds number Re represents the flow behavior in the microchannel, while Peclet number (Pe) represents the ratio between advection and diffusion. Thus, these two numbers are suitable for characterizing the operation point of a micromixer. From the definitions (2.195) and (2.196), the relation between Pe and Re is:
(2.204)
B9781437735208000024/si289.gif is missing
where B9781437735208000024/si290.gif is missingD, v, and Sc are the mean velocity, the diffusion coefficient, the kinematic viscosity, and the Schmidt number (2.193), respectively. The hydraulic diameter Dh and the mixing path Lmixing are usually on the same order; therefore, we can assume Lmixing/Dh1. The kinematic viscosity of liquids and the diffusion coefficient is on the order of v=10−6m2/s, while the diffusion coefficient ranges from D=10−9m2/s to D=10−11m2/s. The Schmidt number is about 103<Sc<105. On a Pe–Re diagram, the area between the two lines Pe1000Re and Pe100,000Re represents the operation range of micromixers. Operation points of micromixers are expected to be in this area.
In micromixers, the process of mixing and chemical reaction are related. Initially, mixing occurs first and is then followed by the chemical reaction. Subsequently, both mixing and chemical reaction occur in parallel. The ratio between the characteristic mixing time tmixing and reaction time treaction is called the Damköhler number:
(2.205)
B9781437735208000024/si291.gif is missing
A small Damköhler number means reaction is much slower than mixing. Thus, the reaction rate is determined by reaction tmixing. A large Damköhler number means reaction is faster than mixing. The extent of mixing determines the rate of reaction.
References
[1] Hirschfelder, J.C.; Gurtiss, J.C.; Bird, R.M., Molecular Theory of Gases and Liquids. (1954) Wiley, New York.
[2] Bird, R.B.; Stewart, W.E.; Lightfoot, E.N., Transport Phenomena. second ed (2001) Wiley, New York.
[3] Garcia, A.L., Numerical Methods for Physics. second ed (2000) Prentice Hall, Englewood Cliffs.
[4] Wu, Z.; Nguyen, N.T., Convective-diffusive transport in parallel lamination micromixers, Microfluidics and Nanofluidics 1 (2004) 208217.
[5] Bown, W., Brownian motion sparks renewed debate, New Scientist 133 (1992) 25.
[6] R. Brown, A brief account of microscopical observations made in the months of June, July and August, 1827, on the Particles contained in the pollen of plants; and on the general existence of active molecules in organic and inorganic bodies, Philosophy Magazine (4) 161–173, 1828.
[7] Cussler, E.L., Diffusion: Mass Transfer in Fluid Systems. second ed (1997) Cambridge University Press.
[8] Tayor, G.I., Diffusion by continuous movements, Proceedings of the London Mathematical Society vol. A20 (1921) 196.
[9] Einstein, A., Investigation on the Theory of Brownian Movement. (1956) Dover, New, York.
[10] Taylor, G.I., Dispersion of soluble matter in solvent flowing slowly through a tube, Proceedings of the Royal Society A 219 (1953) 186203.
[11] Brenner, H.; Edwards, D.A., Macrotransport Processes. (1993) Butterworth-Heinemann, Boston.
[12] Aris, R., On the dispersion of a solute in a fluid flowing through a tube, Proceedings of the Royal Society A 235 (1956) 6777.
[13] Aris, R., On the dispersion of a solute in a pulsating flow through a tube, Proceedings of the Royal Society A 259 (1960) 370376.
[14] Fife, P., Singular perturbation problems whose degenerate forms have many solutions, Applicable Analysis 1 (1972) 331358.
[15] Fife, P.; Nicholes, K., Dispersion in flow through small tubes, Proceedings of the Royal Society A 344 (1975) 131145.
[16] Gill, W.N., A note on the solution of transient dispersion problems, Proceedings of the Royal Society A 298 (1967) 335339.
[17] Gill, W.N.; Sankarasubramanian, R., A note on the solution of transient dispersion problems, Proceedings of the Royal Society A 316 (1970) 341350.
[18] Yamanaka, T., Projection operator theoretical approach to unsteady convective diffusion phenomena, Journal of Chemical Engineering of Japan 16 (1983) 2935.
[19] Watt, S.D.; Roberts, A.J., The accurate dynamic modelling of contaminant dispersion in channels, SIAM Journal of Applied Mathematics 55 (1995) 10161038.
[20] Johns, L.E.; Degance, A.E., Dispersion approximations to the multicomponent convective diffusion equation for chemically active systems, Chemical Engineering Science 30 (1975) 10651067.
[21] Yamanaka, T.; Inui, S., Taylor dispersion models involving nonlinear irreversible reactions, Journal of Chemical Engineering of Japan 27 (1994) 434435.
[22] Bloechle, B.W., On Taylor Dispersion of Reactive Solutes in a Parallel-Plate Fracture-Matrix Syste. PhD Thesis (2001) University of Colorado.
[23] Dutta, D.; Ramachandran, A.; Leighton Jr., D.T., Effect of channel geometry on solute dispersion in pressure-driven microfluidic systems, Microfluidics and Nanofluidics 2 (2006) 275290.
[24] Ajdari, A.; Bontoux, N.; Stone, H.A., Hydrodynamic dispersion in shallow microchannels: the effect of cross-section shape, Analytical Chemistry 78 (2006) 387392.
[25] Aref, H., Stirring by chaotic advection, Journal of Fluid Mechanics 143 (1984) 121.
[26] Wiggins, S.; Ottino, J.M., Foundations of chaotic mixing, Proceedings of the Royal Society A 362 (2004) 937970.
[27] Lorenz, E.N., Deterministic nonperiodic flow, Journal of the Atmospheric Sciences 20 (1963) 130141.
[28] Dean, W.R., Note on the motion of a fluid in a curved pipe, Philosophy Magazine 4 (1927) 208223.
[29] Berger, S.A.; Talbot, L.; Yao, L.S., Flow in curved pipes, Annual Reviews of Fluid Mechanics 15 (1983) 461512.
[30] Gratton, M.B., The Effects of Torsion on Anomalous Diffusion in Helical Pipe. Thesis (2002) Department of Mathematics, Harvey Mudd College.
[31] Germano, M., On the effect of torsion on helical pipe flow, Journal of Fluid Mechanics 203 (1982) 18.
[32] Jones, S.W.; Thomas, O.M.; Aref, H., Chaotic advection by laminar flow in a twisted pipe, Journal of Fluid Mechanics 209 (1989) 335367.
[33] Hadamard, J., Movement permanent lent D’une sphere liquide et visqueuse dans un liquide visqueux, Comptes Rendus 152 (1911) 1735.
[34] Stone, Z.B.; Stone, H.A., Imaging and quantifying mixing in a model droplet micromixer, Physics of Fluids 17 (2005) 111.
[35] Grigoriev, R.O.; Schatz, M.F.; Sharma, V., Chaotic mixing in microdroplets, Lab Chip 6 (2006) 13691372.
[36] Rodd, L.E.; Scott, T.P.; Boger, D.V.; Copper-White, J.J.; McKinley, G.H., The inertio-elastic planar entry flow of low-viscosity elastic fluids in micro-fabricated geometries, Journal of Non-Newtonian Fluid Mechanics 129 (2005) 122.
[37] Chen, C.H.; Lin, H.; Lele, S.K.; Santiago, J.G., Convective and absolute electrokinetic instability with conductivity gradients, Journal of Fluid Mechanics 524 (2005) 263303.
[38] Levich, V.G., Physicochemical Hydrodynamics. (1962) Prentice-Hall, Englewood Cliffs.
[39] Rosenzweig, R.E., Ferrohydrodynamics. (1985) Cambridge University Press, Cambridge.
[40] Love, L.J.; Jansen, J.F.; McKnight, T.E.; Roh, Y.; Phelps, T.J., A magnetocaloric pump for microfluidic applications, IEEE Transactions on Nanobioscience 3 (2004) 101110.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.135.191.86