· web viewtwo newtonian flows with similar flow rates and physicochemical properties are...

42
Tracking time-dependent production in electrokinetic Y-microreactors Hamed Helisaz a , Masoud Babaei b , Arman Sadeghi c,1 a Center of Excellence in Energy Conversion (CEEC), School of Mechanical Engineering, Sharif University of Technology, Tehran 11155-9567, Iran b School of Chemical Engineering & Analytical Science, University of Manchester, Manchester M13 9PL, United Kingdom c Department of Mechanical Engineering, University of Kurdistan, Sanandaj 66177- 15175, Iran Abstract We perform a theoretical study on the transient reaction-diffusion kinetics in an electrokinetic Y-shaped microreactor. The flow is assumed to be both steady and fully developed. The governing equations are solved in dimensionless form utilizing a 3D finite-volume based numerical algorithm, assuming a second-order irreversible reaction between the components. Analytical solutions are also obtained for cross-stream diffusion without reaction under a uniform velocity distribution. It is shown that the well-known butterfly-shaped form of 1 Corresponding author E-mail addresses: [email protected] (H. Helisaz), [email protected] (M. Babaei), [email protected] (A. Sadeghi)

Upload: others

Post on 21-Feb-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

Tracking time-dependent production in

electrokinetic Y-microreactors

Hamed Helisaz a, Masoud Babaei b, Arman Sadeghi c,1

a Center of Excellence in Energy Conversion (CEEC), School of Mechanical Engineering, Sharif University

of Technology, Tehran 11155-9567, Iran

b School of Chemical Engineering & Analytical Science, University of Manchester, Manchester M13

9PL, United Kingdom

c Department of Mechanical Engineering, University of Kurdistan, Sanandaj 66177-15175, Iran

Abstract

We perform a theoretical study on the transient reaction-diffusion kinetics in an electrokinetic Y-shaped

microreactor. The flow is assumed to be both steady and fully developed. The governing equations are

solved in dimensionless form utilizing a 3D finite-volume based numerical algorithm, assuming a second-

order irreversible reaction between the components. Analytical solutions are also obtained for cross-stream

diffusion without reaction under a uniform velocity distribution. It is shown that the well-known butterfly-

shaped form of the production concentration profile is not immediately created and it is established only

after the system is sufficiently close to its steady-state. Furthermore, the inclination of the concentration

peak toward the component of lower diffusivity or inlet concentration is less significant at the earlier stages

of the production. Finally, it is demonstrated that the short-term influence of the parameters affecting the

advection of mass on the total efficiency, defined as the ratio of the total production to the amount of the

limiting reactant within the device, is quite the opposite of that at the steady-state. That is, whereas

increasing each of the dimensionless Debye-Hückel parameter, the pressure-driven velocity to

1 Corresponding authorE-mail addresses: [email protected] (H. Helisaz), [email protected] (M. Babaei), [email protected] (A. Sadeghi)

Page 2:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

electroosmotic velocity ratio, and the Péclet number leads to larger short-term efficiencies, the opposite is

true at the steady-state.

Keywords: Electroosmotic flow, Lab-on-a-chip, Microreactor, Numerical modeling, Analytical solution

1. Introduction

The micro accelerometer developed in 1979 at Stanford University marked the beginning of numerous

attempts to replace the conventional large laboratories with their micro counterparts referred to later on as

lab-on-a-chip (LOC) devices [1]. As it is implied by the name, LOCs are miniaturized laboratory platforms

which are machined on a wafer of glass or silicon and are able to perform one or several laboratory

functions quickly and simultaneously. Less consumption of normally expensive samples and reagents, large

surface-to-volume ratio, and high controllability are among the main advantages of LOC devices attracting

the attention of many researchers.

LOCs are usually fabricated for a certain purpose. Therefore different devices consist of different

components. However microreactors are necessary in more or less all LOC devices due to prevalence of

mixing/reaction processes in most of laboratory functions. In their most basic design, microreactors take a

T- or Y-shaped form [2] in which fluids are entered by two arms and introduced into a chamber wherein

components are mixed together and a chemical reaction occurs among them. This form has been placed as a

base for developing other designs, such as parallel and serial lamination [3, 4], which enhance the mixing

and reaction processes. Nevertheless, considering these designs is indeed accompanied by not only an

increasing fabrication cost, but also difficulties in integrating microreactors into LOC platforms; as a result,

much research attention is still given to typical T/Y-shaped microreactors. The pioneering studies on

microscale reactors include the work of Salimi-Moosavi [5] and that of Kamholz et al. [6], both conducted

in 1990s, in which microreactors are utilized for synthesizing an azo dye and calculating a reaction rate

constant, respectively. The succeeding research works on microreactors focused on more complicated

Page 3:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

applications, including diagnostic platforms [7-10], drug development [11-14], and nanoparticle synthesis

[15-18].

Different methods have been proposed for creating fluid flow in microstructures [19] amongst which

electroosmotic flow (EOF) is used in the growing number of microfluidic applications. This is largely

thanks to EOF’s favorable characteristics such as obviating the need for moving parts and creating pulse-

free plug flows [20]. In electroosmosis, the interaction of the electrolyte with polarized dielectric walls

form an accumulation of ions of the same electric charge adjacent to the walls. Applying an external

electric field then makes the ions move toward the electrode of opposite charge, thereby creating a net flow

upon the interaction of the ions and the solvent molecules. Since late 1990s, when the first studies on EOF-

based microsystems were performed [21, 22], electroosmosis has been considered in many microsystems

either as the only driving force [23-28] or accompanied by pressure force [29-34]. In these works, however,

more attention was paid to micromixers rather than microreactors. Few available researches on

microreactors include the experimental works of Wilson and McCreedy [35] and Watts et al. [36] whose

results proved more efficient reaction by adopting EOF-based microreactors instead of large-scale reactors.

More recently, Sharma et al. [24] and Yousefian and Saidi [37] presented theoretical

results of species transport in EOF-based catalytic microreactors.

Both mixing and reaction processes require time to complete, especially in passive microchannels

wherein flows are laminar and so mixing occurs via the typically slow mechanism of molecular diffusion

[38]. Given the transient time in current commercial micromixers/microreactors [39], it is not surprising to

see the transient time of a microreactor lasting for tens of seconds. Therefore, thorough understanding of

diffusion-reaction mechanism in microreactors is not developed unless both its transient and steady-state

behaviors are investigated. In our previous work [32], for the first time, a comprehensive parametric study

on steady-state responses of diffusion-reaction mechanism was performed by 3D finite-volume based

numerical modeling of an EOF-based Y-microreactor. Since, to our best knowledge, no time-dependent

study has yet been conducted on diffusion-reaction mechanism in T/Y-shaped microreactors, the present

study aims to fill this gap in the literature by extending our recent study to transient conditions. We

Page 4:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

investigate the effects of various operational parameters on the performance of the microreactor before

reaching the steady-state. It is attempted to broaden the scope of research by eliminating the simplifying

assumptions made in the preceding studies. For example, the flow field, which is created by the

simultaneous action of electroosmotic and pressure forces, is evaluated without imposing any additional

approximation like Helmholtz-Smoluchowski approximation. Additionally, no predefined values are

assumed for the ratios between diffusivities and inlet concentrations of the components. We believe the

present study together with our previous work [32] pave the way for a better design of the future

microreactors.

2. Problem formulation

Consideration is given toward mass transport in an electrokinetic-based Y-shaped microreactor from the

onset of reaction to steady-state. Fig. 1 provides top and front views of the microreactor under

consideration which is comprised of two arms coinciding together at the origin of the coordinate system.

Two Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms

under the combined action of electroosmotic and pressure forces. The fluids, one containing reactant A at a

number concentration of c0 , A and the other containing reactant B at a concentration of c0 , B, run side-by-

side in a chamber wherein they mix together and a second-order non-reversible chemical reaction

(A+B→C) occurs among them. The solute-liquid solutions are considered dilute enough to allow

neglecting the influences of the diffusion-reaction process on the flow field which is assumed to be steady.

Thanks to the intrinsic low Reynolds number of microflows, the flow in the main channel can be

considered hydrodynamically fully developed in the presence of a uniform wall zeta potential. The last

premise is that the Debye length is much smaller than the minimum channel dimension so that there is no

EDL overlap. It should be also pointed out that, taking advantage of symmetry, we solve the flow domain

only for the channel’s first quarter and we solve concentration domain for the channel upper half to reduce

the computational costs.

Page 5:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

2.1. Electrical potential distribution

For evaluation of the electroosmotic velocity, it is first required to predict the volumetric body force which

is dependent upon the electrostatic potential. The total electrical potential φ is related to the net ionic charge

in the channel ρe through the Poisson equation, given as

∇2 φ=−ρe

ε(1)

in which ε denotes the fluid permittivity. The electric potential in our system includes the first part

generated by the external electric field E x and the second part induced by electric charges residing on the

channel walls. Reminding that the electric field is uniform, the first part is a linear function of only x and it

can be calculated by considering the reference value of ϕ0 atx=0. The second part can be considered to be

a function of y andz, denoted here by ψ ( y , z ), since the zeta potential throughout the channel walls is

uniform and the flow is fully developed. Therefore, the following formula for the electric potential is

obtained

φ ( x , y , z )=( ϕ0−x E x )+ψ ( y , z ) (2)

Invoking the Boltzmann distribution, the net ionic charge can be related to the ionic concentration at neutral

conditionsn∞, proton chargee, Boltzmann constantk B, and absolute temperature T as

ρe=−2e Z n∞ sinh ( e Z ψk BT ) (3)

Note that this equation is obtained by assuming a symmetric electrolyte of valence Z [40]. Substituting Eqs.

(2) and (3) into Eq. (1), it may be rewritten as

∂2ψ∂ y2 + ∂2ψ

∂ z2 =2 eZ n∞

εsinh ( e Z ψ

kB T ) (4)

Page 6:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

To generalize, Eq. (4) is made dimensionless by defining y¿= y / H ,z¿=z /H , ψ¿=e Z ψ /k B T , and

Κ=H / λD in which λD=(2 n∞ e2 Z2/ε kB T )−1 /2 indicates the Debye length. We will then have

∂2ψ¿

∂ y¿2 + ∂2 ψ¿

∂ z¿2 =Κ2 sinhψ¿ (5)

The solution to ψ¿ must satisfy the following four boundary conditions

∂ψ¿

∂ y¿|y¿=0

= ∂ ψ¿

∂ z¿ |z ¿=0

=0 ,ψ¿ ¿y ¿=1=ψ¿¿z ¿=α=ζ ¿(6)

wherein α=W /H is the channel aspect ratio and ζ ¿=e Z ζ /k B T is the dimensionless zeta potential.

2.2. Velocity distribution

For evaluation of the velocity field, the mathematical representation of the momentum conservation law

should be solved. For a laminar incompressible flow of Newtonian fluids with constant physical properties,

the general form of the momentum equation reads

ρ D uDt

=−∇ p+μ∇2u+ f (7)

wherein t , ρ, p, μ, and u represent the time, density, pressure, dynamic viscosity, and velocity vector,

respectively. Moreover, f is the body force vector reflecting the interaction of the electric field with the free

ions within the solution. Reminding that the flow is assumed steady and fully developed, the left hand side

of Eq. (7) vanishes. The momentum conservation equation, therefore, can be solved for velocity since both

the applied pressure gradient ∇ p and the body force f are known; particularly in the streamwise direction,

∂ p/∂ x=d p /d x which is known a priori and f x= ρe Ex which can be calculated with the aid of the

electric potential solution. Accordingly, the momentum conservation equation in the axial direction reduces

to

Page 7:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

0=−d pd x

+μ( ∂2 ux

∂ y2 +∂2 ux

∂ z2 )+ ρe Ex (8)

wherein ux is the axial velocity. This equation is made dimensionless by scaling the velocity asu¿=ux /uHS

with uHS=−εζ Ex /μ being the Helmholtz-Smoluchowski velocity. Since there is a pressure-driven

contribution to the fluid velocity, the parameter Γ=uPD /uHS is required to quantify the ratio of the

pressure-driven velocity scale uPD=−H2 (d p /d x ) /2 μ to the electroosmotic velocity scale uHS. Based on

the newly defined quantities, Eq. (8) modifies into

∂2u¿

∂ y¿2 +∂2 u¿

∂ z¿2 =−2 Γ−Κ2

ζ ¿ sinhψ¿ (9)

The four necessary boundary conditions for Eq. (9), which are provided by symmetry and no-slip

conditions, take the forms

∂u¿

∂ y¿|y ¿=0

=∂ u¿

∂ z¿|z¿=0

=0 , u¿¿ y¿=1=u¿¿z ¿=α=0 (10)

2.3. Concentration distribution

Generally, advection, diffusion, and reaction are mechanisms responsible for species transport. Here, we

intend to study the effects of these mechanisms over time and so we need to consider the transient form of

reaction-advection-diffusion equation for each species, that is

∂ c A

∂t+u

∂ c A

∂ x=DA ∇2 c A−k cA cB (11)

∂ c A

∂t+u

∂ cB

∂ x=DB ∇2 cB−k c A cB (12)

∂ c A

∂t+u

∂ cC

∂ x=DC ∇2 cC+k c A cB (13)

Page 8:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

In the above equations, c andD represent, respectively, the number concentration and diffusion coefficient

of the species whose type is shown as subscript. The terms including the chemical rate constant k are

reaction terms written based on the law of mass action [41]. Eqs. (11) to (13) can be scaled by introducing

new dimensionless parameters as x¿= x /H , β=DA / DB, γ=DA / DC, c¿=c /c0 , A, and t ¿=t DA/ H2.

Substituting the already defined parameters into Eqs. (11) to (13), there appear two well-known

dimensionless quantities namely the Péclet numberPe=uHS H / DA and Damköhler number

Da=Hk c0 , A/uHS. The dimensionless forms of Eqs. (11) to (13), hence, become

∂ c A¿

∂t ¿ +Pe u¿ ∂ c A¿

∂ x¿ =( ∂2 cA¿

∂ x¿2 +∂2 c A

¿

∂ y¿2 +∂2 c A

¿

∂ z¿2 )−PeDac A¿ cB

¿ (14)

∂ cB¿

∂ t¿+ βPeu¿ ∂c B

¿

∂ x¿ =( ∂2 cB¿

∂ x¿2 +∂2c B

¿

∂ y¿2 +∂2 cB

¿

∂ z¿2 )−βPeDac A¿ cB

¿ (15)

∂ cC¿

∂ t ¿ +γPeu¿ ∂ cC¿

∂ x¿ =( ∂2 cC¿

∂ x¿2 +∂2 cC

¿

∂ y¿2 +∂2cC

¿

∂ z¿2 )+γPeDa cA¿ cB

¿ (16)

Eqs. (14) to (16) require the following boundary and initial conditions to be solved

∂2c A ,B ,C¿

∂ x¿2 |x¿=l

=∂c A ,B ,C

¿

∂ y¿ |y¿=0,1

=∂ c A ,B ,C

¿

∂ z¿ |z¿=± α

=0

(17)c A

¿|x¿=0={ 1 z¿≥ 0¿0 z¿<0

, cB¿|x ¿=0={ 0 z¿>0

¿1/ω z¿≤0, cC

¿|x¿=0=0

c A, B , C¿ |t¿=0=0

wherein l=L/ H and ω=c0 , A/c0 , B.

2.4. Numerical method

The governing equations in dimensionless form are solved by adopting the finite volume method in which

the first step is generating the grid structure. Considering the problem physics, we expect higher electric

potential and velocity gradients at the vicinity of the walls; the opposite is true for solute concentration as

Page 9:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

chemical reaction produces intense concentration gradients at the fluid-fluid interface. Therefore, the grid

structure for solving Eqs. (5) and (9) should be different from that of Eqs. (14) to (16), not to mention that

for the latter the grids should also be clustered near the channel entrance so as to capture high gradients

therein. To avoid the computational difficulties of dealing with non-uniform grid structures, we transform

the current coordinatesx¿, y¿, andz¿ into new ones named x, y, and z, along which grid nodes are uniformly

distributed, by applying the same transformations as those given in our previous work [32]. Adopting this

new coordinate system, Eqs. (5), (9), and (14) to (16) are converted to

Q12 ( y ) ∂2ψ¿

∂ y2 +Q2 ( y ) ∂ ψ¿

∂ y+Q3

2 ( z ) ∂2ψ¿

∂ z2 +Q4 ( z ) ∂ ψ¿

∂ z=Κ 2sinh ψ¿

(18)

Q12 ( y ) ∂2u¿

∂ y2 +Q2 ( y ) ∂u¿

∂ y+Q3

2 ( z ) ∂2 u¿

∂ z2 +Q4 ( z ) ∂ u¿

∂ z=−2 Γ− Κ2

ζ ¿ sinh ψ¿(19)

∂ c A¿

∂t ¿ +Pe u¿Q5 ( x )∂ c A

¿

∂ x=[Q 5

2 ( x )∂2 c A

¿

∂ x2 +Q6 ( x )∂ cA

¿

∂ x+

∂2 c A¿

∂ y2 +Q72 ( z )

∂2c A¿

∂ z2 +Q8 ( z)∂ cA

¿

∂ z ]−PeDac A¿ c B

¿ (20)

∂ cB¿

∂ t¿+ βPeu¿Q5 ( x )

∂ c B¿

∂ x=[Q5

2 ( x )∂2c B

¿

∂ x2 +Q6 ( x )∂ cB

¿

∂ x+

∂2c B¿

∂ y2 +Q72 ( z )

∂2 cB¿

∂ z2 +Q8 ( z )∂ cB

¿

∂ z ]−βPeDa cA¿ cB

¿ (21)

∂ cC¿

∂ t ¿ +γPeu¿Q5 ( x )∂ cC

¿

∂ x=[Q5

2 ( x )∂2 cC

¿

∂ x2 +Q6 ( x )∂ cC

¿

∂ x+

∂2 cC¿

∂ y2 +Q72 ( z )

∂2 cC¿

∂ z2 +Q8 ( z )∂ cC

¿

∂ z ]+γPeDa c A¿ cB

¿ (22)

in which the functions Q1 ,.. , 8 are given in the Appendix. Now, it is time to integrate the equations over each

volume segment and then place the variables at cell centers by practicing the power-law scheme; the

transient terms are also handled by integrating Eqs. (20) to (22) implicitly over the time. This process

introduces a set of algebraic equations each of the governing equations. The algebraic equations

corresponding to Eqs. (18) and (19) have already been introduced in our previous work [32] and those

pertinent to Eqs. (20) to (22) are as follows

aP , A c A,i , j , k¿ −aW , A c A ,i−1 , j , k

¿ −aE , A c A , i+1 , j ,k¿ −aB , A c A ,i , j−1, k

¿ −aT , A c A ,i , j+1, k¿ −aS , A c A, i, j , k−1

¿ −aN , A c A,i , j , k+1¿ =SA(23)

Page 10:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

aP , B cB, i , j , k¿ −aW , B cB,i−1 , j ,k

¿ −aE , B cB, i+1 , j ,k¿ −aB ,B cB ,i , j−1 ,k

¿ −aT , B cB , i , j+1 ,k¿ −aS ,B c B,i , j ,k−1

¿ −aN , B cB ,i , j , k+1¿ =SB(24)

aP , CcC , i, j , k¿ −aW ,C cC ,i−1 , j ,k

¿ −aE ,C cC ,i+1 , j , k¿ −aB, C cC , i , j−1 , k

¿ −aT ,C cC ,i , j+1 , k¿ −aS , C cC , i, j , k−1

¿ −aN , C cC , i , j , k+1¿ =SC

(25)

Here, i, j, and k denote the location of volume blocks in x, y, and z directions, respectively. As seen, Eqs.

(23) to (25) relate the variables at each central cell P to those at the south, north, west, east, bottom, and top

neighbor cells denoted respectively by S, N , W , E, B, and T . Note that, because of considering implicit

time integration, the values of variables are all calculated at time levelt+∆ t and the results of the previous

time step are reflected only in the source terms S which are listed in the Appendix along with the other

coefficients appeared in the equations. The algebraic equations (23) to (25) are solved by applying the Tri-

Diagonal Matrix Algorithm (TDMA): we start from the electric potential equation whose solution is

required for solving velocity field. The velocity equation is then recalled and its results are extracted for

solving the mass transport equations (concentration fields). It is noteworthy that, because of using different

grid systems for the velocity and concentration fields, the velocity values at the concentration grid points is

obtained utilizing the cubic spline interpolation, prior to their use in the concentration fields. Ultimately,

Eqs. (23) to (25) are solved to get the concentrations of components A, B, and C at desired locations and

times.

For validation of the numerical method developed, the results for t → ∞ were compared with those of

our previous study [32] in which the steady-state of the present problem was investigated, whereby a very

good agreement was observed. This validation, however, does not necessarily confirm the transient

responses of our numerical algorithm; these responses are verified by using an analytical solution

developed for a special case in the next section.

2.5. Analytical solution for u¿=1 and Da=0

In this section, we intend to present an analytical solution for the problem by considering two more

assumptions including

Page 11:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

EDLs are so thin that the velocity over the majority of the channel cross-sectional area equals uHS (

u¿ (x , y , z )=1).

No chemical reaction occurs among the components, that is Da=0, implying that the inlet components

only mix together.

The first assumption obviates the need for solving the electric potential and velocity equations; it also

reduces our problem to a two-dimensional one dependent only on x and zcoordinates [42]. Moreover, the

reaction terms in mass transport equations vanish by virtue of the second assumption; hence Eqs. (14) and

(15) are no longer and can be handled separately. Here, we develop our analytical model only for c A¿ as

exactly the same procedure can be followed for cB¿ . Based on the new premises, Eq. (14) simplifies to

∂ c A¿

∂t ¿ +Pe∂ c A

¿

∂ x¿ =∂2 cA

¿

∂ x¿2 +∂2 c A

¿

∂ z¿2 (26)

We express c A¿ as the summation of a steady-state component c ( x¿ , z¿ ) and a transient component

c ( x¿ , z¿ , t ¿), governed respectively by the following equations

Pe ∂ c∂ x¿=

∂2 c∂ x¿2 +

∂2 c∂ z¿2

(27)

∂ c∂ t¿

+Pe ∂ c∂ x¿=

∂2 c∂ x¿2 +

∂2 c∂ z¿2 (28)

whose solutions must conform to the following boundary conditions

∂2 c∂ x¿2|

x¿=l

= ∂2 c∂ x¿2|

x ¿=l

= ∂ c∂ z¿|

z ¿=± α= ∂ c

∂ z¿|z¿=± α

=0

(29)

c|x¿=0={ 1 z¿≥0¿0 z¿<0

, c|x ¿=0=0 , c|t ¿=0=−c

Eqs. (27) and (28) are solved invoking the separation of variables method; the analytical solution of the

former satisfying the boundary conditions (29) can be readily derived as

Page 12:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

c ( x¿ , z¿ )=12−∑

n=1

∞ 2nπ

sin(nπ2 )cos [nπ

2 ( z¿

α+1)]exp¿¿¿ (30)

wherein

δ−¿

+¿=Pe2

±√ Pe2+n2 π2

α2

2¿¿

(31)

Solving Eq. (28), however, poses more difficulties due to the presence of three independent variables. The

general solution of c satisfying the pertinent boundary conditions may be written as

c ( x¿ , z¿ , t ¿)=−2i∑m=1

A0 m exp( Pe2

x¿)sin( mπl

x¿)exp [−( Pe2

4+ m2 π2

l2 ) t ¿]−2i∑n=1

∑m=1

Anm exp( Pe2

x¿)sin(mπl

x¿)cos [ nπ2 ( z¿

α+1)]exp [−( Pe2

4+ m2 π2

l2 + n2 π2

4 α 2 ) t¿ ](32)

Applying the initial condition and following the orthogonality conditions, the coefficients A0m and Anm

appeared in Eq. (32) are calculated as

A0 m=−i

πm2 [1−cos ( πm) exp(−Pe

2l)]

Pe2 l2

4+m2 π2

(33)

Anm=−i 2 mn

sin( nπ2 )¿¿ (34)

Therefore, considering c A¿ =c+c, the final expression for c A

¿ is obtained as

c A¿ (t ¿ , x¿ , z¿)=1

2−∑

n=1

∞ 2nπ

sin( nπ2 )cos [nπ

2 ( z¿

α+1)]exp¿¿¿ (35)

Now that we have presented an analytical solution for mixing in electrokinetic microreactors at thin EDL

limit, it is convenient to quantify the degree of mixing. Different parameters are used in the literature to

quantify the degree of mixing in T/Y-shaped microchannels among which is the mixing index. The

mathematical representation of this parameter appropriate to the present analysis is given as [43]

I mix(x¿ , t¿)=2∫

−W

0

c A ( x , z , t ) d z

W c0 , A=

2∫−α

0

cA¿ ( x¿ , z¿ , t¿ ) d z¿

α c0 , A¿

(36)

Page 13:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

Substituting Eq. (35) into Eq. (36) provides

I mix(x¿ , t¿)=1−∑n=1

∞ 8n2 π 2 sin2( nπ

2 )exp¿¿¿ (37)

The predictions of Eq. (37) are compared with the results of the numerical simulations based on a uniform

fluid velocity in Fig. 2 at different x¿ and Pe. A good agreement is observed between the results, revealing

the correctness of the numerical procedure conducted; the small deviation between the results stems from

applying the no-slip boundary condition in the numerical code. That is why the error is smaller for a lower

x¿ at which there is more dominance of the diffusion over advection due to larger concentration gradients.

3. Results and discussion

In this section, we intend to investigate the reaction-diffusion behavior of components over time by

studying the effect of each governing parameter on the system transient response. The results are all

presented assumingα=2andl=20. To start with, the axial progress of the concentration zones over the

time is depicted in Fig. 3 where red and blue contours correspond respectively to components A and B. For

drawing this figure, the inlet concentrations of the reactants are assumed to be the same but the diffusion

coefficient of component A is twice that of B. Therefore, while A puts its effort into moving both in x and

z directions, B is more focused on moving along the streamwise direction and so, as observed in the figure,

it takes less time for B to pass the channel. Bearing in mind that the components are introduced with the

same flow rate, this implies an important conclusion: providing the same concentrations of the reactants in

the channel, which is crucial for efficient production, requires supplying more amount of the reactant of

lower diffusivity. It should be noted that the contours in Fig. 3 are primarily depicted for the mid-plane (

y¿=0¿ wherein the components are faster than those traveling adjacent to the wall; therefore, less progress

of the concentration zones is expected close to the walls as it is demonstrated by the dashed lines.

Fig .3 can be redrawn for the product of reaction to achieve Fig. 4 wherein the time development of C

concentration profile is shown at both x¿=7 and y¿=0. It is interesting that the expected butterfly-shaped

Page 14:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

profile, which has been explained in our previous work [32], does not appear immediately and C profile

firstly emerges in a spindle-like form. That is because in a short duration, slower axial movement of the

reactants traveling near the horizontal walls gives rise to less accumulation of A and B therein as compared

with the mid-plane; so less product is generated near the walls forming the spindle-like profile observed in

Fig. 4a. Giving more time, A and B find more opportunity to cover the entire channel and now the

difference between the residence times of the components shapes the C concentration profile: at the

centerline, where the velocity is maximum and the components have the least time to react with each other,

the production is minimum whereas in the near-wall regions the opposite is true and the production reaches

its maximum (Figs.4b and 4c).

Like Fig. 3, DA is assumed to be double of DB in Fig. 4 causing more penetration of component A into

B stream, thereby forcing the product concentration peak toward there. This inclination can be seen even in

Fig. 4a where reaction is in its beginning stages. Another point worth noting is that the upstream sections

obtain their steady-state in advance of the downstream locations. This can be readily recognized by

comparing Figs. 4b and 4c in which although the y-z profiles do not change, x-z profiles experience

significant alteration in the downstream region. In order to consider this matter in more depth, we plot the

transverse distributions of component C at the mid-plane for differentx¿andt ¿ in Fig. 5. As observed, the

concentration profile of C at x¿=5 does not change over time meaning that it has obtained its steady-state

before other axial positions, especially x¿=20in which the production has not even started at t ¿=0.1. The

larger transient times of the downstream sections let components located therein to participate more in

diffusion process, thereby increasing the near-wall production of C by moving toward the channel end. It is

exactly for the same reason that the inclination of the concentration peak is magnified at larger values of x¿

(Fig. 5c).

3.1. Effect of velocity on production

This section is dedicated to investigating the effect of the velocity profile on the system transient response.

The velocity here is assumed to be created by the combined action of the electroosmotic and pressure forces

Page 15:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

and, therefore, Κ ad Γ are the governing parameters specifying how fast the solutes are moving by the

flow. Based on the definition, Κ is inversely proportional to the Debye length and so the larger Κ is the

thinner EDL becomes. Accordingly, the electroosmotic body force is concentrated more near the wall for a

higher K, leading to a faster movement of the species, especially near the wall. The time development of

the transverse C concentration profile at the mid-plane for pure electroosmotic flow at thick and thin EDLs

are respectively studied in Figs. 6a and 6b. The two remaining parts of Fig. 6 are assigned to examine the

time development of the concentration profile at positive and negative values of Γ . Recalling its definition,

Γ measures the share of the pressure gradient force in the driving components and its sign indicates

whether the pressure gradient is favorable (Γ>0) or unfavorable (Γ<0) to electroosmotic force; therefore,

increasingΓ is equivalent to rising the fluid velocity.

The first point drawing attention in Fig. 6 is that the inclination of the concentration peak increases

with time. Moreover, by comparing the left and right parts of the figure, one can conclude that producing

the same amount of C at a fixed axial position takes much less time when the flow rate increases either by

shrinking EDL or applying a favorable pressure gradient. For example, whereas the profile of t ¿=0.25 in

Fig. 6b passes through c¿=0.18 at its peak, its counterpart in the LHS figure does not even exceed

c¿=0.05. This is firstly because charging the channel with reacting components takes more time when the

particles are slower, causing a delay in the start of production. Moreover, slower flow rates allow the

reacting components to spread throughout a larger area and so the aggregation of C at a given point takes

longer. In return for the decline in the production rate, introducing the components with a lower velocity

gives them a longer time to participate in the reaction-diffusion process and so leads to higher production at

the steady-state, particularly in the lateral regions.

3.2. Efficiency

One of the most important attributes of a microreactor, especially when it comes to application, is its total

production. This parameter, however, is dependent not only on the microreactor performance, but also on

the inlet concentration of the reactants; hence, it cannot be considered as the only factor for examining the

Page 16:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

microreactor performance. Therefore, in order to omit the effect of the inlet concentration of the reactants

on the total production, we measure the total amount of C with respect to the limiting reactant, that is to say

that the microreactor efficiency is calculated. Recalling that in this studyω≥ 1 for all the cases, component

B can always be treated as the limiting reactant and so the total efficiency of system in percent can be

written as

η (t )=∫−W

W

∫0

H

∫0

L

cC ( x , y , z , t ) d xd y d z

∫−W

W

∫0

H

∫0

L

cB (x , y , z ,t ) d x d yd z×100=

∫−α

α

∫0

1

∫0

l

cC¿ ( x¿ , y¿ , z¿ , t¿ ) d x¿d y¿ d z¿

∫−α

α

∫0

1

∫0

l

cB¿ ( x¿ , y¿ , z¿ ,t ¿) d x¿d y¿ d z¿

×100 (38)

Due to its importance, the efficiency is studied in Fig. 7 under the effect of all the governing

parameters, namelyΓ ,Κ ,β ,ω,Pe, and Da. As mentioned in the previous section, increasing the flow rate

diminishes the production in long-term and so it is not surprising that the efficiency grows by decreasing

either ΓorΚ . The supplementary point added by this figure is the extent to which each of these parameters

is effective: according to Figs. 7a and 7b, the efficiency seems to be more influenced by the pressure

gradient than EDL thickness. Consequently, applying an unfavorable pressure gradient should be

considered as the first hydrodynamic factor in designing efficient microreactors. This conclusion, however,

is reversed in short-term intervals: considering an unfavorable pressure gradient in Fig. 7a leads to the

lowest efficiency beforet ¿≅ 0.17. This is the immediate result of a higher production rate in the presence of

higher velocities which was pointed out to in the previous section. The opposite behavior of the system in

short-term duration is important especially when it is not allowed to obtain its steady-state; as an example,

for cases in which there is a time-dependent injection of the reacting components instead of a constant

injection rate.

The second row of Fig. 7 is dedicated to specifying the effect of component B on the system

performance for which the efficiency is plotted versus the diffusivity and inlet concentration of this reactant

in parts (c) and (d). It should be noted that although βandωare dependent on both A and B, changing them

at fixed Pe and Da only affects the properties of B component. Enhancing DB (decreasingβ) increases the

Page 17:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

region over which B can penetrate into A stream and so raises the production of C and efficiency. The RHS

figure, however, tells a more interesting story: increasingc0 , Braises the initial concentration of the limiting

reactant, which of course increases the production; nonetheless, this should not be interpreted as efficiency

enhancement since the production growth occurs solely at the expense of injecting more reactants. In fact,

the ratio of the production to the limiting reactant, that is the efficiency, is reduced by increasing c0 , B

because a lower number of A particles surround each B particle.

The two remaining figures in Fig. 7 are devoted to studying the influences of the Péclet and

Damköhler numbers on η. Considering the fact that only one governing parameter is permitted to vary in

each figure, the increase of the former means a decrease in DA while the latter is altered only through

changing the chemical rate constant. Accordingly, an increase in Pe will be accompanied by a lower cross-

stream diffusion of A particles, which enhances their overall streamwise movement; this leads to more

accumulation of the reacting components downstream and, ultimately, larger short-term efficiencies. At

long-term, however, higher cross-stream diffusion of component A provides a wider engagement of the

reacting components, this is to say that the efficiency is magnified by decreasing Pe (Fig. 7e). The last

figure we discuss here is Fig. 7f in which the impact of the Damköhler numberon the system efficiency is

demonstrated. Due to its definition,Da is directly proportional to the chemical rate constant and, therefore,

the observed monotone-increasing relation between Da and η is not strange.

4. Conclusions

A theoretical study was conducted to investigate the transient response of a Y-shaped microreactor. It was

assumed that the fluid flow, created under a combined influence of the pressure and electroosmotic forces,

is steady and fully developed. A second-order irreversible chemical reaction was assumed to take place

between the components. In general, the problem was treated numerically utilizing a 3D finite-volume-

based numerical code; however, analytical solutions were also presented assuming that the velocity is

uniform and there is no chemical reaction between the components. The analytical and numerical results

Page 18:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

were shown to be in a good agreement. A complete parametric study revealed that the well-known

butterfly-shaped profiles of the product concentration are not immediately established. In fact, spindle-like

profiles are created at the beginning and are evolved into butterfly-shaped profiles only near the steady-

state. The inclination of the concentration peak toward the component with either less diffusivity or less

inlet concentration, which has already been reported for a steady-state, was shown to be less significant at

the earlier stages of the production. Moreover, the inspection of the total efficiency, defined as the ratio of

the whole production to the amount of the limiting reactant within the device, demonstrated that the long-

term influence of the factors affecting the advection of mass is quite the opposite of their impact on the

transient response. More precisely, whereas increasing (a) the dimensionless Debye-Hückel parameter, (b)

the velocity scale ratio, and (c) the Péclet number give rise to higher efficiencies at short time scales; quite

the opposite is true at the steady-state. That is, increasing (a), (b) and (c) at steady state decreases the

efficiencies. The influence of other parameters on the efficiency is not time-dependent. For example,

increasing either the inlet concentration ratio or the Damköhler number is accompanied by better

efficiencies for non-steady and steady states. In contrast, increasing the diffusivity ratio always decreases

the efficiency.

Appendix

The functions Qi=1 ,… ,8appeared in Eqs. (18) to (22) are as follows

Q1 ( y )= eΩ y+e−Ω y+22 τ Ω

, Q2 ( y )= e2 Ω y+2eΩ y−2e−Ω y−e−2 Ω y

4 τ2 Ω(A1)

Q3 ( z )= eΩ z+e−Ω z+22 ατ Ω

,Q 4 ( z )= e2 Ω z+2eΩ z−2 e−Ω z−e−2Ωz

4 α2 τ2 Ω

Q5 ( x )= eΩ' ( x−1)+e−Ω' ( x−1 )+22l τ x

' Ω' , Q6 ( x )= e2Ω' (x−1 )+2 eΩ' (x−1 )−2 e−Ω' ( x−1)−e−2 Ω' ( x−1)

4 l2 τ x' 2 Ω'

Page 19:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

Q7 ( z )=sinh(0.5 τ z

' )

τ z' α cosh [ τ z

' ( z−0.5 ) ], Q8 ( z )=

−sinh2(0.5 τ z' )sinh [τ z

' ( z−0.5)]τ z

' α2 cosh3[τ z' ( z−0.5)]

where Ω=ln ( τ+1τ−1 ), Ω'=ln ( τ x

' +1τ x

' −1 ), τ=1.01, τ x' =1.05, andτ z

' =10. The coefficients of algebraic

equations (23) to (25) are also given as

aP , A=aW , A +aE, A +aS , A+aN , A +aB, A+aT , A+PeDa cB, P¿ g Δ3+Δ3/δ

(A2)

aP , B=aW , B+aE , B+aS , B+aN ,B+aB ,B+aT , B+βPeDa cA , P¿ g Δ3+Δ3 /δ

aP , C=aW ,C+aE , C+aS ,C+aN , C+aB,C+aT ,C+ Δ3/δ

aW , A=Δ2 [ Pe u¿Q5 ( x )−Q6 ( x ) ] (1+G1 ) , aW , B=Δ2 [ βPe u¿Q5 ( x )−Q6 ( x ) ] (1+G2 )

aW ,C=Δ2 [γPeu¿Q5 ( x )−Q6 ( x ) ] ( 1+G3 ) ,aE , A=Δ2 [ Pe u¿Q5 ( x )−Q6 ( x ) ] G1

aE , B=Δ2 [ βPe u¿Q5 ( x )−Q6 ( x ) ] G2 , aE, C=Δ2 [ γPeu¿Q5 ( x )−Q6 ( x ) ]G3

aS , A=aS , B=aS ,C=−Δ2 Q8 ( z ) (G4+1 ) , aN , A=aN , B=aN ,C=−Δ2 Q8 ( z ) G4

aB , A=aB , B=aB ,C=Δ, aT , A=aT ,B=aT ,C=Δ

SA=c A , P¿h Δ3/δ , SB=cB, P

¿h Δ3/δ ,SC=γPeDac A ,P¿ c B , P

¿ Δ3+cC , P¿h Δ3/δ

in which Δ is the spatial interval. A 100 ×100 ×100 mesh structure is considered in this study as it has

been found to provide sufficiently mesh independent results. Moreover, δ is the time interval which is

multiplied by 1.05 in each time step and its value for the first time step is δ 0=5 ×10−4; this value is

specified because further time refinement does not affect the results. In addition, the superscripts gandh

mark the results corresponding to previous iteration and previous time step, respectively. Finally, the

parameters Gi=1 ,⋯ ,4 are defines as

Page 20:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

G1= {e[ Pe u¿Q5 ( x )−Q6 ( x )]/Q52 ( x ) ∆−1}−1

,G 2={e [ βPe u¿ Q5 ( x )−Q6 ( x )] /Q52( x ) ∆−1}−1

(A3)

G3= {e [γPeu¿ Q5 ( x )−Q6 ( x ) ]/Q52 ( x ) ∆−1}−1

,G4=[e−Q8 ( z )/Q72 ( z )∆−1 ]−1

Acknowledgment

The authors sincerely thank the Iranian National Science Foundation (INSF) for their financial support

during the course of this work.

References

[1] J. B. Angell, P. Barth, and S. C. Terry, "Silicon micromechanical devices," Scientific American, vol. 248, pp. 44-55, 1983.

[2] L. Capretto, W. Cheng, M. Hill, and X. Zhang, "Micromixing within microfluidic devices," in Microfluidics, ed: Springer, 2011, pp. 27-68.

[3] Z. Wu and N.-T. Nguyen, "Convective–diffusive transport in parallel lamination micromixers," Microfluidics and Nanofluidics, vol. 1, pp. 208-217, 2005.

[4] M. S. Munson and P. Yager, "Simple quantitative optical method for monitoring the extent of mixing applied to a novel microfluidic mixer," Analytica Chimica Acta, vol. 507, pp. 63-71, 2004.

[5] H. Salimi-Moosavi, T. Tang, and D. J. Harrison, "Electroosmotic pumping of organic solvents and reagents in microfabricated reactor chips," Journal of the American Chemical Society, vol. 119, pp. 8716-8717, 1997.

[6] A. E. Kamholz, B. H. Weigl, B. A. Finlayson, and P. Yager, "Quantitative analysis of molecular interaction in a microfluidic channel: the T-sensor," Analytical chemistry, vol. 71, pp. 5340-5347, 1999.

[7] C. Luo, Q. Fu, H. Li, L. Xu, M. Sun, Q. Ouyang, et al., "PDMS microfludic device for optical detection of protein immunoassay using gold nanoparticles," Lab on a Chip, vol. 5, pp. 726-729, 2005.

[8] J. Wu, F. Yan, J. Tang, C. Zhai, and H. Ju, "A disposable multianalyte electrochemical immunosensor array for automated simultaneous determination of tumor markers," Clinical chemistry, vol. 53, pp. 1495-1502, 2007.

[9] W.-T. Liu, L. Zhu, Q.-W. Qin, Q. Zhang, H. Feng, and S. Ang, "Microfluidic device as a new platform for immunofluorescent detection of viruses," Lab on a Chip, vol. 5, pp. 1327-1330, 2005.

[10] D. Ferraro, J. Champ, B. Teste, M. Serra, L. Malaquin, J.-L. Viovy, et al., "Microfluidic platform combining droplets and magnetic tweezers: application to HER2 expression in cancer diagnosis," Scientific reports, vol. 6, p. 25540, 2016.

[11] J. Wegner, S. Ceylan, and A. Kirschning, "Flow chemistry–a key enabling technology for (multistep) organic synthesis," Advanced Synthesis & Catalysis, vol. 354, pp. 17-57, 2012.

[12] C. Wiles and P. Watts, "Micro reaction technology in organic synthesis," ed: CRC Press, 2016, pp. 363-382.

[13] P. Eribol, A. Uguz, and K. Ulgen, "Screening applications in drug discovery based on microfluidic technology," Biomicrofluidics, vol. 10, p. 011502, 2016.

Page 21:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

[14] R. Ran, Q. Sun, T. Baby, D. Wibowo, A. P. Middelberg, and C.-X. Zhao, "Multiphase microfluidic synthesis of micro-and nanostructures for pharmaceutical applications," Chemical Engineering Science, 2017.

[15] M. NaderiNasrabadi, Y. Mortazavi, and A. A. Khodadadi, "Highly sensitive and selective Gd 2 O 3-doped SnO 2 ethanol sensors synthesized by a high temperature and pressure solvothermal method in a microreactor," Sensors and Actuators B: Chemical, vol. 230, pp. 130-139, 2016.

[16] V. Sebastian, S. Basak, and K. F. Jensen, "Continuous synthesis of palladium nanorods in oxidative segmented flow," AIChE Journal, vol. 62, pp. 373-380, 2016.

[17] A. Dobhal, A. Kulkarni, P. Dandekar, and R. Jain, "A microreactor-based continuous process for controlled synthesis of poly-methyl-methacrylate-methacrylic acid (PMMA) nanoparticles," Journal of Materials Chemistry B, vol. 5, pp. 3404-3417, 2017.

[18] M. V. Bandulasena, G. T. Vladisavljević, O. G. Odunmbaku, and B. Benyahia, "Continuous synthesis of PVP stabilized biocompatible gold nanoparticles with a controlled size using a 3D glass capillary microfluidic device," Chemical Engineering Science, 2017.

[19] G. Karniadakis, A. Beskok, and N. Aluru, "Microflows and nanoflows: fundamentals and simulation. 2005," Cited on, p. 123.

[20] C.-C. Chang and R.-J. Yang, "Electrokinetic mixing in microfluidic systems," Microfluidics and Nanofluidics, vol. 3, pp. 501-525, 2007.

[21] M. U. Kopp, H. J. Crabtree, and A. Manz, "Developments in technology and applications of microsystems," Current Opinion in Chemical Biology, vol. 1, pp. 410-419, 1997.

[22] A. Manz, "Miniaturized chemical analysis systems based on electroosmotic flow," in Micro Electro Mechanical Systems, 1997. MEMS'97, Proceedings, IEEE., Tenth Annual International Workshop on, 1997, pp. 14-18.

[23] Z. Wu and D. Li, "Mixing and flow regulating by induced-charge electrokinetic flow in a microchannel with a pair of conducting triangle hurdles," Microfluidics and nanofluidics, vol. 5, pp. 65-76, 2008.

[24] H. Sharma, N. Vasu, and S. De, "Mass transfer during catalytic reaction in electroosmotically driven flow in a channel microreactor," Heat and mass transfer, vol. 47, pp. 541-550, 2011.

[25] C.-C. Cho and C.-L. Chen, "Mixing enhancement of electrokinetically-driven non-Newtonian fluids in microchannel with patterned blocks," Chemical engineering journal, vol. 191, pp. 132-140, 2012.

[26] A. A. Yazdi, A. Sadeghi, and M. H. Saidi, "Electrokinetic mixing at high zeta potentials: Ionic size effects on cross stream diffusion," Journal of colloid and interface science, vol. 442, pp. 8-14, 2015.

[27] B. Srinivas, "Electroosmotic flow of a power law fluid in an elliptic microchannel," Colloids and Surfaces A: Physicochemical and Engineering Aspects, vol. 492, pp. 144-151, 2016.

[28] L. Martínez, O. Bautista, J. Escandón, and F. Méndez, "Electroosmotic flow of a Phan-Thien–Tanner fluid in a wavy-wall microchannel," Colloids and Surfaces A: Physicochemical and Engineering Aspects, vol. 498, pp. 7-19, 2016.

[29] H. Song, Y. Wang, and K. Pant, "Scaling law for cross-stream diffusion in microchannels under combined electroosmotic and pressure driven flow," Microfluidics and nanofluidics, vol. 14, pp. 371-382, 2013.

[30] A. A. Yazdi, A. Sadeghi, and M. H. Saidi, "Rheology effects on cross-stream diffusion in a Y-shaped micromixer," Colloids and Surfaces A: Physicochemical and Engineering Aspects, vol. 456, pp. 296-306, 2014.

[31] A. Sadeghi, "Depletion of cross‐stream diffusion in the presence of viscoelasticity," AIChE Journal, vol. 61, pp. 4533-4541, 2015.

[32] H. Helisaz, M. H. Saidi, and A. Sadeghi, "3D modeling of reaction-diffusion dynamics in an electrokinetic Y-shaped microreactor," Sensors and Actuators B: Chemical, vol. 235, pp. 343-355, 2016.

Page 22:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

[33] W. Zhao, F. Yang, K. Wang, J. Bai, and G. Wang, "Rapid mixing by turbulent-like electrokinetic microflow," Chemical Engineering Science, vol. 165, pp. 113-121, 2017.

[34] H. Helisaz, M. H. Saidi, and A. Sadeghi, "Reduction of production rate in Y-shaped microreactors in the presence of viscoelasticity," Analytica Chimica Acta, 2017.

[35] N. G. Wilson and T. McCreedy, "On-chip catalysis using a lithographically fabricated glass microreactor—the dehydration of alcohols using sulfated zirconia," Chemical Communications, pp. 733-734, 2000.

[36] P. Watts, S. J. Haswell, and E. Pombo-Villar, "Electrochemical effects related to synthesis in micro reactors operating under electrokinetic flow," Chemical Engineering Journal, vol. 101, pp. 237-240, 2004.

[37] Z. Yousefian and M. H. Saidi, "Mass transport analysis of non-Newtonian fluids under combined electroosmotically and pressure driven flow in rectangular microreactors," Colloids and Surfaces A: Physicochemical and Engineering Aspects, vol. 508, pp. 345-359, 2016.

[38] Z. Lu, J. McMahon, H. Mohamed, D. Barnard, T. R. Shaikh, C. A. Mannella, et al., "Passive microfluidic device for submillisecond mixing," Sensors and Actuators B: Chemical, vol. 144, pp. 301-309, 2010.

[39] J. M. Reckamp, A. Bindels, S. Duffield, Y. C. Liu, E. Bradford, E. Ricci, et al., "Mixing Performance Evaluation for Commercially Available Micromixers Using Villermaux–Dushman Reaction Scheme with the Interaction by Exchange with the Mean Model," Organic Process Research & Development, 2017.

[40] A. Sadeghi, Y. Kazemi, and M. H. Saidi, "Joule heating effects in electrokinetically driven flow through rectangular microchannels: An analytical approach," Nanoscale and Microscale Thermophysical Engineering, vol. 17, pp. 173-193, 2013.

[41] C. Guldberg and P. Waage, "Experiments concerning chemical affinity," German translation by Abegg in Ostwald’s Klassiker der Exacten Wissenshaften, pp. 10-125, 1899.

[42] A. Sadeghi, "Analytical solutions for species transport in a T‐sensor at low Péclet numbers," AIChE Journal, vol. 62, pp. 4119-4130, 2016.

[43] J. M. Chen, T.-L. Horng, and W. Y. Tan, "Analysis and measurements of mixing in pressure-driven microchannel flow," Microfluidics and Nanofluidics, vol. 2, pp. 455-469, 2006.

Page 23:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

Fig. 1. 2D views of the microreactor under consideration including the coordinate system, dimensions,

and arrangement of components.

Page 24:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

t*

I mix

0 0.2 0.4 0.6 0.8 10

0.1

0.2

0.3 NumericalAnalytical

(b)Pe = 5

x*= 2

x*= 1

t*

I mix

0 0.2 0.4 0.6 0.8 10

0.05

0.1

0.15 NumericalAnalytical

(a)Pe = 1

x*= 1

x*= 2

Fig. 2. Comparison among the analytical results of I mix over time and the predictions of the numerical

code considering a uniform velocity field at two different axial positions for a) Pe=1 and b) Pe=5.

Page 25:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

x*

z*

0 5 10 15 20

(a)

x*

z*

0 5 10 15 20

(b)

x*

z*

0 5 10 15 20

(d)

x*

z*

0 5 10 15 20

(c)

Fig. 3. Axial movement of the concentration front over the time for components A (red counters) and B

(blue contours); the corresponding times for each part include a) t ¿=0.01, b)t ¿=0.04, c)t ¿=0.07, and d)

t ¿=0.1. The contours are depicted for the mid-plane ( y¿=0) while the dashed lines delineate the

concentration zones at the upper wall ( y¿=1). The other parameters include Κ=10, Γ=0, Pe=100,

β=2, ω=1, γ=1, and Da=1.

Page 26:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2

(a)

(b)

(c)

Fig. 4. Product contours at the mid-plane and at x¿=7 for a)t ¿=0.1, b)t ¿=0.2, and c)t ¿=0.3. The

evolution of the concentration profile from spindle-like form into butterfly-shaped form is observed in the

cross-sectional contours; the asymmetry of the production zone is also evident in these figures. The

parameter set considered include Κ=10, Γ=0, Pe=100, β=2, ω=2, γ=1, and Da=1.

Page 27:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

z*

c*

-2 -1 0 1 20

0.05

0.1

0.15

0.2

x* = 5x* = 10x* = 15

(a)

z*

c*

-2 -1 0 1 20

0.05

0.1

0.15

0.2

x* = 5x* = 10x* = 15x* = 20

(b)

z*

c*

-2 -1 0 1 20

0.05

0.1

0.15

0.2

x* = 5x* = 10x* = 15x* = 20

(c)

Fig. 5. Profiles of c¿ vs. z¿ at the mid-plane and at different axial positions for a) t ¿=0.1, b)t ¿=0.2, and

c)t ¿=0.3. The profile of x¿=20 is not shown in the first figure because the production has not already

Page 28:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

started therein. The parameter set considered include Κ=10, Γ=0, Pe=100, β=2, ω=2, γ=1, and

Da=1.

z*

c*

-2 -1 0 1 20

0.05

0.1

0.15

0.2

SSt* = 0.25t* = 0.2t* = 0.15

(b)

z*

c*

-2 -1 0 1 20

0.05

0.1

0.15

0.2

SSt* = 0.45t* = 0.35t* = 0.25

(a)

z*

c*

-2 -1 0 1 20

0.05

0.1

0.15

0.2

SSt* = 0.90t* = 0.75t* = 0.60

(c)

z*

c*

-2 -1 0 1 20

0.05

0.1

0.15

SSt* = 0.15t* = 0.125t* = 0.1

(d)

Fig. 6. Profiles of c¿ vs. z¿ at different times while keeping x¿=20 and y¿=0 for a)Κ=2, b)Κ=50, c)

Γ=−1, and d)Γ=1. Besides the transient profiles, the results of the steady-state, denoted by SS, are also

shown in the figures. The default parameters include Κ=10, Γ=0, Pe=100, β=2, ω=2, γ=1, and

Da=1.

Page 29:  · Web viewTwo Newtonian flows with similar flow rates and physicochemical properties are introduced by these arms under the combined action of electroosmotic and pressure forces

t*

t*

(%

)

(%

)

0 0.2 0.4 0.6 0.8 10

10

20

30

40

= -1 = 0 = 1

(a)

t*

(%

)

0 0.2 0.4 0.6 0.8 10

10

20

30

K = 2K = 10K = 50

(b)

t*

(%

)

0 0.1 0.2 0.3 0.40

5

10

15

20

= 1 = 2 = 4

(c)

t*

t*

(%

)

(%

)

0 0.2 0.4 0.6 0.8 10

5

10

15

20

25

30

Pe = 50Pe = 100Pe = 500

(e)

t*

(%

)

0 0.1 0.2 0.3 0.40

5

10

15

20

25

Da = 4Da = 2Da = 1

(f)t*

(%

)

0 0.1 0.2 0.3 0.40

5

10

15

20

25

= 4 = 2 = 1

(d)

Fig. 7. Efficiency enhancement over the time at different values of a)Γ , b)Κ , c)β, d)ω, e)Pe, and d)Da.

The default parameters include Κ=10, Γ=0, Pe=100, β=2, ω=2, γ=1, and Da=1.