Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/275087399

Experimental Investigation of Microscale Effects in Perforated Plate


Aerodynamics

Article  in  Journal of Fluids Engineering · December 2013


DOI: 10.1115/1.4024962

CITATIONS READS

9 688

4 authors, including:

Tomasz Lewandowski
Siemens Industry Software, Leuven, Belgium
10 PUBLICATIONS   49 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Definition of a physical model of flow through perforated plates by means of numerical simulations View project

The effect of wood origin on the cutting power for the sawing process View project

All content following this page was uploaded by Tomasz Ochrymiuk on 21 September 2018.

The user has requested enhancement of the downloaded file.


Ryszard Szwaba
e-mail: rssz@imp.gda.pl Experimental Investigation of
Tomasz Ochrymiuk
e-mail: tomasz.ochrymiuk@imp.gda.pl Microscale Effects in Perforated
Tomasz Lewandowski
e-mail: tomasz.lewandowski@imp.gda.pl
Plate Aerodynamics
Institute of Fluid Flow Machinery, This paper contains an extensive analysis of the flow in microholes based on an experi-
Polish Academy of Sciences, mental investigation. Experiments of the gas flow past a perforated plate with microholes
Fiszera 14, (110lm) were carried out. A wide range of pressure differences between the inlet and the
Gdansk PL80-952, Poland outlet were investigated for that purpose. Two distinguishable flow regimes were
obtained: the laminar flow with the slip effects and the turbulence transition regime for a
very low Reynolds number. The results are in good agreement with the theory, simula-
Justyna Czerwinska1 tions, experiments for large scale perforated plates, and compressible flows in micro-
Artorg Center,
tubes. The relation between the mass flow rate and the Knudsen, Reynolds, and Mach
University of Bern,
numbers for the laminar and transitional regime was obtained. It is a quadratic function
Murtenstrasse 50,
of the Reynolds and Knudsen numbers (ReKn) based on the hole’s diameter. The value of
Bern CH 3010, Switzerland
the first order tangential momentum accommodation coefficient was estimated. It shows a
e-mail: justyna.czerwinska@artorg.unibe.ch
strong relation to the inlet Knudsen number. [DOI: 10.1115/1.4024962]

1 Introduction Another characteristic parameter of the gas relates its molecular


properties to the continuum quantities. The mean free path repre-
Perforated plate aerodynamics has been extensively studied over
sents an average distance between the molecules of the gas and is
many decades by experiments [1–8] and theory [9–13]. The appli-
defined as
cation interest has varied: from sound absorption [14], microfilters
and separators [5], and heat exchangers [15], to the method of pro- pffiffiffiffiffiffiffiffiffi
duction of a uniform turbulent flow [2] and a new concept of a k ¼ kk lðTÞ 2RT =P (2)
blade turbine cooling technique [16]. The concept of microchannel pffiffiffi
cooling for the gas turbine blades is a natural application of ther- where kk ¼ p=2. The viscosity in that relation also depends on
modynamics and heat transfer. It allows us to accomplish two the temperature and can be described as
goals: first, to spread out the cooling network in a series of smaller
and highly distributed channels in order to provide a much better lðTÞ ¼ ðT=Tref Þx (3)
uniformity of the cooling and thermal gradients and second, to
bring the cooling fluid closer to the blade surface and create a The nondimensional parameters determining the relevant scales
more efficient heat transfer. In the past, the film jet used in the film for the flow are the Mach number ðMa ¼ U=aÞ, the Reynolds
cooling process was so thick that the amount of cooling air used number based on the hole’s diameter and average velocity
could effectively reduce the performance of a gas turbine engine ðRe ¼ qUD=lÞ, and the Knudsen number based on the hole’s di-
[17]. This paper is focused on the investigation of the microscale ameter ðKn ¼ k=DÞ. After simple arithmetical transformations the
effects from the global macroscale experimental data. For that pur- relation between nondimensional parameters is as follows
pose, the experiments were performed in air on a perforated plate pffiffiffiffiffiffi
with a perforation of 5:2% and holes 110 lm in diameter. Various ReKn ¼ 2jkk Ma (4)
ranges of pressure differences were studied. This allowed us to
obtain the characteristics of two distinguishable regimes: a laminar Therefore, the nondimensional numbers relate to each other and
one with an evident first order slip boundary condition and the cannot be considered separately.
transition to turbulence at around Re ¼ 60. The studies were per-
formed for various Mach numbers from ðMa  0:01Þ up to the
critical conditions (choking flow in the microholes). 2 Measurement Technique
The theory of compressible gas flow in channels has been stud- The experimental investigations were carried out in the test sec-
ied by continuum and molecular approaches and can be found in tion shown in Fig. 1. The flow direction is from the left to the right
various textbooks; for example, continuum [18], molecular [19], as indicated by an arrow. The flow is a result of an imposed pres-
and simple analytical relations [20]. Here, we briefly describe the sure difference between the ambient condition and pressure in the
basic theory required to extract information from experimental vacuum tanks, which was placed behind valve 7. The ambient
data. The relation between the gas parameters is given as an equa- condition corresponds to the air parameters at the laboratory
tion of state (P=q ¼ RT). Based on that equation the speed of space, e.g., temperature about 20  22  C and atmospheric pres-
sound can be defined as sure. The pressure in the vacuum tanks was equal to 1:6 kPa. The
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffi capacity of the vacuum tanks was 120 m3 , since it allows a long
a ¼ jRT ¼ jP=q (1) enough time for measurements at very low pressures where the
mass flow is only about 12 standard liters per minute (SLPM).
The air goes from the ambient flow through the flow meter 1,
through the control valve 2 and compensatory chamber 3 to frame
1
Corresponding author. 4, where the membrane with microholes is mounted. Downstream
Contributed by the Fluids Engineering Division of ASME for publication in the
JOURNAL OF FLUIDS ENGINEERING. Manuscript received January 28, 2013; final
of the membrane the flow goes through another chamber 5, a sec-
manuscript received June 27, 2013; published online September 23, 2013. Assoc. ond valve 6 (which controls the flow condition downstream of the
Editor: Prof. Ali Beskok. microholes), and the cut-off valve 7 to the vacuum tanks. The

Journal of Fluids Engineering Copyright V


C 2013 by ASME DECEMBER 2013, Vol. 135 / 121104-1

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 04/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 2 Schematic representation of the perforated plate
(D 5 110 lm, L 5 500 lm) and the microscopic picture

comparison with the global parameters and other experimental


data is made. Finally, an estimation of the microscale effects, such
as the velocity slip, is discussed.

3.1 General Parameters. The results are presented in a se-


ries of inlet pressure variations of 2:5; 3; 4, 1st and 2nd series,
5; 10; 20; 40 and 55 kPa. Nondimensional outlet parameters are
Fig. 1 Top: view of the test section: (1) flow meter, (2) control plotted in Fig. 3.
valve, (3) compensatory chamber, (4) membrane location, (5) The Reynolds number is defined by the hole’s diameter and the
second chamber, (6) second valve, and (7) cut-off valve to the
vacuum tanks. Bottom: schematic representation of the
outlet parameters. The Mach number in the hole is a function of
experiment. the Knudsen and Reynolds numbers (see Eq. (4)). As can be noted

change of the stagnation parameters (pressure and density) was


obtained by means of control valve 2. However, the pressure
downstream of the microholes was controlled by means of valve
6. The particular characteristics can be measured throughout an
adequate manipulation of these two control valves.
The measurement of the stagnation parameters was located in
compensatory chamber 3. The pressure was measured by means
of a Prandtl probe with a Kulite pressure transducer with the full
scale (FS) accuracy of 0:1%, e.g., 60:1 kPa. The temperature was
measured by means of the thermocouple element and was, on av-
erage, 203 K with an accuracy of 0:1 K. The static pressure
upstream and downstream of the membrane was measured by
means of a pressure scanner with the FS accuracy of 0:05%, e.g.,
60:05 kPa. The mass flow rate was measured by means of three
different laminar flow meters, depending on the range of the mass
flow (Alicate: 20, 100, and 1500 SLPM; at the ambient pressure
equal to 101:3kPa and the temperature of 298 K, the air density
for these conditions was equal to 1:184 kg=m3 ). The accuracy of
these flow meters was equal to 60:01 SLPM of the measured
value. The measurements were comprised of the barometric pres-
sure; stagnation parameters upstream of the microholes, static
pressure upstream and downstream of the microholes, and the
mass flow rate. The pressure versus mass flow characteristics at
constant pressure upstream of the measurement membrane was
obtained. The following upstream microholes pressures were cho-
sen: 2:5; 3; 4; 5; 10; 20; 40; and 55 kPa;
The membrane diameter was 60 mm. The membrane was pro-
duced from steel. The diameter of a single microhole was equal to
110 lm and the length was equal to 500 lm. The membrane perfo-
ration, e.g., the ratio of the surface of all holes to the whole mem-
brane surface was equal to 5:2%; which means that the membrane
contained about 15; 471 microholes. A detailed view of the mem-
brane is shown in Fig. 2. The accuracy of the estimation of the
microhole diameter and length is 0:01 mm and 0:002 mm,
respectively.

3 Results and Discussion


The analysis of the results is divided into several parts. In the Fig. 3 Nondimensional number dependencies: (a) the Mach
first part, the experimental data are presented. In the second sec- number as a function of the Reynolds number, and (b) the
tion, a global analysis of the results is made. In the next part, a Knudsen number as a function of the Reynolds number

121104-2 / Vol. 135, DECEMBER 2013 Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 04/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 4 Mass flow rate as a function of the outlet: (a) Reynolds
number, and (b) Knudsen number Fig. 5 Pressure drop distribution as a function of: (a) the
Reynolds and Knudsen numbers at the outlet, and (b) the mass
flow rate
from Fig. 3(a), there are two distinguishable flow regimes: one for
low Reynolds numbers (Re < 10) and one for high Reynolds
numbers (Re > 10). The latter is related to the turbulence transi- laminar flow DP ¼ Pin  Pout  U and for the transition to turbu-
tion and the evidence for that is provided later. In Fig. 3(b), the lence regime DP ¼ Pin  Pout  U 1:75 (see Ref. [20]). Figure 7
Knudsen and Reynolds numbers are plotted. The Knudsen number shows the mass flow rate changes in relation to the pressure drop.
is based on the hole’s diameter D and the outlet parameters. In The characteristics are similar to the velocity changes. Table 1
that graph, both of the regimes behave similarly and the changes shows the best fit values for the velocity, where the function is as
are related to the adiabatic transition between the states of the gas. follows
Figure 4 shows the mass flow rate as a function of the Reynolds
and Knudsen numbers. The nature of the curves is not surprising DP ¼ AU B ; A; B are fitting coefficients (5)
since the velocity from the mass flow rate was taken into account
to estimate the Reynolds number. Hence, the linear dependency It can be seen that the laminar flow has, indeed, a linear velocity
between the mass flow rate Qh and the Reynolds number Reout is dependency, while the transition to the turbulence regime is dif-
observed and plotted in Fig. 4(a). The relation between the Knud- ferent. Figure 6(b) shows an enlargement of the laminar part. It
sen number Knout and the mass flow rate Qh in Fig. 4(b) behaves can be noted that the results obtained in our experiments are in
in a similar way. agreement with the literature.
In Fig. 5 the pressure difference is plotted in two ways: in rela- The point of this work was to study microscale effects and evi-
tion to the multiplication factor of the Reynolds and Knudsen dence of these effects is much more difficult to find in the turbu-
numbers and in relation to the mass flow rate. In both cases the lence transitional regime due to the insufficient theoretical
two flow regimes are clearly seen: the laminar one and the transi- description. Hence, our further investigation is mainly restricted
tion to turbulence. It is shown in the following text that the results to the laminar flow. Figure 6(c) shows the mass flow rate as a
for the laminar case fit well with the microchannel experiments function of the pressure difference. It can be noted that the com-
and simulations for the low Knudsen number regime [21]. pressibility effects are large in the transitional regime. Addition-
In Fig. 6 the velocity in the hole is plotted as a function of the ally, it can be seen from our experiments that the transition occurs
pressure difference. This standard plot was first proposed by more rapidly than for a pipe of the same diameter. We obtained a
Hagen [22] and shows most clearly the transition to turbulence transitional Reynolds number around Re ¼ 60 and, in some cases,
which occurs for cases for the inlet pressure of Pin ¼ 10  55 kPa. it has occurred for Re ¼ 81:5 for a pipe [23]. For pressure meas-
The velocity of the flow should obey the following law for the urements along the microchannel in the air in a specially designed

Journal of Fluids Engineering DECEMBER 2013, Vol. 135 / 121104-3

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 04/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Table 1 Coefficients for the velocity pressure relation used to
estimate transition to turbulence (see Eq. (5))

Pin ðkPaÞ 2:5  5:0 10:0 20:0 40:0 55:0

A 0.62 0.04 0.02 0.02 0.04


B 0.99 1.35 1.54 1.71 1.57

tion to the transitional Reynolds number; hence, the discrepancy


between our results and the experiments [24]. Our channels are
four times shorter (500 lm), therefore the transition occurs for
smaller Reynolds numbers. However, a different experimental
setup is needed to precisely estimate it.
Finally, Fig. 8 shows the global characteristic behavior. All of
the data are arranged in one line; mainly, the normalized mass
flow rate Qh =Q as a function of the outlet Knudsen and Reynolds
numbers Reout Knout . The normalization is performed by the theo-
retical critical mass flow rate given by Eq. (6). This curve has the
equation

Qh =Q ¼ AReout Kn2out þ BReout Knout þ C (6)

Again, two Reynolds number-dependent regimes can be noted in


this figure. Here, A, B, and C are the fitting parameters and are
given in Table 2.
In this section experimental results were plotted in various con-
figurations in order to estimate the major contributions in flow
behavior. It was proven that some cases between
Pin ¼ 10  55 kPa were in the turbulence transition region and
some cases could be considered as a laminar flow. Additionally,
the best fit for all of the data was shown as a function of a normal-

Fig. 6 Transition to turbulence in plots of the velocity as a


function of the pressure drop: (a) the laminar and turbulence
transition regime and (b) only the laminar regime

microchip [24], the transition to turbulence occurred for the


Reynolds number Re ¼ 2300 for hydraulic diameters between
25  100 lm and the channel length was 2000 lm. As was proved
from the theory in Ref. [23] the pipe length has a significant rela-

Fig. 8 Normalized mass flow rate to the critical mass flow rate
as a function of the Knudsen and Reynolds numbers

Table 2 Coefficients for the mass flow rate ReKn dependency


(see Eq. (6))

Pin ðkPaÞ A B C

3:0 2.2371 0.7649 0.0005


4:0 Number 1 1.5838 0.7799 0.0007
4:0 Number 2 1.6002 0.7797 0.0008
5:0 1.1829 0.7794 0.0015
10:0 0.3329 0.887 0.0009
20:0 0.2618 0.8381 0.0518
40:0 0.2946 0.9911 0.0262
Fig. 7 Transition to turbulence and compressibility effects 55:0 0.2719 0.9745 0.027
shown as changes of the mass flow rate

121104-4 / Vol. 135, DECEMBER 2013 Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 04/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


ized mass flow rate and an outlet Reynolds-Knudsen number. This
relation proves that the flow is greatly affected by the compressi-
bility effects and in a less pronounced manner by viscous effects
or microscale-related issues such as the velocity slip.

3.2 Perforated Plate Aerodynamics. Perforated plate aero-


dynamics has been studied for many years, mostly focusing on
plates with large holes. Both the laminar and turbulent regimes
have been considered. Some specific examples can be found in
the literature; for example, the pressure loss coefficient over plate
thickness [25], the turbulent flow over a perforated plate [2], and a
very high Reynolds number (Re > 10; 000) [26].
In this section, however, we will compare the obtained results
with two specific papers. One considers various large scale holes
in different types of plates [27] and the other is a study of various
shapes of microholes on a large plate [5].
The mass flow rate is given by the equation from the empirical
studies in Ref. [27]

Q ¼ FSqh Uh (7)

which, by implementing the stagnation parameters, transforms to


rffiffiffiffiffiffiffiffi
Mah j
Q ¼ FS   P 0 (8)
j1 2 jþ1=2ðj1Þ RT0
1 þ Ma 2 h

or, in the form dependent on the Knudsen and Reynolds numbers


rffiffiffiffiffiffiffiffi
Reout Knout j
Q ¼ FS pffiffiffiffiffiffi  P
jþ1=2ðj1Þ 0 (9)
j1 2 2 RT0
2jkk 1 þ 4jk2 Reout Knout
k

The experimental relation derived in Ref. [27] leads to

Mah ¼ 1:2ðDP=P0 Þ0:55 ; Reout Knout ¼ CðDP=P0 Þ0:55 (10)


pffiffiffiffiffiffi
where C ¼ 1:2 2jkk ¼ 1:736. Equation (9), is plotted in Fig.
9(a). Two regimes are clearly seen. The transition to turbulence
fits well to the relation derived from the experiments in Ref. [27].
The laminar flow behaves differently. The exact fitting coeffi- Fig. 9 Comparison with the other experiments for the perfo-
cients are presented in Table 3, where the function is given by the rated plates: (a) perforated plates in parallel flow [27] and (b)
equation perforated plates with various microholes [5]

Mah ¼ AðDP=P0 ÞB (11) Another equation has been proposed in Ref. [4]
 
It can be seen that the coefficient for the laminar flow is twice as 3lQh 16 L
large as the one for the transition to turbulence BL  2BT . This DP ¼ þ 1 ½1  f ðbÞ (14)
ðD=2Þ3 3p D
means that at lower Reynolds numbers, for the same pressure
drop, there is an increase in the importance of viscous and rarefac-
tion effects. The results confirm that, despite the small scale of the It can be noted that all of the approaches have a very similar
holes in our experiments, the global plate aerodynamics fits very form. A global comparison of the previously presented equations
well to the large scale investigation. and experiments on plates having various microhole shapes has
There are some other theoretical approaches for studying perfo- been performed in Ref. [5]. For a comparison with our experi-
rated plate aerodynamics [28] ments, we chose to plot the nondimensional equation
 
lQh 16 L
DP ¼ þ 3 (12) Table 3 Coefficients for the Mach number pressure drop de-
ðD=2Þ3 p D pendency (see Eq. (11))

where the former term corresponds to the single long channel flow Pin ðkPaÞ A B
and the latter is a correction for the entry effects. For the plate
3:0 0.18 1.01
[12] it is 4:0 : Number 1 0.25 0.96
4:0 : Number 2 0.24 0.95
3lQh X
1
5:0 0.32 0.93
DP ¼ 3
½1  f ðbÞ; f ðbÞ ¼ Ci bð2iþ1Þ=2 (13)
ðD=2Þ 10:0 1.16 0.65
i¼1
20:0 1.30 0.53
40:0 1.43 0.50
where C1 ¼ 0:3389, C2 ¼ 0:1031, C3 ¼ 0:0558, and 55:0 1.95 0.89
C4 ¼ 0:0364.

Journal of Fluids Engineering DECEMBER 2013, Vol. 135 / 121104-5

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 04/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


  
DP 2 16 L b
K¼1 2 ¼b p þ 3 C1 þ C2 (15)
2qUin
3p D Uin D

where C1 ¼ 10:7 and C2 ¼ 0:27. Our experiments and also line


corresponding to Eq. (15) with an adjustment for a circular lattice
instead of a square one, are shown in Fig. 9(b). It can be seen that
the laminar regime fits very well to that empirical equation. The
transition to the turbulence part does not fit that well. This equa-
tion is derived in order to focus on the viscous effects (the com-
pressibility influence was neglected) and therefore has better
agreement for the laminar flow.
In this section two distinguishable experimental comparisons
for a flow past a perforated plate were performed. One considers
the equation derived to account for the compressibility effects
(see Eq. (10)) and the other is described by the equation for the
viscous flow resistance See Eq. (15)). In both limits our experi-
ments fit very well for the adequate relevant case.
3.3 Microscale Effects. In the previous section the global
characteristic of the perforated plate was studied. In this paragraph
we will focus on the microscale effects of a single microhole in
order to establish the presence and influence of the velocity slip
on the measured mass flow rate. The microscale effects have been
studied in various ways, both in single channels and in perforated
plates (microfilters or membranes). Some numerical investigations
have been performed with a continuum flow and a slip boundary
[29–32], with the Direct Simulation Monte Carlo method
[19,33,34], the linearized Boltzmann equation [35], other compu-
tational discrete methods [36,37], or other types of theoretical
models [38,39].
The experimental investigation has been performed: in long
microtubes [40,41], for isothermal microchannels [42], in square
microchannels with heating [43], for subchoking effects in
long microchannels [44], for flow through micro-orfices [3],
measurements for a high Knudsen number [45,46], and the
detailed determination of the accommodation coefficients from
experiments [47]. The review for various velocity slip and temper-
ature jump effects can be found in Ref. [48].
It has recently been shown that the compressible microchannel
flow is very sensitive to the compressibility effects, even for very Fig. 10 Comparison with the experimental data in Ref. [50]: (a)
low compressibility and low Knudsen numbers. A theoretical pre- the pressure drop as a function of the Reynolds and Knudsen
diction has been found and compared with many experimental numbers and (b) the mass flow rate as a function of the Knud-
data for microflows [49] with the assumption of no slip boundary sen number
condition. Slip effects in such a case have been almost
undistinguishable.
In our experiments we have observed the significant influence A similar equation has been derived for the temperature on the
of the compressibility effects and, despite that fact, we are provid- wall [53]. In general, for a high Knudsen number, the condition
ing a way to analyze velocity slip effects. First, a comparison with can be written in a more general form
the experimental data in a microtube was made [50]. Figure 10(a)
shows the mass flow rate as a function of the pressure drop. In the  
2  rv @U k2 @ 2 U k3 @ 3 U
experiments of Ref. [50] the pressure drop has been small and U g  Uw ¼ k jw þ j w þ j w þ ::: (17)
almost constant. In our case, it was different. Additionally, the rv @y 2! @y2 3! @y3
experiments in question were performed for a constant tempera-
ture. Hence, in Fig. 10(b) the mass flow rate as a function of the The Knudsen numbers in our experiments are small enough that
Knudsen number has a very different characteristic than in our ad- we can neglect higher orders than the first one
iabatic flow with various pressure drops. However, for the small We performed the estimation for the slip effect’s influence on
pressure difference between the inlet and outlet, our experiments the mass flow rate for a laminar flow in two ways. The former is a
get close enough to the isothermal curve. comparison of the data with the existing theory. The latter is a com-
For another comparison, we plotted the data normalized to the parison with the numerical solutions. In Fig. 11(a) we have plotted
critical mass flow rate. Figure 11(a) shows our experiments, the our experimental data, the data from experiments in a microtube
data of Ref. [50], and the numerical simulations [51] for a three- presented in Ref. [50], and three-dimensional Navier–Stokes
dimensional microchannel with no-slip and slip boundary condi- simulations for a compressible gas flow with no-slip and a 1st-order
tions. It can be seen that the relations fit very well to the previ- slip boundary condition [51]. It can be noted that the results fit well
ously described curves (see Eq. (6)). with the one line and, in particular the no-slip and slip
The gas-wall interaction on a microscale has been proposed in numerical solutions vary indistinguishably between themselves.
Ref. [52] in the form of a boundary condition for the velocity In Fig. 11(b) we compare the theoretical prediction for the iso-
thermal slip flow with our experimental data. The selected data,
which are the closest to the isothermal solution (see Fig. 10(b))
2  rv @U are compared with the predictions of a compressible channel flow
Ug  Uw ¼ k j (16)
rv @y w with: the no-slip boundary condition, the 1st-order slip, and the

121104-6 / Vol. 135, DECEMBER 2013 Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 04/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 12 Normalized mass flow rate between the slip and no-
slip value as a function of the outlet Reynolds and Knudsen
numbers for laminar flow. The plot shows our experiments and
3-D compressible Navier–Stokes simulations with 1st-order slip
boundary conditions [51].

The mass flow rate is normalized in such a way that the plate
resistance is irrelevant, since the adequate slip and no-slip solution
is taken to account. The value is as follows

Q ¼ ðQexp =Qns  Qexp =Qs ÞQns =Qs (19)

which, for simulations where Qexp ¼ Qs , gives ðQs  Qns Þ=Qs .


The normalized mass flow rate is shown as a function of the Reyn-
olds and Knudsen numbers. It can be noted from numerical simu-
lations where the slip was fixed to the first-order solution and for
the velocity accommodation coefficient rv ¼ 0:3 that the differ-
ence in mass flow rate is almost constant and equals 1%. It
depends on the Reynolds number only, but it does not depend on
the shape of the inlet. The simulations were performed for a sharp
Fig. 11 Normalized mass flow rate as a function of the Reyn- and smooth entrance and the difference was almost undistinguish-
olds and Knudsen numbers at the outlet. The plot contains our able (see Ref. [51] for details). For the experimental part it means
experimental data, experiments [50], (a) 3-D compressible that even some irregularities in the shapes of the holes will not
Navier–Stokes simulations with 1st-order slip boundary condi- affect the results in a distinguishable manner. For our experi-
tions [51] and (b) the theoretical prediction for isothermal gas
flow in the channel without slip, with 1st- and 2nd-order slip
ments, Eq. (19) leads to a difference in the mass flow rate of about
boundary conditions (see Eq. (18)). 0:5%. The discrepancy between the experiments and the simula-
tions lies in the differences in the velocity accommodation coeffi-
cient (simulation rv ¼ 0:3; experiment rv ¼ 0:89).
To determine the first-order slip coefficient, the following equa-
2nd-order slip solution. The equation has been given in various tion was used
forms in Refs. [18,50,54] and the one used in our comparison is 2
rv ¼ (20)
Qh ðPin þ Pout ÞðPin  Pout Þ A1 =8 þ 1
¼ D2
Sh 32RLlout Tout

 (18) The corresponding parameters were collected in Table 4. Addi-
Pin Pin þ Pout tionally, some reported values of the tangential momentum
1 þ 8AKn þ 16BKn2 Ln
Pout Pin  Pout accommodation coefficient (TMAC) and corresponding flow con-
ditions from other authors are presented in Table 5. Other methods
where A; B ¼ 0 for the no-slip pipe flow, A1 ¼ 1:7 and B1 ¼ 0 for and the rv value dependencies are extensively discussed in the
the 1st-order slip solution, and A2 ¼ 1:7 and B2 ¼ 1:08 for the review in Ref. [55]. It can be noted that the values are strongly
2nd order slip solution. The obtained fitting constants are taken influenced by inlet Knudsen number. This can be justified by sev-
from the experiments in a microtube flow [45,50] for larger Knud- eral conditions. The first is that the inlet and the outlet effects
sen numbers than ours. It can be seen that our results fit well to were not taken into account in the theory used for slip estimation
the 1st-order slip assumption. However, the difference between (only channel flow). For the slower flows corresponding to the
the 1st- and the 2nd-order slip for that Knudsen number range and higher inlet Knudsen number Knin , the gas interacts with the solid
the coefficient fit taken from literature is undistinguishable. The walls more, not only in the holes, but also at the inlet and the out-
experimental coefficient for the slip is rv ¼ 0:89. let. Hence, it leads to the smaller tangential momentum accommo-
The second comparison for the slip effects has been performed dation coefficient. Similar results were observed in experiments
with numerical simulations [51]. It is plotted in Fig. 12. [56] where the membrane has much smaller holes and a much

Journal of Fluids Engineering DECEMBER 2013, Vol. 135 / 121104-7

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 04/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Table 4 Experimental values of rv estimated for laminar flow models for a large scale [27], viscous effects in slow flows [5],
and experiments in a microtube [50]. For all of these cases, the
Pin ðkPaÞ Knin A1 Confidence rv measured values agreed very well. The estimation of the first-
order tangential momentum accommodation coefficient has
3:0 0.025 7.6 0.912 1.02 shown a strong connection dependency on the inlet Knudsen num-
4:0 : Number 1 0.019 15.44 0.968 0.68
4:0 : Number 2 0.019 16.76 0.987 0.64
ber. It could be a result of the plate’s perforation and inlet related
5:0 0.016 21.60 0.965 0.54 effects. To confirm that, a more extensive study with various plate
perforation values would need to be performed.

Acknowledgment
Table 5 Examples of experimental TMAC values and flow
characteristics This research was supported by the Polish Ministry of Science
and Education under the project entitled “Efficient Turbomachi-
Reference [50] [58] [56] nery Blade Cooling With Various Outlet Coolant Configurations,”
(Grant No. N501 367134).
Gas type N2 CO2 , N2 , Ar O2 , Ar
Pin ðkPaÞ 1.2 130–430 Not given
DPðkPaÞ 1–1.4 30–330 14–149.5 Nomenclature
Porosity Not applicable Not applicable 4 105 a¼ speed of sound
Diameter ðlmÞ 25.2 52 1.3 0.01 A; B; C ¼ coefficients
Kn 0.003–0.289 0.03–0.44 7–100 cp ¼ heat capacity at constant pressure
rv 1:02 0:75  0:85 0:25  0:3
cv ¼ heat capacity at constant volume
D¼ hole diameter
E¼ energy
larger Knudsen number and the observed TMAC was also very F¼ perforation
small. The second reason could be related to the fact that the G¼ source term in momentum equation
Knudsen relation is not a linear function. The second order term k¼ thermal conductivity
strongly depends on the pressure drop (see Eq. 18). The experi- K¼ nondimensional value of pressure
ments presented here were performed for various pressure drops. Kn ¼ Knudsen number based on hole’s diameter
Therefore, it is not possible to fit the data to the estimated second L¼ hole length; perforated plate thickness
order phenomena. Additionally, as was observed in Ref. [57], the Ma ¼ Mach number
temperature or Knudsen relation for the tangential momentum Nh ¼ total number of holes
coefficient can be expected. In our experiments the contribution of P¼ pressure
each of these effects to the estimated value of the rv is unknown Pr ¼ Prandtl number
and additional experiments need to be performed in the future; for Q¼ mass flow rate
example, with various porosities of the plate and hole diameters. R¼ gas constant
In conclusion, it is clearly seen that for the presented compari- Re ¼ Reynolds number based on the hole’s diameter
sons the velocity slip estimation gives a satisfying result. Despite S¼ perforated plate surface
the large scale of the experimental setup and relatively low Knud- T¼ temperature
sen number, the evidence of the slip effects is clearly proven. U¼ velocity
The points on the plot in Fig. 12 were determined based on ex- x¼ wall parallel direction
perimental data; namely, on the pressure and temperature meas- y¼ wall normal direction
urements. The values of these points were calculated using Eqs. Z¼ distance between the holes
(18) and (19). Based on these equations and using the total deriva- b¼ opening factor of the filter (area of the holes over the
tive method, the error in Fig. 12 can be estimated. This error total area)
amounts toapproximately 0:24% of the given value (for details, d¼ error
see the Appendix). It was assumed that the flow in the microholes j¼ specific heat constant
hold at adiabatic conditions. For the subcritical flows, which k¼ mean free path of the gas
mostly occur in the experiment, the deviation from the adiabatic l¼ dynamic viscosity
condition at such measurement conditions can be neglected. n¼ permeability
q¼ density
rv ¼ tangential momentum accommodation coefficient
4 Conclusion (TMAC)
This paper focuses on the experimental investigation of the per- u¼ inertial resistance factor
forated plate aerodynamics for wide ranges of Reynolds and x¼ viscosity index
Knudsen numbers. It was shown that two Reynolds dependent
regimes existed: the laminar one and the transition to turbulence Indices
occurring for Re ¼ 60. The observed Reynolds number value for ad ¼ adiabatic
the transition to turbulence is very low. It confirms observations avg ¼ average
performed by other researchers, that in microchannel transition to eff ¼ effective
turbulence is present for Reynolds numbers much smaller than in h¼ hole
the macroscale. Our experiment is the first one which shows the i; j ¼ index
turbulence transition on the perforated plate with microscale influ- in ¼ inlet parameters
ence. The perforated plate had holes of a micrometer size L¼ laminar
(110 lm). Hence, we also investigated the velocity slip effects on na ¼ nonadiabatic
the mass flow rate. It was shown that there was clear evidence of out ¼ outlet parameters
the slip on the order of 0:5% influencing the mass flow rate. This ref ¼ reference value
is not a significant alternation. However, it can be utilized in T¼ transition to turbulence
many applications where even very small changes can bring large w¼ wall parameters
gains, such as boundary layer control for turbomachinery. The ex- 0¼ stagnation parameters
perimental results were compared for compressible experimental ?¼ critical parameters

121104-8 / Vol. 135, DECEMBER 2013 Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 04/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Appendix: Errors Estimation Table 6 Device measurement errors and the CFD estimated
nonadiabatic error for the temperature
The mass flow rate is normalized in such a way that the plate
resistance is irrelevant as the adequate slip and no-slip solution is dT (K) dTna (K) dP(kPa) dL (mm) dD (mm)
taken to account. Based on Eq. (19), analogically to Eq. (A2), the
error of the curve determination could be expressed in the form 0.1 0.15 103 0.065 0.002 0.01

 
1 Qexp Qexp !
Qns  Q1s Qns dQexp Qexp 
Qns Qs
dQ ¼ þ  þ dQns If we are modeling a perforated plate it is recommended to elimi-
Qs Qns Qs Qs nate the permeability term and use the inertial loss term alone,
0   1 (A1) yielding the following simplified form of the porous media
Qexp Qexp
Q Q
exp ns Q  Q Qns equation
þ@ AdQs
ns s
þ
Q3s Q2s

X
3
1
rP ¼  uij qUj jUj (A5)
j¼1
2
where dQexp ; dQs ; and dQns are the mass flow rate error in the
measurements, with slip and no-slip assumptions, respectively.
Based on Eq. (18), the error in the curve determination could be To predict the heat transfer between the microholes in air and the
expressed in the following form part of the membrane surface with the ambient condition we need
to solve the standard energy transport equation in porous media
0 1 regions with modifications to the conduction flux and the transient
~ 2~2 terms only. The effective thermal conductivity in the porous me-
B D Px D P C ðdTout þ dTna Þ dium keff ¼ Skf þ ð1  SÞks is computed as the volume average of
dQ ¼ @  x þ  x A
Tout 2 Tout 2
RL Tref Tout RL Tref Tout 32 the air conductivity kf and the microhole membrane conductivity
ks . Some calculations for critical and subcritical conditions were
1 ~
D2 Px carried out. During 10s of simulation time only insignificant heat
þ   ðdTref þ dTad Þ
32 RL Tout x T T transfer values were observed (6–10 W) for an ambient surface. A
Tref out ref value of dTna ¼ 0:15 103 K was considered. All of the
1 D2 ðDPin ðPin  Pout Þ þ DPout ðPin  Pout ÞÞ values of the determination curve of the error are collected in
þ  x dPin Table 6.
32 RL Tout T Tref out
2
1 D ðDPin ðPin  Pout Þ  DPout ðPin  Pout ÞÞ Table 7 Error estimated according to Eq. (18), (A1), and (A2)
þ  x dPout
32 RL Tout T Tref out
Pin ðkPaÞ Pout ðkPaÞ dQ=Qð%Þ dQ=Qð%Þ
2
1 D P 1 DP
   dL þ   dD Pin ¼ 5 kPa
32 RL Tout x T 16 RL Tout x T 5.071 2.456 0.3819 0.2361
Tref out Tref out
5.037 2.855 0.3816 0.2368
(A2) 5.021 2.905 0.3816 0.2368
5.064 3.006 0.3815 0.2370
5.017 3.106 0.3814 0.2371
where 5.044 3.407 0.3813 0.2374
5.067 3.659 0.3811 0.2377
5.067 4.212 0.3809 0.2382
P~ ¼ Pin ðPin  Pout Þ þ Pout ðPin  Pout Þ (A3) 5.016 4.664 0.3807 0.2386
Pin ¼ 4 kPa :Number 1
4.024 2.163 0.3818 0.2363
and dTref ; dTout is the temperature measurement error, dPin ; dPout
4.023 2.262 0.3817 0.2367
is the pressure measurement error, dD is the hole’s diameter 4.023 2.313 0.3816 0.2368
error, dL is the thickness of the plate error. It is a classical form 4.024 2.414 0.3815 0.2369
of a systemic error with a nonadiabatic correction dTna . All of 4.073 2.564 0.3814 0.2372
the values of the errors are collected in Table 7. It was assumed 4.023 2.765 0.3813 0.2375
that the flow in the microholes held the adiabatic conditions. The 3.995 2.865 0.3811 0.2377
only exception from the adiabatic assumption was the existence 4.021 3.518 0.3809 0.2382
of a very small contact between the membrane and the ambient 4.070 3.818 0.3807 0.2386
air and it was due to the measurement requirements. This contact Pin ¼ 4 kPa :Number 2
4.043 2.031 0.3818 0.2365
(surface to surface) of all microholes was smaller by about two
4.040 2.229 0.3816 0.2367
orders (0.05). The measurement time of one point of the charac- 4.096 2.336 0.3816 0.2368
teristics was about 10 s. The maximum temperature change due 4.045 2.385 0.3815 0.2370
to the recovery factor at critical conditions was 5:4 K. For sub- 4.046 2.537 0.3814 0.2371
critical flows the deviation from the adiabatic condition for the 4.040 2.732 0.3813 0.2375
heat exchange could be neglected. However, in order to ensure 4.039 2.932 0.3812 0.2377
that our claim was justified we estimated a nonadiabatic correc- 4.044 3.337 0.3808 0.2384
tion in the experiment by a computational fluid dynamics (CFD) 4.037 3.735 0.3807 0.2386
study. To simplify the CFD calculation we assumed that the per- Pin ¼ 3 kPa
3.060 2.004 0.3813 0.2373
forated plate was modeled by porous media by adding a momen-
3.013 2.510 0.3809 0.2382
tum source term to the standard fluid flow equations expressed in 3.098 2.870 0.3807 0.2386
the form Pin ¼ 2:5 kPa
2.410 1.958 0.3810 0.2380
2.462 2.361 0.3807 0.2386
Gi ¼ ðnUi =l þ uqjUjUi =2Þ (A4)

Journal of Fluids Engineering DECEMBER 2013, Vol. 135 / 121104-9

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 04/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


References [31] Colin, S., 2012, “Gas Microflows in the Slip Flow Regime: A Critical Review
on Convective Heat Transfer,” ASME J. Heat Transfer, 134, p. 020908.
[1] Budoff, M., and Zorumski, W., 1971, “Flow Resistance of Perforated Plates in [32] Tamayol, A., and Hooman, K., 2011, “Slip-Flow in Microchannels of Non-
Tangential Flow,” NASA Report No. TM X-2361. Circular Cross Sections,” ASME J Fluids Eng., 133, p. 091202.
[2] Liu, R., and Ting, D., 2007, “Turbulent Flow Downstream of a Perforated Plate: [33] Yang, J., Ye, J., Zheng, J., Wong, I., Lam, C., Xu, P., Chen, R., and Zhu, Z.,
Sharp-Edged Orifice Versus Finite-Thickness Holes,” ASME J. Fluids Eng., 2010, “Using Direct Simulation Monte Carlo With Improved Boundary Condi-
129, pp. 1164–1171. tions for Heat and Mass Transfer in Microchannels,” ASME J. Heat Transfer,
[3] Mishra, C., and Peles, Y., 2005, “Incompressible and Compressible Flows 132, p. 041008.
Through Rectangular Microorifices Entrenched in Silicon Microchannels,” J. [34] Wang, M., and Li, Z., 2004, “Simulations for Gas Flows in Microgeometries
Microelectromech. Syst., 14(5), pp. 1000–1012. Using the Direct Simulation Monte Carlo Method,” Int. J. Heat Fluid Flow,
[4] van Rijn, C., van der Wekken, M., Hijdam, W., and Elwenpoek, M., 1994, 25(6), pp. 975–985.
“Deflection and Maximum Load of Microfiltration Membrane Sieves Made [35] Garcia, R., and Siewert, C., 2010, “Viscous-Slip, Thermal-Slip, and
With Silicon Micromachining,” J. Microelectromech. Syst., 6, pp. 48–54. Temperature-Jump Coefficients Based on the Linearized Boltzmann Equation
[5] Yang, J., Ho, C., Yang, X., and Tai, Y., 2001, “Micromachined Particle Filter (and Five Kinetic Models) With the Cercignani–Lampis Boundary Condition”.
With Low Power Dissipation,” ASME J. Fluids Eng., 123, pp. 899–908. Eur. J. Mech. B/Fluids, 29, pp. 181–191.
[6] Zierep, J., Bohning, R., and Doerffer, P., 2003, “Experimental and Analytical [36] Jain, V., and Lin, C., 2006, “Numerical Modeling of Three-Dimensional Com-
Analysis of Perforated Plate Aerodynamics,” J. Therm. Sci., 12(3), pp. 193–197. pressible Gas Flow in Microchannels,” J. Micromech. Microeng., 16, pp.
[7] Zhang, T., Jia, L., Li, C., Yang, L., and Jaluria, Y., 2011, “Experimental Study 292–302.
on Single-Phase Gas Flow in Microtubes,” ASME J. Heat Transfer, 133, p. [37] Szalmas, L., and Valougeorgis, D., 2002, “Rarefied Gas Flow of Binary Mix-
111703. tures Through Long Channels With Triangular and Trapezoidal Cross
[8] Lahjomri, J., and Obarra, A., 2013, “Hydrodynamic and Thermal Characteris- Sections,” Microfluid. Nanofluid., 9(2), pp. 471–487.
tics of Laminar Slip Flow Over a Horizontal Isothermal Flat Plate,” ASME J. [38] Albertoni, S., Cercignani, C., and Gotusso, L., 1963, “Numerical Evaluation of
Heat Transfer, 135, p. 021704. the Slip Coeffcient,” Phys. Fluids, 6, pp. 993–996.
[9] Barashkin, S., Porodnov, B., and Syromyatnikov, S., 1989.,“Density Distribu- [39] Shen, S., Chen, G., Crone, R., and Anaya-Dufresne, M., 2007, “A Kinetic-
tion in Rarefied Gas Flow Behind a Perforated Plate,” Fluid Dyn., 24(3), pp. Theory Based First Order Slip Boundary Condition for Gas Flow,” Phys Fluids,
482–484. 19, p. 086101.
[10] Edwards, D., Shapiro, M., Bar-Yoseph, P., and Shapira, M., 1990, “The Influ- [40] Arkilic, E., Schmidt, M., and Breuer, K., 1997, “Gaseous Slip Flow in Long
ence of Reynolds Number Upon the Apparent Permeability of Spatially Peri- Microchannels,” J. Microelectromech. Syst., 6(2), pp. 167–178.
odic Arrays of Cylinders,” Phys. Fluids A, 2(1), pp. 45–55. [41] Hsieh, S., Tsai, H., Lin, C., Huang, C., and Chien, C., 2004, “Gas Flow in a
[11] Pozrikidis, C., 2010, “Slip Velocity Over a Perforated or Patchy Surface,” J. Long Microchannel,” Int. J. Heat Mass Transfer, 47(17), pp. 3877–3887.
Fluid Mech., 643, pp. 471–477. [42] Yang, Z., and Garimella, S., 2009, “Rarefied Gas Flow in Microtubes at Differ-
[12] Tio, K., and Sadhal, S., 1994, “Boundary Conditions for Stokes Flows Near a ent Inlet-Outlet Pressure Ratios,” Phys Fluids, 21, p. 052005.
Porous Membrane,” Appl. Sci. Res., 52, pp. 1–20. [43] van Male, P., de Croon, M., Tiggelaar, R., van den Berg, A., and Schouten, J.
[13] Duan, Z., 2011, “Incompressible Criterion and Pressure Drop for Gaseous Slip C., 2004, “Heat and Mass Transfer in a Square Microchannel With Asymmetric
Flow in Circular and Noncircular Microchannels,” ASME J. Fluids Eng., 133, Heating,” Int. J. Heat Mass Transfer, 47, pp. 87–99.
p. 074501. [44] Yao, Z., He, F., Ding, Y., Shen, M., and Wang, X., 2004, “Low-Speed Gas
[14] Maa, D., 1975, “Potentials of Micro Perforated Absorbers,” J. Acoust. Soc. Flow Subchoking Phenomenon in a Long-Constant-Area Microchannel,” AIAA
Am., 104(5), pp. 2866–2868. J., 42(8), pp. 1517–1521.
[15] Babak, V., Babak, T., Kholpanov, L., and Malyusov, V., 1986, “A Procedure [45] Ewart, T., Perrier, P., Graur, I., and Meolans, J., 2007, “Mass Flow Rate Meas-
for Calculating Matrix Heat Exchangers Formed From Perforated Plates,” J. urements in a Microchannel, From Hydrodynamic to Near Free Molecular
Eng. Phys., 50(3), pp. 330–335. Regimes,” J. Fluid Mech., 584, pp. 337–356.
[16] Bunker, R., 2007, “Gas Turbine Heat Transfer: Ten Remaining Hot Gas Path [46] Pitakarnnop, J., Varoutis, S., Valougeorgis, D., Geoffroy, S., Baldas, L., and
Challenges,” ASME J. Turbomach., 129(2), pp. 193–201. Colin, S., 2010, “A Novel Experimental Setup for Gas Microflows,” Microfluid.
[17] Gau, C., and Huang, W., 1990, “Effect of Weak Swirling Flow on Film Cooling Nanofluid., 8(1), pp. 57–72.
Performance,” ASME J. Turbomach., 112(4), pp. 786–791. [47] Rebrov, A., Morozov, A., Plotnikov, M., Yu, Z., Timoshenko, N., and Maltsev,
[18] Howarth, L., 1953, Modern Developments in Fluid Dynamics: High Speed V., 2003, “Determination of Accommodation Coefficients of Translational and
Flow, Vol. 2, Clarendon, Oxford, UK. Internal Energy Using a Thin Wire in a Free-Molecular Flow,” Rev. Sci. Ins-
[19] Bird, G., 1994, Molecular Gas Dynamics and the Direct Simulation of Gas trum., 74(2), pp. 1103–1106.
Flows, Oxford University Press, New York. [48] Sone, Y., 2000, “Flows Induced by Temperature Fields in a Rarefied Gas and
[20] White, F., 1998, Fluid Mechanics, McGraw-Hill, New York. Their Ghost Effect on the Behavior of a Gas in the Continuum Limit,” Ann.
[21] Beskok, A., and Karniadakis, G., 1999, “A Model for Flows in Channels, Pipes Rev. Fluid Mech., 32, pp. 779–811.
and Ducts at Micro and Nano Scale,” Microscale Thermophys. Eng., 3, pp. 43–77. [49] Venerus, D., and Bugajsky, D., 2010, “Compressible Laminar Flow in a
[22] Hagen, G., 1839, “Uber den bewegung des wassers in engen cylindrischen Channel,” Phys Fluids, 22, p. 046101.
rohren,” Ann. Phys. Chem., 46, pp. 423–442. [50] Ewart, T., Perrier, P., Graur, I., and Meolans, J., 2006, “Mass Flow Rate Meas-
[23] Eckhardt, B., 2008, “Turbulence Transition in Pipe Flow: Some Open Ques- urements in Gas Micro Flows,” Exp. Fluids, 41, pp. 487–498.
tions,” Nonlinearity, 21, pp. 1–11. [51] Lewandowski, T., Jebauer, S., Czerwinska, J., and Doerffer, P., 2009, “Entrance
[24] Kohl, M., Abdel-Khalik, S., Jeter, S., and Sadowski, D., 2005, “A Microfluidic Effects in Microchannel Gas Flow,” J. Therm. Sci., 18(4), pp. 345–352.
Experimental Platform With Internal Pressure Measurements,” Sens. Actuators, [52] Maxwell, J., 1879, “On Stresses in Rarified Gases Arising From Inequalities of
A, 118(2), pp. 212–221. Temperature,” Philos. Trans. R. Soc. London, 170, pp. 231–256.
[25] Gan, G., and Riffat, S., 1997, “Pressure Loss Characteristics of Orifice and Per- [53] Smoluchowski, M., 1898, “Uber den temperatursprung bei warmeleitung in
forated Plates,” Exp. Therm. Fluid Sci., 14(2), pp. 160–165. gasen,” Akad. Wiss. Wien, 107, pp. 304–329.
[26] Cuhadaroglu, B., Akansu, Y., and Turhal, A., 2007, “An Experimental Study on [54] Cai, C., Sun, Q., and Boyd, I., 2007, “Gas Flows in Microchannels and Micro-
the Effects of Uniform Injection Through One Perforated Surface of a Square tubes,” J. Fluid Mech., 89, pp. 305–314.
Cylinder on Some Aerodynamic Parameters,” Exp. Therm. Fluid Sci., 31, pp. [55] Agrawal, A., and Prabhu, S. V., 2008, “Survey on Measurement of Tangential
909–915. Momentum Accommodation Coefficient,” J. Vac. Sci. Technol. A, 26(4), pp.
[27] Doerffer, P., and Bohning, R., 2000, “Modelling of Perforated Plate Aerody- 634–645.
namics Performance,” Aerosp. Sci. Technol., 4, pp. 525–534. [56] Roy, S., Cooper, S. M., Meyyappan, M., and Cruden, B. A., 2005, “Single
[28] Dagan, Z., Weinbaum, S., and Pfeffer, R., 1982, “An Infinite Solution for the Component Gas Transport Through 10 nm Pores: Experimental Data and
Creeping Motion Through an Orifice of Finite Length,” J. Fluid Mech., 115, pp. Hydrodynamic Prediction,” J. Membr. Sci., 253, pp. 209–215.
505–523. [57] Cao, B.-Y., Chen, M., and Guo, Z.-Y., 2005, “Temperature Dependence of the
[29] Ahmed, I., and Beskok, A., 2002, “Rarefaction, Compressibility, and Viscous Tangential Momentum Accommodation Coefficient for Gases,” Appl. Phys.
Heating in Gas Micro Filters,” J. Thermophys. Heat Transfer, 16(2), pp. 161–170. Lett., 86, pp. 091905-3.
[30] Roy, S., Raju, R., Cruden, H. C. B., and Meyyappan, M., 2003, “Modeling Gas [58] Arkilic, E., Breuer, K., and Schmidt, M., 2001, “Mass Flow and Tangential Mo-
Flow Through Microchannels and Nanopores,” J. Appl. Phys., 93(8), pp. mentum Accommodation in Silicon Micromachined Channels,” J. Fluid Mech.,
4870–4879. 437(5), pp. 29–43.

121104-10 / Vol. 135, DECEMBER 2013 Transactions of the ASME

DownloadedViewFrom:
publicationhttp://fluidsengineering.asmedigitalcollection.asme.org/
stats on 04/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like