Microscale - Capson -Commen- Modified Presure Equation - Final11

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 68

Development and validation of an exact analytical second-order slip

velocity model for flow in three-dimensional microchannel Array

Hariom Saran Singha*, Amit Kakkarb and P.M.V. Subbaraoc

abc
Department of Mechanical Engineering, Indian Institute of Technology Delhi, New Delhi,
Delhi, India 110016

* e-mail: om.saran7@gmail.com

1
Development and validation of an exact analytical second-order slip
velocity model for flow in three-dimensional microchannel Array

Experimental measurements and three-dimensional simulations are performed for


surface structure silicon microchannel with length 5400μm, width (average)
487.98µm and a height of 1.02µm. The nitrogen gas is taken as a working fluid, and
the mass flow rate of nitrogen gas is measured for the pressure ratio ranging from
1.24-1.96. The simulations data with general second-order slip velocity with
appropriate slip coefficients (C 1=1.18) and (C2=0.34) have a good agreement with the
experimental results.

Keywords: General second-order slip boundary condition, Slip coefficient, Knudsen


number.

1. Introduction

The study of fluid flow through microchannels has gained attention nowadays due to the

application of microdevices in the cooling of electronic circuits, computer chips, fuel cells,

and medical devices [1]. Microscale devices have characteristic dimensions in micron,

which is comparable to the mean free path of the gas molecule even at atmospheric

conditions. The gas flow through these devices is called rarefied flow. The rarefaction is

quantified as the ratio of the mean free path (λ) of gas molecules to the characteristic

dimension of the system (l), which is called the Knudsen number (Kn) [2]. The range of

Knudsen numbers between 0.001 to 0.1 causes slight rarefaction due to which slip occurs at

the wall. It is called a slip regime. Flow problems of this regime can be modelled using

continuum fluid mechanics with slip boundary conditions. Gad-Al-Huk [3] had done flow

regime classification using the Knudsen number and studied boundary conditions in the slip

regime. In rarefied gas flow, fractional reduction of tangential momentum of gas molecules

occurs due to momentum exchange with the wall. This reduction in the momentum of gas

molecules is called the tangential momentum accommodation coefficient (TMAC).

2
Some researchers [4]–[7] used Boltzmann and Burnett equations to model the rarefied gas

flow . Many authors [8]–[11] obtained experimental data and solved the problem of slip

regime analytically using the Navier-Stokes equation with slip boundary condition. Arkilic

et al. [9] used the perturbation method and solved the Navier-Stokes equation with the first-

order slip boundary condition. The authors also validated it against the obtained

experimental results. Some other notable works of researchers [12]–[15] have compared

their experimental results with the analytical solution of Arkilic et al. [9]. Zahid et al. [16]

studied compressibility, rarefaction in gas flow through the microchannel.

Besides the plane model, Ebert et al.[17] and Morini et al.[18], [19] suggested a new

rectangular model using first-order slip boundary conditions.

Apart from molecular dynamics, Lattice Boltzmann approach, computational fluid dynamics

is also used to simulate the gas flow through microchannel using the Navier-Stokes

equation. The Navier–Stokes equation with first-order slip and temperature jump boundary

condition was solved by Chen et al.[20] using a finite-difference methodology. The results

of pressure distribution were validated with Pong et al. [21] and the mass flow rate with

Arkilic et al. [9]. Roy et al. [22], [23] developed 2D code using the Navier -Stokes equation

with first-order slip velocity and temperature jump boundary condition. The authors solved

it using the Finite Element Method. The simulation was carried out by Morini et al. [18]

using the Navier-Stokes equation as a governing equation with a first-order slip model. They

analysed the effect of rarefaction on the flow. Cai et al. [24] further extended the work done

by Arkilic et al.[9] with the inclusion of the energy equation and the removal of the

isothermal flow assumption. Authors obtained analytical solution using compressible

Navier-Stokes equation with first-order slip boundary and temperature jump boundary

condition. They have validated results with the numerical solution.

3
Besides 2D Analysis, some limited studies of the gas flow in the 3D channels are also

available in the literature. Researchers [25]–[28] developed a numerical procedure and

solved 3D rarefied gas flow using first-order slip boundary condition and studied the

rarefied gas flow through microchannel.

To check the Navier-Stokes equations validity in the transition regime requires a higher-

order boundary condition. Towards this, initially, Schamberg et al. [29] obtained second-

order boundary conditions using Burnett approximate solution of the Boltzmann equation.

Although it was observed that the solution did not always have an agreement with

experimental data of Schaaf et al. [30] and in many cases, even Navier -Stokes equation

worked better than that. Deissler et al. [31] and Cercignani et al. [32] derived second-order

boundary conditions for slip flow at a higher Knudsen number. Dongari et al. [33] used

second-order boundary condition and obtained analytical solution for higher Knudsen

number flow. Few researchers performed experiments and validated their experimental data

using second order slip boundary condition. Sreekanth et al. [34] proposed a modified

boundary condition and given the general form of boundary condition containing

independent slip coefficient C1 and C2, based on experimental data obtained the flow in

circular tube. In addition to this, the authors also carried out the experimental measurement

of gas flow through the microchannel. Colin et al. [35], [36] used Deissler’s second-order

boundary condition and obtained mass flow rate expression. The measured data was

validated with analytical solution up to Knudsen number 0.22. The plane model was also

validated with first-order boundary conditions only up to Kn = 0.05. Authors inferred

variation between the plane model and rectangular model results for channel of aspect ratio

(ar) of 0.01 and Kn = 0.053. The increase in the aspect ratio (ar) causes more deviation of

models.

4
Maurer et al. [37] have obtained experimental result up to mean Knudsen number 0.8 for

helium and upto 0.6 for nitrogen respectively. They obtained nondimensional mass flow rate

as function of mean Knudsen number as controlling parameter. The variation of slip

coefficients was obtained fitting the experimental data with polynomial of second order.

Ewart et al. [38] obtained experimental data with helium for Knudsen number from 0.03 to

50, and results compared with solution obtained from kinetic approach. Moreover, slip

coefficient was extracted by comparing the non-dimensional mass flow rate using first order

slip boundary for Knudsen number 0.03 to 0.3 and for 0.03 to 0.7 using polynomial of

second degree. In the transition and free molecular regime kinetic approach is used for slip

coefficient calculation. Graur et al. [39] measured the highly accurate data for different

gases for flow through microchannel to obtain the slip coefficient and study the flow

behaviour for gas flow in slip to free molecular regime (0.03-30). They have obtained

nondimensional mass flow rate expression for first and second order slip boundary condition

using continuum approach for Knudsen number 0-0.03. The slip coefficients for various

gases were obtained. The experimental data for slip to free molecular regime has been

compared with kinetic approach and analytical solution has been fitted using value of

variable slip coefficient as fitting parameter. They have concluded that irrespective of

Knudsen number and modelling approach TMAC decreases with increase in molecular

weight of gases. The values of TMAC are below unity this indicates that using full

accommodation coefficient is not right approach to deals with gas surface interaction.

Beskok et al. [40] proposed universal model to study the validity for wide range of Knudsen

number (0<Kn<∞) and compared it with DSMC and experimental data. They concluded that

model has good agreement for Kn ≤ 1, however the deviation between experimental and

simulation results was found 21% for Kn>1. Model used are based on diffused reflection

which is one of the limitations for accepting it as general slip model. The different slip value

5
are listed by Barber and Emersion [41] for model having two slip coefficients.

Vadiraj et al.[42] carried out an experimental measurement of the gas flow through a

trapezoidal microchannel having dimensions with a depth of 103µm, top and bottom width

as 1143µm, 998µm with the length of 2cm and bonded using quartz (dissimilar material).

They have performed experiment and obtained slip coefficient variation using the

methodology of Maurer et al.[37]

Srinivasan et al. [43]–[45] fabricated microchannels with different aspect ratios (0.002, 0.01

and 0.1), channels were etched on the silicon wafer and bonded with another wafer using

fusion bonding. This fabrication method led to a controlled surface structure microchannel.

They have performed experiments and simulation modelling using second order boundary

condition and studied the flow behaviour. Reference [43] new analytical model was

developed using second-order slip boundary conditions to check the validity of Navier–

Stokes equations from slip to transition regime Authors model the experimental data up to

outlet Knudsen number 1.01 for nitrogen gas flow also determined the slip coefficient. In

Reference [44] further they modelled up to free molecular regime and concluded that up to

Kn ≤1 measured and experimental mass flow rate deviation are within 2% using slip

coefficients C1=1.31 and C2=0.11, and for Knudsen number ≥1 normalised volume flow rate

had a 3.30% deviation with other available experimental data. The Knudsen minimum was

also seen in the results.

Moreover, mass flow rate, pressure distribution and velocity distribution were plotted to

study fluid flow behaviour. Reference [45] they have studied flow behaver through smooth

and rough channel. Due to low roughness slip is high and TMAC value predicted as

0.851for smooth channel.

6
Few researchers [36], [37], [42] obtained experimental data where channel was bonded with

dissimilar material and researchers [38], [39] with channel bonded with same material

(silicon to silicon) used the first and second-order boundary condition checked the

applicability and

using second order boundary condition was derived without any approximation to check the

validity of previous proximate model and further used to model the three-dimensional

rarefied gas flow through microchannel. To replicate the actual experimental condition,

microchannel having a trapezoidal cross-section with inlet and outlet plenums (as a

reservoir) is used for 3D simulation. The simulation results are compared with experimental

data obtained from the controlled surface structure microchannel.

2. Development of Analytical Model for Slip Velocity


Whenever a gas flows past a solid surface, there is always the formation of the Knudsen

layer or kinetic layer and, its thickness is the order of one mean free path. When the mean

free path is comparable to the characteristic dimension of flow, the flow is called rarefied

flow. In rarefied gas flow, the collision frequency in Knudsen layer regions is insufficient to

bring quasi thermodynamic equilibrium. The partial exchange of momentum and energy in

this layer causes slip flow on the wall.

A first-order slip boundary condition is initially used to model this problem, which has

accurately modelled the flow up to Kn = 0.10. It is evident from Figure 1 that there is a

deviation between actual slip velocity and extrapolated first-order slip velocity.

Furthermore, the increase in Knudsen number leads to a higher mean free path which causes

an increase in velocity deviation.

For higher Knudsen number, many researchers used Navier-Stokes equation with second-

order slip boundary condition and modelled the rarefied flow taking approximation in

7
expression as included in the introduction part. Although the second-order solution was the

approximate solution, its modelling ability is better than the first order. This indicates that

the magnitude of second-order slip velocity lies between actual and extrapolated slip

velocity corresponding to first-order slip boundary conditions like point Q in Figure 1.

For the analytical modelling, we can refer to Srinivasan et al. [43]. Nondimensionalised

governing equations and second-order slip boundary conditions are used here directly.

The mathematical analysis steps are included to obtain slip velocity at the wall using the

general second-order slip boundary condition without approximating pressure expression.

First-order boundary condition

(1)

General second-order slip boundary condition

(2)

y
Velocity profile

z x

Knudsen Layer
(O) λ

Q
Wall Actual
First order

Figure 1. Knudsen layer near the wall

8
The streamwise (x-direction) velocity profile is obtained after nondimensionalization, order

of magnitude comparison and integration of x-direction momentum equation; we get

(3)

Now, the boundary conditions that were applied in equation (3) are given below.

(i) At

(ii) At
Implementing the boundary conditions, the value of constants is evaluated and given as:

inserting the value of A and B in equation (3), we get

(4)
In the above equation, Kn is variable and function of x coordinate. Knudsen number is based

on channel height, and it can correlate the local Knudsen number to the outlet Knudsen

number using the relation given below.

Where is the Knudsen number based on the outlet condition of the microchannel.

Substituting the value of streamwise velocity in nondimensionalised continuity equation and

integrating with respect to ~y equation for ~v can be obtained.

(5)

9
The value of constant is evaluated using the boundary condition for v along the axis. From

the literature survey, Roy et al. [22] found that the velocity along the y-direction will vanish

at the centre and asymmetric along the x-axis. Applying the boundary conditions.

at , hence C = 0

Now putting the value of C in equation (5), we get

(6)

Putting the boundary condition in equation (6) that the wall-normal velocity must be zero

(no penetration condition) at the wall and taking bottom wall condition.

So, at ,

(7)

Integrating equation (7) with respect to x, we get

(8)

Where D and E are constants.

The presence of a logarithmic term in the pressure equation makes the equation highly

nonlinear. In literature, reference [43] has done an order of magnitude analysis and

compared different terms in pressure equation (10) for slip regime using slip regime

experimental data. In literature reference [34], [37] it was found that C1 is of order one and

C2 is one order less.

Reference [43] used nondimensional pressure variation data of Roy et al. [22] slip regime

data and performed the order of magnitude comparison using values C 1=1.13 and C2=0.13.

Based on the magnitude comparison of each term in the pressure equation, they replaced

C2ln~
p with a constant because the magnitude of the logarithmic term is negligible compared

to other the two terms.

10
The present work solved a highly nonlinear pressure equation (9) using the Newton Raphson

scheme to get pressure profile variation along the flow direction. It has seen from the

equation (9) for low Knudsen number influence of the third term 24Kn o2C2ln~
p is negligible.

However, as the Knudsen number increases, the logarithmic term will affect the pressure

equation and slip velocity, suggesting that it should be retained in expression at a higher

value of the Knudsen number.

The pressure boundary conditions are used, and values of constants D and E are calculated.

Inlet and outlet pressure boundary conditions have been given as:

(i) At
(ii) At

Where P denotes the ratio of pressure is represented by,

Putting the above boundary condition in equation (8), we get

Now we know that.

(9)

(10)

placing the value of D in equation (10)

(11)

Plot of nondimensional pressure variation with approximation and exact solution are shown

in Figure 2 for slip and transition regimes data. It was found that in slip regime both profiles

11
are quite matching shown in Figure 2 (i) for Kno=0.067, pressure ratio 1.84. However, in

early transition regime both profile are different the plot shown in Figure 2 (ii) for

Kno=0.22, pressure ratio 1.95, both data was taken from reference [47]. Another data of end

of transition regime taken from reference [44] shown in Figure 2 (iii) for Kno=0.90 pressure

ratio 1.93 the deviation clearly indicates two different profiles and it indicates that profile

deviation is increases with increase in Knudsen number in transition regime.

(i)

(ii)

(iii)

Figure 2. Variation of approximate and exact nondimensional pressure distribution (i)


Kno=0.667 pressure ratio 1.84 (ii) Kno=0.22 pressure ratio 1.95 (iii) Kno=0.90 pressure ratio
1.93.
Therefore, in the transition regime analysis exact analytical model must be used for accurate

analysis. Further using exact analysis slip velocity profile has been developed to apply on

wall to implement slip boundary condition.

12
Substituting the value of pressure gradient in equation (4) from equation (11), we get

(12)

To evaluate the slip velocity on the top and bottom wall, the value of ~y=± 1/2 have

substituted in above the equation.

(13)

Here

Kno- Knudsen number at channel outlet based on channel height.

H- Depth of channel

-Nondimensional pressure

P- Pressure ratio

The slip velocity is calculated from the above equation (13) at the top and bottom wall,

which is a function of variable ~


p , and that depends on ~
x only.

H-Depth of the microchannel

-Nondimensional Pressure and

Slip velocity at the wall was evaluated using equation (13) along the flow direction. The slip

is denormalised and implemented using UDF in Fluent with a suitable slip coefficient and

converged inlet pressure has been compared with experimental data.

3. Fabrication Characterisation and Bonding

Fabrication, characterisation, and bonding have been carried out by reference [43] . Here

microchannel system used for the study of 3D flow is etched on the silicon wafer and

13
bonded to other silicon wafer using fusion bonding. The dimensions of the microchannel are

given below in Table 1.

Table 1. Dimensions of the microchannel system


Length(µm) Width(µm) Depth(µm) Aspect Number of
ratio(ar) channels

System1 5400 Wtop= 488.70 1.02 0.002 20


Wbot= 487.26
Uncertainty ±15 ±1 ±0.005
Note-Surface roughness of channel ( Rarms =15 nm± 3)

4. Computational fluid dynamics Analysis and Model Validation

The creation of the geometry and mesh generation was done in ICEM CFD 18.1. The fine

mesh was used near the wall to capture the sharp gradient and coarse mesh as it approaches

the fluid core. The finite volume method was used to solve the governing equations in each

cell. The second-order upwind scheme is used in discretising momentum and energy

equation. The pressure discretisation is also second-order accurate. The convergence criteria

for continuity was obtained by dividing the experimental mass flow rate with the number of

cells present in the cross-section of the computational domain. The validation in gas flow

through the 3D microchannel was obtained for first-order and general second-order slip

boundary conditions. The first-order slip boundary condition was validated with Arkilic et

al. [9] experimental work and general second-order boundary condition with Colin et al.

experimental data [36] and [48]. A two-dimensional simulation is also performed for

velocity profiles comparison.

4.1. Computational fluid dynamic Analysis


The computational fluid dynamics was used to solve the steady 3D isothermal compressible

flow with the slip velocity profile using the ideal gas model. The channel was having a

trapezoidal cross-section due to selective wet etching. The angle (θ) between the top wall

14
and sidewall of the channel is 54.7° shown in Figure 3. However, the present test section has

an array of twenty channels with a common inlet and outlet plenum. It was assumed that in

steady-state flow conditions, the flow was uniformly distributed to each channel. Therefore,

a single microchannel with a trapezoidal cross-section having an inlet and outlet plenum was

used for numerical simulation. The fluid domain of the microchannel with the inlet and

outlet plenum was discretised with structured hexahedral mesh. The 3D steady compressible

Navier-Stokes equation and the continuity equation were used as the governing equation.

The general second-order slip boundary condition was used on the top and bottom walls.

The viscosity of the working fluid is assumed to be constant. The equation of state was used

to define the pressure in modelling the steady isothermal flow.

Inlet Plenum Outlet Plenum


y

z x

θ H

Figure 3. Microchannel with inlet and outlet plenum

4.1.1. Governing equation and Boundary conditions:


The governing equations for 3D steady-state gas flow through microchannel simulation:

Continuity equation:

(14)

x-direction momentum equation:

(15

15
)

y-direction momentum equation:

(16

z-direction momentum equation:

(17
)

Equation of state:

(18)

This flow is compressible because of the substantial pressure drop along the flow direction

due to friction. So, to obtain the pressure in the flow field ideal gas model is used. In the

numerical simulation, the experimental mass flow rate through a single channel was taken as

the inlet boundary condition and pressure as the outlet boundary condition. Therefore,

general second-order slip velocity boundary condition was applied only on the top and

bottom walls. In the CFD simulation, general second-order slip boundary conditions cannot

be applied directly. So, to implement this boundary condition, the slip velocity profile was

solved analytically and implemented on the top and bottom wall using a user-defined

function.

The expression for slip boundary condition is given as:

To implement the slip velocity on sidewall outlet Knudsen number is calculated based on

channel width as characteristic dimension. The Knudsen number based on channel width

was 0.00021, which lies in the continuum regime. Therefore, no-slip boundary condition is

16
applied on the sidewalls of the channel. The experimental flow was assumed to be

isothermal because the material of silicon wafer has thermal conductivity ̴ 38 W/mK at room

temperature.

The temperature boundary condition for isothermal flow

(19)

The numerical simulation was carried out because the momentum equations introduce

substantial complexity in the solution of flow problems due to the presence of the nonlinear

advection term and the second-order derivative viscous term. There is no general analytical

solution for the system of the equations because it cannot be directly converted into an

ordinary differential equation for 3D flow even with no-slip boundary conditions.

Many fluid flow problems with complex geometry and boundary conditions have no

analytical solution and were solved using the numerical technique. Nevertheless, the

problem with the analytical solution provides a significant relationship among the solution

parameters if it can be obtained. The combination of the numerical and analytical solution

was used to model the problem because, by default, Fluent has a no-slip boundary condition

at the wall. The slip boundary profile obtained along the streamwise direction analytically is

implemented on the top and bottom surface of the channel using UDF. For the slip velocity

profile, prerequisites are pressure gradient and slip coefficients; the pressure gradient also

depends on slip coefficients.

4.1.2. Grid independency test


The variation of microchannel inlet pressure with mesh size is showed in Figure 4. The

optimal mesh selected for simulation has 0.025% of relative error of inlet pressure with

respect to the next level of refined mesh. A mesh having 5219900 nodes was selected in the

grid independence test. The convergence criteria of order 10-16 was taken for continuity.

17
Figure 4. Grid independency test for simulation of present experimental data.

4.1.3. Calculation of slip coefficients for second-order boundary condition


To provide the general slip boundary condition, the slip coefficients C 1 and C2 were

required. The appropriate value of slip coefficient C 1 was obtained using the velocity profile

of the first-order slip boundary condition on the top and bottom wall of the channel.

Furthermore, using the slip coefficient C1 and a guess value of C2, the second-order slip

velocity was obtained and again simulated for the appropriate value of C2.

Steps involves for calculating the slip coefficients for first, general second-order boundary

conditions:

Step I

18
Step II

Step III

Step IV

Step V

Step VI

In every step, expression of the first-order slip boundary condition is written in the first row,

and the second one is related to the general second-order slip boundary condition. In the first

step, streamwise velocity profiles obtained from the analytical solution is given, which is a

function of streamwise (~
x ), its normal coordinates ~y and slip coefficients. In the second

step, the velocity profile normal to streamwise distribution is obtained using a continuity

equation. The aim of calculating this profile is to get the pressure profile. Now obtained

pressure profile expressions were given in the third step. In the fourth step, pressure gradient

expressions are given. The streamwise velocity is given in step fifth and slip velocity at wall

expression is given in step sixth.

19
Estimation of first-order slip coefficients C1:

Using the guess value of slip coefficient C1 in the first-order pressure expression, the

pressure gradient is calculated using the fourth step. The calculated pressure gradient is used

in first-order velocity expression in the first step. Now, step fifth has been achieved after

substituting the pressure gradient in the first step. This first-order velocity expression is only

a function of nondimensional coordinates~


x , ~y . To get the expression of slip velocity at the

top/ bottom wall as a function of ~


x , ~y=± 1/2 substituted in step fifth. The expression of

velocity as a function of ~
x is obtained are shown in step sixth.

Now in step sixth, we have got the first order slip velocity expression as a function ~
x only.

For the different values of ~


x , slip velocity ~
u at the wall along the streamwise direction was

obtained. Now denormalisation was done to obtain the actual value of slip velocity, and

UDF is used to implement it on the wall of the channel. After, that 3D simulations are

performed (mass flow rate as inlet and pressure as outlet boundary conditions), and inlet

pressure was obtained and compared with experimental results. This process has been

repeated until we got appropriate values of first-order slip coefficient C 1. From Table 2, We

can see the simulated inlet pressure values and their percentage deviation (%δ) with respect

to experimental data using the first-order slip boundary condition. Here, C 1=1.18 is taken as

the best value for the first-order slip coefficient.

Table 2. Percentage deviation between measured and simulated inlet pressure for different
first-order slip coefficients for data of P=1.58.

20
Estimation of slip coefficients C1, C2:

The slip coefficient C1 (obtained from iterative method) and a guessed value of second slip

coefficient (C2) is substituted in pressure expression obtained from the second-order analysis

given in step three. Now pressure gradient has been obtained using step fourth. The obtained

pressure gradient is substituted in second-order velocity expression in step first and the

attained streamwise velocity expression shown in step fifth. Now in step fifth, velocity

expression obtained from the second-order analysis is a function of coordinates ~


x , ~y . To get

the expression of slip velocity at the top/bottom wall as a function of ~


x , ~y=± 1/2 substituted

in step fifth. The expression of velocity as a function of ~


x was obtained in the next step.

Now in step sixth, velocity profile expression is a function ~


x only, for different values of ~
x

velocity along the streamwise direction has been calculated. Now denormalising the velocity

and using UDF, this slip profile is implemented on the walls of the channel (top and

bottom). After implementing the second-order slip velocity profile, simulation was done

(mass flow rate as inlet and pressure as outlet boundary conditions), and inlet pressure was

obtained. This simulated pressure was compared with experimental results. This process was

repeated for different values of slip coefficient C 2. From Table 3, for the second-order

boundary condition lowest deviation is with the value of C 1 =1.18, C2 =0.15. However,

keeping the value of C1 =1.18, and varying value of C2 for all experimental data of the same

Knudsen number, the simulations have been performed and obtained deviations are shown

in Table 4. It can be observed that C 1=1.18, C2 =0.34 has minimum deviation. The

coefficients C1 =1.18, C2 =0.15 have not been selected because the reduction in C 2 might be

capturing uncertainty in experimental data. That is why it has more accurate for single

experimental data but for all six data average deviation more. So, finally, C 1=1.18 and

C2=0.34 is selected for final second-order slip coefficients.

21
Table 3. Percentage deviation between measured and simulated inlet pressure for a fixed

value of C1 and different values of C2 for data of P = 1.58.

Here formula for percentage deviation is given as:

[∑ ( ( ( ]
k=6
%δ=average P i simulated −Pi experimental ) /P experimental )∗100 )k
k =1

Table 4. Average percentage deviation between simulated and measured inlet pressure for a

fixed value of C1 and different value of C2 for all six data of Kno= 0.1028.

4.2. Model Validation

A new numerical model is required to validate with experimental and numerical data

available in the literature. In the present work, two sets of simulations were performed based

on data available in previous research works. The first set of simulations was performed to

validate the first order, and the other one was for validating the second-order general slip

boundary conditions. To provides the slip on walls of the channel, the slip velocity profile

was developed along the height and width of the channel separately. The 3D simulation for

a first-order slip model was validated against empirical results of Arkilic et al. [9] for Kno =

22
0.156. The second-order general slip boundary conditions for Kn o = 0.09 and 0.47 were

validated with experimental data (inlet pressure data) of Colin et al. [36], [48].

4.2.1. First-order model validation


The simulation is carried out for measured mass flow rate data employing first-order

boundary conditions for the pressure ratio 1.15 to 2.48, outlet pressure 100.8 kPa, and outlet

Knudsen number 0.156. The working fluid used was helium. The 3D simulations with no-

slip boundary conditions have been performed on different mesh sizes for the grid

independence test. The optimal mesh having 0.01% of relative error in the value of inlet

pressure to the next level of refined mesh has been selected for simulations. The variation of

pressure at the microchannel inlet with respect to mesh size is shown in Figure 5 (i) and

Figure 5 (ii). A mesh having 450000 nodes was selected in the grid independence test to

carry out further simulations.

(i) (ii)
Figure 5. Grid independency test for first-order validation (i) with reference [9] and (ii) with
reference [20].

The convergence criteria for continuity have been found of order 10 -15. So, to achieve this,

the mass flow rate is taken as inlet and pressure as outlet boundary conditions. For the

Knudsen number 0.156, the slip velocity profile was obtained for each pressure ratio flow.

The default no-slip boundary condition is replaced with a slip velocity profile using UDF to

23
implement slip boundary conditions on the channel walls. The simulation was performed for

the first-order model validation, and inlet pressure was plotted with respect to experimental

data of mass flow rate in Figure 6. It was found that the maximum deviation in the data of

Chen et al.[20] and Roy et al.[22] is 3.89% and 5.14%, respectively and, the maximum

deviation for the present simulation was 3.83% for inlet pressure.

Figure 6. Variation of mass flow rate with inlet pressure for slip condition at = 0.156
reference [9], and comparison with [20], [22].
Moreover, validation has also been done against numerical data of Roy et al. [23] for

velocity distribution along the centreline direction and transverse velocity of streamwise

direction. The 2D simulations were performed on different mesh sizes. The optimal mesh

selected for simulation was having 0.014% of the relative error of the mass flow rate at the

outlet with respect to the next level of refined mesh. A mesh having 41000 quadrilaterals

was selected in the grid independence test. The variation of mass flow rate at the inlet of

microchannel along with mesh size is shown in Figure 5(ii). The convergence criteria for

continuity was taken as 10-12 (for pressure boundary condition as the lowest criteria for

continuity convergence). The maximum variations in streamwise and transverse direction

velocity are within 1.30% and 1.68%, respectively. The plots are shown in Figure 7 (i) and

(ii).

24
(i)

(ii)

Figure 7. Comparison of (i) Centreline velocity and (ii) transfer velocity along the length of
the channel for pressure ratio 1.34 and 2.70 with reference [23].

4.2.2. Second-order model validation


The inlet pressure variation of the microchannel with mesh size is shown in Figure 8. The

meshes having 450000 nodes for Kn=0.09 shown in Figure 8 (i) and 330000 nodes for

kn=0.47 shown in Figure 8 (ii) have been selected for the grid-independence test. For both

cases the mesh chosen has a relative error below of 0.03% of inlet pressure with respect to

the next level of refined mesh. The convergence criteria of the order 10-16 were taken for

continuity.

25
(i) (ii)
Figure 8. Grid independency test for validation of experimental data of (i) reference [36]
and (ii) reference [48].

So, to achieve convergence criteria, mass flow rate as inlet and pressure as outlet boundary

condition was used. The simulations were performed by applying the second-order

boundary conditions, and a converged value of pressure at the inlet of the microchannel was

obtained. The working fluid used was nitrogen and helium.

Colin et al. [36] performed an experiment for gas flow through the microchannel and

developed an analytical model for the same. They found that for a flow-through channel

having an aspect ratio (ar) ≥0.01, the plane model underestimates the flow rate for Knudsen

number (Kn)≥0.05. They have also reported that the lateral wall affects the channel flow for

aspect ratio (ar) ≥0.01.

Furthermore, they found that their analytical model based on Deissler’s boundary conditions

have good agreement between experimental and analytical solutions for TMAC=0.93, up to

Knudsen number 0.22.

An exact analytical solution was used to obtain the slip velocity boundary profile in the

present work. The simulations were performed for flow through microchannel with slip

coefficients C1 =1.15 for the first order and general second-order boundary conditions with

slip coefficients C1 =1.15 and C2 =0.34. For Kn =0.09, the present simulation is in good

26
agreement with the experimental data, as shown in Figure 9 (i). The maximum deviation of

the first-order and Colin’s analytical solution are 0.50% and 2.36% w.r.t experimental value

and, the deviation of second-order simulation is 1.43%. The first-order, general second-

order and Colins model have good agreement with experimental data of slip regime. Figure

9 (ii) shows that Colin’s rectangular model overestimates the mass flow rate, due to which

inlet pressure is underestimated.

(i)

(ii)
Figure 9. Comparison of inlet pressure against mass flow rate for (i) channel having an
aspect ratio of 0.055 at Kno= 0.09 (ii) channel having an aspect ratio of 0.01 at Kno= 0.47.
For Knudsen number 0.47 using the same slip coefficient as C 1=1.15 and C2=0.34,

simulations were performed. The obtained inlet pressure has good agreement with

experimental data having deviation 0.59%. The Colin’s analytical solution obtained by the

rectangular model underestimated inlet pressure up to 7.47 %. This deviation is due to the

overestimation of the mass flow rate. Based on the above experimental and numerical

27
validation, it can be inferred that the present modelling technique can correctly predict the

experimental results up to Knudsen number 0.47.

5. Experimental Set-Up
The experimental set-up used in this work is developed in the turbomachinery Laboratory at

the Indian Institute of Technology Delhi. A cylinder filled with nitrogen gas up to 120 bar is

the source of the working fluid. The pressure regulator was connected to the gas cylinder to

provide the gas flow at constant pressure during experimental work. The Festo filters of

specifications (5, 1, 0.01µm) are used to filter the particle of bigger size and other impurities

present in the working fluid. Pressure transducer having an accuracy of ±0.1% of full-scale

is used for measuring inlet and outlet pressure of microchannel. The temperature

measurement was done by a resistance temperature detector (RTD) sensor with accuracy

±0.5% of full scale, and the temperature was maintained constant during the experiments (at

35.5℃). Mass flow meters of various ranges, like 0.5, 1, 2 Sccm was used. It showed

accuracy ±0.15% of full scale connected to the computer with data acquisition system. The

vacuum generator is used before the inlet of the microchannel to create low pressure. The

vacuum assembly is connected at the channel outlet to create a vacuum to achieve the

required Knudsen number. The whole experimental set-up is assembled on class 10 clean

bench.

6. Result and Discussion


Experimental data of gas flow in the microchannel having an aspect ratio (ar) 0.002 were

recorded for outlet Knudsen number of 0.1028. The mass flow rate variations were

measured and recorded for various pressure ratios (1.24-1.96). The results of inlet pressure

obtained from 3D simulations are compared with the experimental results.

28
6.1. The Methodology and Experimental Data
The detailed methodology is presented for the experiment of a gas flowing through the

microchannel. To achieve the required outlet Knudsen number, the vacuum is created inside

the vacuum chamber and pressure is provided at the inlet of the microchannel as per the

design pressure ratio. The system is first allowed to achieve a steady-state, which generally

takes about 30-45 min. The experimental data of pressure at the inlet, outlet and mass flow

rate are recorded using a data acquisition system.

For outlet Knudsen number 0.1028, outlet pressure is maintained at 6.51×10 4 Pa, and inlet

pressure varies from 8.07×104 Pa to 1.28×105 Pa. Experiments was performed for six

designed pressure ratios range from 1.24 to 1.96. Experimental data of nitrogen gas flow

through the microchannel is given in Table 5.

Table 5. Experimental condition for Nitrogen gas flow

S.No Parameter Range (Mean value)


1 Outlet Pressure 6.51×104 Pa
2 Outlet Knudsen number Kno= 0.1028
3 Temperature 305.5K
4 Pressure ratio 1.24 to 1.96

In this work, the UDF has been developed and implemented in Fluent to provide slip

velocity at channel walls (top and bottom). The slip velocity is obtained using the exact

analytical solution with the appropriate independent slip coefficient C1=1.18 and C2=0.34.

The simulations are carried out for six different pressure ratios. The experimental mass flow

rate was taken as inlet boundary condition and pressure as an outlet boundary condition. The

converged values of inlet pressures were obtained using slip and no-slip boundary

conditions at the wall to verify the slip effect. An average deviation of error between

experimental and simulation results was calculated. The results of simulated inlet pressure

with no-slip boundary condition have a deviation of 17.44% with respect to the measured

value of inlet pressure. The positive error indicates that higher inlet pressure is required due

to the no-slip condition at the wall to maintain the same flow rate. This research focused on

29
the applicability of the second-order model with an exact analytical solution over a higher

range of Knudsen numbers.

6.2. Description of Flow Through the Microchannel


The 3D simulation is carried out for flow through trapezoidal cross-section microchannel

having length 5400 µm, top width 488.7 µm, bottom width 487.26 µm and height 1.02 µm.

The simulation was performed by employing first-order and general second-order boundary

conditions with the appropriate slip coefficient (C 1=1.18 and C2=0.34). The uncertainty has

also been calculated for the accuracy of the experimental results, which is shown in plot of

Figure 10.

Figure 10. Variation of experimental and simulation inlet pressure with mass flow rate for
Kno= 0.1028.
The variation of simulated inlet pressure against experimental mass flow rate is plotted in

Figure 11. The first-order slip model results deviated from the experiment results by 2.77%,

and the general second-order boundary condition deviation was 0.51%.

30
Figure 11. Variation of experimental and simulation inlet pressure with mass flow rate for
Kno= 0.1028.
It indicates that the general second-order model can better predict the inlet pressure

conditions at a higher Knudsen number. This model also works better than Colin et al. [36]

at a higher Knudsen number.

For the study of microflow behaviour, pressure distribution along the streamwise direction is

presented in Figure 12. It shows its nonlinear nature along the flow direction and constant in

cross-section because no pressure gradient is applied in the transverse direction.

31
Figure 12. Pressure distribution along the length of the microchannel
X-component of velocity (u) is plotted in the x-y plane passing through centreline at

different locations of x-coordinate as x=1800µm, 2800µm and 4500µm and shown in Figure

13. It is evident from the plot that fluid is continuously accelerating along the flow direction.

Figure 13. U-velocity variation at various cross sections of the microchannel


The y -component of velocity (v) is also plotted in the same plane, and different x locations

are shown in Figure 14. Due to symmetry, velocity is zero at the centre, and because of no

penetration condition, velocity is again zero at the wall. The v velocity achieves maximum

between centre and top and bottom wall.

Figure 14. V- velocity variation at various cross sections of the microchannel

32
The variation of z-component of velocity was plotted in x-z plane containing centreline axis

at different value of x coordinates which is shown in Figure 15. The velocity at the centre is

zero, and velocity at the wall is also zero because of no penetration condition. There is no

slip velocity at the sidewall along the flow direction because the Knudsen number based on

channel width as a characteristic dimension lies in the continuum regime for present case.

Figure 15. W- velocity variation at various cross-sections of the microchannel


The plots of different components of velocity of flow through trapezoidal cross-section

channel shows that the y and z-component of velocity are zero at centre and wall, achieving

maximum between centre and wall. The streamwise velocity is maximum at the centre and

has slip conditions on the top and bottom wall. In the plot, it is evident that each velocity

profile is continuously growing in the flow direction. This shows that this 3D flow is

developing in nature. Furthermore, an increase in lateral component and slip velocity in the

flow direction leads to shifting of mass flow toward the wall.

7. Conclusion

Experimental and numerical simulations was carried out to model the rarefied gas flow

through trapezoidal microchannel with inlet and outlet plenums with second order exact

analytical solution used for slip profile. The conclusions are summarised as follows.

1. The new exact analytical solution for the slip velocity is developed using a highly

nonlinear pressure gradient derived from the pressure equation. The pressure

gradient is obtained by solving the nonlinear pressure equation without any

modification in the pressure expression.

2. The analytically developed slip profile is implemented on walls of channel into

Fluent using a user-defined function. The current three-dimensional modelling

33
technique developed is validated with the numerical solution and experimental data

of up to Kno=0.47 the deviation of 3D simulation result with Colin’s experimental

result is 0.59%.

3. This model can be used for 3 dimensional modelling of rarefied flow in transition

regime.

4. The average deviation of present experimental (Kno=0.1028) and simulation data for

inlet pressure using the general second-order slip boundary condition with slip

coefficient (C1=1.18, C2=0.34.) is 0.51%. The simulations results agree well with the

experimental data.

5. The pressure profile plotted along the centreline of the microchannel shows

nonlinear variation. The velocity profile variation in the streamwise direction

indicates that the flow is developing in nature.

34
Nomenclature
A Constants
ar An aspect ratio of the microchannel (H/W)
B, C Constants
C1 and Slip coefficient
C2
D, E Constants
H Depth of microchannel [m]
Kn Local Knudsen number
Kno Outlet Knudsen number
L Length of microchannel [m]
l The characteristic dimension of the system [m]
M Mach number
p Local pressure [Pa]
pi
P
Pressure ratio po
p Nondimensional pressure
pi Inlet pressure [Pa]
Outlet pressure [Pa]
R Gas constant []

T Gas temperature [K]

u velocity component in the streamwise direction [m/sec]


u Area average streamwise velocity [m/sec]
u Nondimensional velocity in the streamwise direction
Slip velocity
us
us Nondimensional slip velocity
uw Nondimensional wall velocity
Velocity in the y-direction [m/sec]
v

v Nondimensional velocity in the y-direction
W Width of the microchannel (average of top and bottom) [m]
X X-axis
x Non-dimensional coordinate w.r.t channel length L
Y Y-axis
y Nondimensional y coordinate w. r.t channel height H
Z Z-axis
Greek Symbols
 The partial derivative of quantity w.r.t variable x, y, z
 Mean free path of gas molecules
 Gas molecular viscosity (Ns m-2)
 Tangential momentum accommodation coefficient
 Gas molecular viscosity (Ns m-2)
 Nondimensional number (ratio of channel height and length)
 Density of gas (kg m-3)

35
δ Deviation in experimental and simulated data
Subscript
i inlet
o Outlet
rms Root mean square
s Surface
w Wall

36
Footnotes
Hariom Saran Singh email: om.saran7@gmail.com Department of Mechanical Engineering Indian

Institute of Technology New Delhi 110016 New Delhi India.

37
8. Reference

[1] M. Gad-el-Hak, The MEMS Handbook, Boka Raton. CRC Press, 1997.

[2] C. Shen, Rarefied gas dynamics: fundamentals, simulations and micro flows. Berlin:

Springer Verlag Berlin Heidelberg, 2005.

[3] M. Gad-el-Hak, “The Fluid Mechanics of Microdevices—The Freeman Scholar

Lecture,” J. Fluids Eng., vol. 121, no. 1, p. 5, 1999, doi: 10.1115/1.2822013.

[4] C. Cercignani, M. Lampis, and S. Lorenzani, “Variational approach to gas flows in

microchannels,” Phys. Fluids, vol. 16, no. 9, pp. 3426–3437, 2004, doi:

10.1063/1.1764700.

[5] H. Xue, H. M. Ji, and C. Shu, “Analysis of micro-Couette flow using the Burnett

equations,” Int. J. Heat Mass Transf., vol. 44, pp. 4139–4146, 2001.

[6] B. Y. Cao, M. Chen, and Z. Y. Guo, “Rarefied gas flow in rough microchannels by

molecular dynamics simulation,” Chinese Phys. Lett., vol. 21, no. 9, pp. 1777–1779,

2004, doi: 10.1088/0256-307X/21/9/028.

[7] S. N, “A new analytical solution of microchannel Gas flow,” J. Micromechanics

Microengineering, vol. 1695, no. 17, 2007, doi: 10.1088/0960-1317/17/8/036.

[8] J. C. Harley, “Gas Flow in Micr-Channels Gas Flow in Micr-Channels,” J. Fluid

Mech., vol. 28, pp. 257–274, 1995.

[9] E. B. Arkilic, M. A. Schmidt, and K. S. Breuer, “Gaseous slip flow in long

microchannels,” J. Microelectromechanical Syst., vol. 6, no. 2, pp. 167–178, 1997,

38
doi: 10.1109/84.585795.

[10] E. B. Arkilic, K. S. Breuer, and M. A. Schmidt, “Mass flow and tangential

momentum accommodation in silicon micromachined channels,” J. Fluid Mech., vol.

437, pp. 29–43, 2001, doi: 10.1017/S0022112001004128.

[11] Y. Zohar et al., “Subsonic gas flow in a straight and uniform microchannel,” J. Fluid

Mech., vol. 472, pp. 125–151, 2002, doi: 10.1017/S0022112002002203.

[12] J. C. Shih et al., “Monatomic and Polyatomic Gas Flow Through Uniform

Microchannels,” ASME, vol. DSC-59, pp. 197–203., 1996.

[13] S. Takuto, A., Soo, K. M., Hiroshi, I., and Kenjiro, “An Experimental Investigation of

Gas Flow Characteristics in Microchannels,” 2000, pp. 155–161.

[14] S. E. Turner, W. Hall, O. J. Gregory, and C. Hall, “An Experimental Investigation of

Gas Flow in Microchannels,” J. Heat Transfer, vol. 126, no. October 2004, pp. 753–

763, 2004, doi: 10.1115/1.1797036.

[15] S. Hsieh, H. Tsai, C. Lin, C. Huang, and C. Chien, “Gas flow in a long

microchannel,” J. Heat Mass Trens., vol. 47, pp. 3877–3887, 2004, doi:

10.1016/j.ijheatmasstransfer.2004.03.027.

[16] W. A. Zahid, Y. Yin, and K. Zhu, “Couette Poiseuille flow of a gas in long

microchannels,” Microfluid. Nanofluid, vol. 3, pp. 55–64, 2007, doi: 10.1007/s10404-

006-0108-5.

[17] W. A. Ebert and E. M. Sparrow, “Slip Flow in Rectangular and Annular Ducts,” J.

Basic Eng., vol. 87, no. 4, pp. 1018–1024, 1965, doi: 10.1115/1.3650793.

39
[18] G. L. Morini, M. Spiga, and P. Tartarini, “The rarefaction effect on the friction factor

of gas flow in microchannels,” Superlattices Microstruct., vol. 35, no. 3–6, pp. 587–

599, 2004, doi: 10.1016/j.spmi.2003.09.013.

[19] G. L. Morini, Y. Yang, H. Chalabi, and M. Lorenzini, “A critical review of the

measurement techniques for the analysis of gas microflows through microchannels,”

Experimental Thermal and Fluid Science, vol. 35, no. 6. pp. 849–865, 2011, doi:

10.1016/j.expthermflusci.2011.02.005.

[20] C. S. Chen, S. M. Lee, and J. D. Sheu, “Numerical analysis of gas flow in

microchannels,” Numer. Heat Transf. Part A Appl., vol. 33, no. 7, pp. 749–762, 1998,

doi: 10.1080/10407789808913964.

[21] Y. Pong, K. C., Ho, C., Liu, J., and Tai, “Non-Linear Pressure Distribution in

Uniform Microchannels.,” . Appl. Microfabr. to fluid Mech., vol. FED-197, pp. 51–

56.

[22] S. Roy, R. Raju, H. F. Chuang, B. A. Cruden, and M. Meyyappan, “Modeling gas

flow through microchannels and nanopores,” J. Appl. Phys., vol. 93, no. 8, pp. 4870–

4879, 2003, doi: 10.1063/1.1559936.

[23] R. Raju and S. Roy, “Hydrodynamic prediction of high speed microflows,” in 33rd

AIAA Fluid Dynamics Conference and Exhibit, no. June, doi: 2003, 10.2514/6. 2003-

4010.

[24] C. Cai and I. D. Boyd, “Compressible gas flow inside a two-dimensional uniform

microchannel,” J. Thermophys. Heat Transf., vol. 21, no. 3, pp. 608–615, 2007, doi:

10.2514/1.29362.

40
[25] C. S. Chen, “Numerical method for predicting three-dimensional steady compressible

flow in long microchannels,” J. Micromechanics Microengineering, vol. 14, no. 7,

pp. 1091–1100, 2004, doi: 10.1088/0960-1317/14/7/032.

[26] A. Agrawal and A. Agrawal, “Three-dimensional simulation of gaseous slip flow in

different aspect ratio microducts,” Phys. Fluids, vol. 18, no. 10, 2006, doi:

10.1063/1.2354546.

[27] V. Jain and C. X. Lin, “Numerical modeling of three-dimensional compressible gas

flow in microchannnels,” J. Micromechanics Microengineering, vol. 16, no. 2, pp.

292–302, 2006, doi: 10.1088/0960-1317/16/2/014.

[28] N. Jeong, C. L. Lin, and D. H. Choi, “Lattice Boltzmann study of three-dimensional

gas microchannel flows,” J. Micromechanics Microengineering, vol. 16, no. 9, pp.

1749–1759, 2006, doi: 10.1088/0960-1317/16/9/001.

[29] R. Schamberg, “The fundamental differential equations and the boundary conditions

for high speed slip-flow, and their application to several specific problems,” Ph.D.

Thesis, Californiya Institute of Technology Pasadena Californiya, 1947.

[30] S. A. S. Paul A. Chambre, “Flow of Rarefied Gases,” in Fundamental of gas

Dynamics, N.J. Princetcenton University Press Princeton, 1958.

[31] R. G. Deissler, “An analysis of second-order slip flow and temperature-jump

boundary conditions for rarefied gases,” Int. J. Heat Mass Transf., vol. 7, no. 6, pp.

681–694, 1964, doi: 10.1016/0017-9310(64)90161-9.

[32] C. Cercignani and A. Daneri, “Flow of a rarefied gas between two parallel plates,” J.

Appl. Phys., vol. 34, no. 12, pp. 3509–3513, 1963, doi: 10.1063/1.1729249.

41
[33] N. Dongari, A. Agrawal, and A. Agrawal, “Analytical solution of gaseous slip flow in

long microchannels,” in International Journal of Heat and Mass Transfer, 2007, vol.

50, no. 17–18, pp. 3411–3421, doi: 10.1016/j.ijheatmasstransfer.2007.01.048.

[34] A. K. Sreekanth and M. Field, “Slip Flow Through Long Circular Tubes,” in

Proceedings of the 6th International Symposium on Rarefied Gas Dynamics, pp. 667–

680, 1969.

[35] C. Aubert and S. Colin, “High-order boundary conditions for gaseous flows in

rectangular microducts,” Microscale Thermophys. Eng., vol. 5, no. 1, pp. 41–54,

2001, doi: 10.1080/108939501300005367.

[36] S. Colin, P. Lalonde, R. Caen, P. Lalonde, and R. Caen, “Validation of a Second-

Order Slip Flow Model in Rectangular Microchannels,” J. Heat Transf. Eng., vol. 25,

no. 3, pp. 23–30, 2004, doi: 10.1080/01457630490280047.

[37] J. Maurer, P. Tabeling, P. Joseph, and H. Willaime, “Second-order slip laws in

microchannels for helium and nitrogen,” Phys. Fluids, vol. 15, no. 9, pp. 2613–2621,

2003, doi: 10.1063/1.1599355.

[38] T. Ewart, P. Perrier, I. A. Graur, and J. G. Méolans, “Mass flow rate measurements in

a microchannel, from hydrodynamic to near free molecular regimes,” J. Fluid Mech.,

vol. 584, pp. 337–356, 2007, doi: 10.1017/S0022112007006374.

[39] I. A. Graur, P. Perrier, W. Ghozlani, and J. G. Méolans, “Measurements of tangential

momentum accommodation coefficient for various gases in plane microchannel,”

Phys. Fluids, vol. 21, no. 10, 2009, doi: 10.1063/1.3253696.

[40] A. Beskok and G. E. Karniadakis, “Report : A model for flows in channels, pipes, and

42
ducts at micro and nano scales,” Microscale Thermophys. Eng., vol. 3, no. 1, pp. 43–

77, 2010, doi: 10.1080/108939599199864.

[41] R. W. Barber and D. R. Emerson, “Challenges in modeling gas-phase flow in

microchannels: From slip to transition,” Heat Transf. Eng., vol. 27, no. 4, pp. 3–12,

2006, doi: 10.1080/01457630500522271.

[42] V. Hemadri, A. Agrawal, and U. V. Bhandarkar, “Determination of tangential

momentum accommodation coefficient and slip coefficients for rarefied gas flow in a

microchannel,” Sadhana - Acad. Proc. Eng. Sci., vol. 43, no. 10, 2018, doi:

10.1007/s12046-018-0929-4.

[43] K. Srinivasan, P. M. V. Subbarao, and S. R. Kale, “Experimental and Numerical

Studies on Gas Flow Through Silicon Microchannels,” J. Fluids Eng. Trans. ASME,

vol. 139, no. 8, pp. 1–12, 2017, doi: 10.1115/1.4036249.

[44] K. Srinivasan, P. M. V. Subbarao, and S. R. Kale, “A comprehensive experimental

and numerical study on gas flow through microchannels from slip to free-molecular

regimes,” J. Micromechanics Microengineering, vol. 28, no. 9, 2018, doi:

10.1088/1361-6439/aac4d5.

[45] K. Srinivasan, P. M. V. Subbarao, and S. R. Kale, “Studies on Gas Flow through

Smooth Microchannel Surface–Fabrication, Characterization, Analysis, and

Tangential Momentum Accommodation Coefficient Comparison,” Heat Transf. Eng.,

vol. 41, no. 6–7, pp. 607–621, 2020, doi: 10.1080/01457632.2018.1546952.

[46] N. G. Hadjiconstantinou, “Comment on Cercignani’s second-order slip coefficient,”

Phys. Fluids, vol. 15, no. 8, pp. 2352–2354, 2003, doi: 10.1063/1.1587155.

43
[47] K. Srinivasan, P. M. V. Subbarao, and S. R. Kale, “Experimental and Numerical

Studies on Gas Flow Through Silicon Microchannels,” Ph.D. dissertation, Dept.

Mech. IIT Delhi, 2009.

[48] S. Colin, “Rarefaction and compressibility effects on steady and transient gas flows in

microchannels,” Microfluid Nanofluid, pp. 268–279, 2005, doi: 10.1007/s10404-004-

0002-y.

44
Acknowledgement
Decalration of Interest
On behalf of all the authors, the corresponding author states that there is no conflict of interest.

45
Tables
Table 6. Dimensions of the microchannel system

Length(µm) Width(µm) Depth(µm) Aspect Number of


ratio(ar) channels

System1 5400 Wtop= 488.7 1.02 0.002 20


Wbot= 487.26
Uncertainty ±15 ±1 ±0.005
Note-Surface roughness of channel ( Rarms =15 nm± 3)

46
Table 7. Percentage deviation between measured and simulated inlet pressure for different
first-order slip coefficients for data of P=1.58.

S.N Inlet Outlet Mass flow Slip Simulated %δ


pressure Knudsen rateṁ (kg/sec) coefficient inlet pressure
Pi(experimental) number (C1) (Pi (simulated (Pa))
(Pa) (Kno)
1 102923 0.1028 2.45×10-11 1.22 105656 2.66
2 102923 0.1028 2.45×10-11 1.21 105654 2.65
3 102923 0.1028 2.45×10-11 1.20 105652 2.65
4 102923 0.1028 2.45×10-11 1.19 105649 2.65
5 102923 0.1028 2.45×10-11 1.18 105647 2.65
6 102923 0.1028 2.45×10-11 1.17 105645 2.64
7 102923 0.1028 2.45×10-11 1.16 105643 2.64
8 102923 0.1028 2.45×10-11 1.15 105640 2.64
9 102923 0.1028 2.45×10-11 1.14 105638 2.64

47
Table 8. Percentage deviation between measured and simulated inlet pressure for a fixed

value of C1 and different values of C2 for data of P = 1.58.

S.N Pi expt (Pa) Kno ṁ (kg/sec) C1 C2 Pi simulated (Pa) %δ


o
1 102923 0.1028 2.45×10-11 1.18 0.69 101854 1.04
2 102923 0.1028 2.45×10-11 1.18 0.34 102591 0.32
3 102923 0.1028 2.45×10-11 1.18 0.20 102855 0.07
4 102923 0.1028 2.45×10-11 1.18 0.15 102979 0.05
5 102923 0.1028 2.45×10-11 1.18 0.13 102986 0.06

48
Table 9. Average percentage deviation between simulated and measured inlet pressure for a

fixed value of C1 and different value of C2 for all six data of Kno= 0.1028.

Second-order slip
coefficients

S.N Kno C1 C2
o
1 0.1028 1.18 0.69 0.84
2 0.1028 1.18 0.34 0.51
3 0.1028 1.18 0.20 0.53
4 0.1028 1.18 0.15 0.53
5 0.1028 1.18 0.13 0.55

49
Table 10. Experimental condition for Nitrogen gas flow

S.No Parameter Range (Mean value)


1 Outlet Pressure 6.51×104 Pa
2 Outlet Knudsen number Kno= 0.1028
3 Temperature 305.5K
4 Pressure ratio 1.24 to 1.96

50
List of figures

Figure 1.

y
Velocity profile

z x

Knudsen Layer
(O) λ

Q
Wall Actual
First order

Figure 16. Knudsen layer near the wall

51
Figure 2.

(i)

(ii)

52
(iii)

Figure 17. Variation of approximate and exact nondimensional pressure distribution (i)
Kn=0.667 presure ratio 1.84 (ii) Kn=0.22 pressure ratio 1.93 (iii) ) Kn=0.90 pressure ratio
1.93.

53
Figure 3

Inlet Plenum Outlet Plenum


y

z x

θ H

Figure 18. Microchannel with inlet and outlet plenum

54
Figure 4

Figure 19. Grid independency test for simulation of present experimental data.

55
Figure 5

(i) (ii)
Figure 20. Grid independency test for first-order validation (i) with Reference [9] and (ii)
with Reference [20].

56
Figure 6

Figure 21. Variation of mass flow rate with inlet pressure for slip condition at = 0.156
reference [9], and comparison with [20], [22].

57
Figure 7

(i)

(ii)

Figure 22. Comparison of (i) Centreline velocity and (ii) transfer velocity along the length of
the channel for pressure ratio 1.34 and 2.70 with reference [23].

58
Figure 8

(i) (ii)
Figure 23. Grid independency test for validation of experimental data of (i) reference [36]
and (ii) reference [48].

59
Figure 9

(i)

(ii)
Figure 24. Comparison of inlet pressure against mass flow rate for (i) channel having an
aspect ratio of 0.055 at Kno= 0.09 (ii) channel having an aspect ratio of 0.01 at Kno= 0.47.

60
Figure 10

Figure 25. Variation of experimental and simulation inlet pressure with mass flow rate for
Kno= 0.1028.

61
Figure 11

Figure 26. Variation of experimental and simulation inlet pressure with mass flow rate for
Kno= 0.1028.

62
Figure 12

Figure 27. Pressure distribution along the length of the microchannel

63
Figure 13

Figure 28. U-velocity variation at various cross sections of the microchannel

64
Figure 14

Figure 29. V- velocity variation at various cross sections of the microchannel

65
Figure 15

Figure 30. W- velocity variation at various cross-sections of the microchannel

66
Figures caption

Notes on contributors

Hariom Saran Singh is research scholar in Mechanical Engineering Department Indian

Institute of Technology Delhi. He received his master’s degree from IIT BHU in 2011.

Earlier he was working in Jaypee University of Engineering and Technology Madhya

Pradesh (Guna) India. He joined IIT Delhi in 2014 and his research work includes

experimental and numerical modelling of rarefied gas flow through straight and complex

microchannels.

Amit Kakkar

He did his Bachelor of Engineering in Mechanical Engineering from Kumaon University,

Nainital in year 2002 (Gold Medalist). He then joined as a faculty member in the

Department of Mechanical Engineering, Maharana Pratap Engineering College, Kanpur. In

Dec. 2007 he joined in the Department of Mechanical Engineering, Indian Institute of

Technology Delhi for pursuing his M.S. (Research) studies in 2010. His main research

interests are in Microfluidics and in fundamentals of Fluid Mechanics.

67
Paruchuri M.V. Subbarao is a professor in Mechanical Engineering Department Indian

Institute of Technology Delhi. He received Phd degree from IIT Kanpur in 1994. His

research interest includes computational and experimental micro fluid mechanics,

computational study of high-speed gas dynamics, turbulent flow fluid mechanics and heat

transfer. He is author 4 book and of more than 100 articles. He is currently Ray W. Herrick

chair professor in Mechanical Engineering Department.

68

You might also like