You are on page 1of 13

International Journal of Heat and Fluid Flow 65 (2017) 60–72

Contents lists available at ScienceDirect

International Journal of Heat and Fluid Flow


journal homepage: www.elsevier.com/locate/ijhff

Pressure drop measurements for woven metal mesh screens used in


electrical safety switchgears
W. Bussière a,∗, D. Rochette a, S. Clain b, P. André a, J.B. Renard c
a
Université Clermont Auvergne, CNRS/IN2P3, LPC, Clermont-Ferrand F-63000, France
b
Centro de Matematica (CMAT), Campus de Gualtar, Universidade do Minho, Braga 4710 – 057, Portugal
c
LPC2E CNRS, 3A Avenue de la Recherche Scientifique, Orléans Cedex 2 45071, France

a r t i c l e i n f o a b s t r a c t

Article history: The paper deals with an experimental investigation of the pressure drops and friction factors induced by
Received 4 May 2016 sets of metal woven screens in the case of an incompressible fluid flow, namely water flow. These woven
Revised 25 January 2017
screens are of metallic type and are used in electrical safety devices, especially at the end of exhaust duct
Accepted 18 February 2017
of low voltage circuit breakers. In a first step, the whole of the set-up is used to check some approxima-
tions dedicated to pressure drops due to sphere beds around 275 μm and 375 μm as diameters. Results
Keywords: are discussed in terms of Darcyan and non-Darcyan permeabilities compared with published data and
Pressure drop are analyzed in terms of the Blake-type friction factor. In the second step, the pressure drops are mea-
Ergun law sured for stacks composed of eight different woven screens (of plain dutch type) formed with millimetric
Drag coefficient
wires. Various formulations of the pressure drops and friction factors published elsewhere are tested with
Friction factor
a special care dedicated to the choice of the geometry for the flow pattern (hydraulic diameter, spher-
Metal woven screens
Porous media ical diameter, cylindrical diameter) and to the consideration of laminar and turbulent contribution. We
then give the formulation that characterizes the fluid flowing through stacks of woven screens used in
electrical safety applications.
© 2017 Elsevier Inc. All rights reserved.

1. Introduction ing solutions and their associated performance especially for the
Medium Voltage (MV) electrical safety (Morel and Rival, 1990). We
Almost all the studies dealing with woven metal screens essen- recall that among the whole of the recommendations in relation
tially refer to topics such as high efficiency heat exchangers, en- with the current study, the filter technologies have to ensure an
ergy storage units, regenerators, electronics coolers, catalytic reac- efficient working of the electrical switchgear and to help the clear-
tors, filtering industry, gas bearing supply systems, gas-liquid two ing of the very high energy level linked to the internal fault arc
phase flow applications ((Wu et al., 2005; Belforte et al., 2009; clearance, with a specific capacity to be proof against very high
Kolodziej et al., 2009; Mahjoob and Vafai, 2008) and references pressure and shock waves on one hand, and the brutal flow out
herein). But up to our knowledge, there is no published work with of hot gases at the exhaust (up to around 10 kK in the center of
experimental measurements of the drag coefficients dealing with the MV cell) on the other hand. Filters are thus designed to limit
stacks of woven metal screens used for electrical safety purposes the temperature, the velocity and the pressure of the gases at the
which is the aim of the present work. exhaust.
Electrical safety practices are of crucial importance in elec- As a consequence innovative solutions have been carried out in
trical devices industry and have to follow the rules im- order to ensure the mandatory level of electrical safety. These im-
posed by standards from IEC (www.schneider-electric.com 2007; provements and the general rules are depicted in various patents,
IEC 2003) and specific safety guidelines for example summarized among them we refer to Morel and Rival (1990), Rival et al.
in (Schneider Electric–Square D 2003). In a recent past, filter tech- (1999) and Faber et al. (2003).
nologies have been widely studied in order to improve exist- Whatever the voltage range considered, MV or LV (Low Volt-
age), the electrical safety switchgears are designed to support the
clearance of the electric fault. This is done by means of the ig-
∗ nition of an electric arc (typically between two metallic contacts
Corresponding author at: Laboratoire de Physique de Clermont, LPC CNRS
UMR6533, Campus Universitaire des Cézeaux, 4 Avenue Blaise Pascal, TSA 60026, or electrodes or splitter plates) and hence of the consecutive arc
CS 60026, Aubière Cedex 63178, France. plasma which has to be considered as a time dependent resistive
E-mail address: william.bussiere@uca.fr (W. Bussière).

http://dx.doi.org/10.1016/j.ijheatfluidflow.2017.02.008
0142-727X/© 2017 Elsevier Inc. All rights reserved.
W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72 61

So it is not realistic to recreate and use such media to obtain re-


Nomenclature liable measurements for Darcyan and non-Darcyan permeabilities
(short duration, high gradients difficult to reproduce for systematic
a specific surface area (m −1 ) measurements, destructive hot power source). In a previous work
A, B coefficients in Ergun equation (-) (Bussière et al., 2006) we have done some measurements with air
A1 , A2 coefficients in Blake-type equation (-) as the test fluid.
C1 , C2 coefficients in Ergun-Bird equation (-) Among the possible risks that have to be taken into account,
CW wall correction factor in inertial term (-) the protection against arc flash is of high interest both for the
d1 warp wire diameter (m) safety of the maintenance staff and for the vicinity of the elec-
d2 weft wire diameter (m) trical apparatus. In the case of MV cells, many studies have
dp particle diameter (m) shown the significant role played by porous filters (Besnard, 2007;
D diameter of a single screen (m) Rochette et al., 2007; Rochette et al., 2008; Rochette et al., 2010;
DC cylindrical diameter model (m) Rochette et al., 2011) which are mainly of mineral kind and com-
Dh hydraulic diameter (m) posed of a deformable granular medium fixed by metallic screens
Dp , Deq particle equivalent diameter (m) (whose mean granulometry and porosity are carefully checked). In
DS spherical diameter model (m) these latter works, we have studied the specific influence of the
f, F friction factor (-) porous filters on fundamental physical properties playing an im-
k1 viscous constant of the Darcy–Forchheimer portant role during the MV switchgears work namely the pressure,
equation (m2 ) or Darcyan permeability term or temperature and velocity distributions in the MV cell and at the
permeability coefficient exhaust of the MV cell.
k2 inertial constant of the Darcy–Forchheimer More recently the filter technologies have been adapted to the
equation (m) or non-Darcyan permeability LV switchgears to take advantage of the improvements obtained
term or inertial coefficient in the MV range (Faber et al., 2003). The filter technology is dif-
L length of the porous medium (m) ferent in so far as the dispersed granular medium is not used
Le effective flow length (m) any more. The filter consists in one or more woven metal mesh
M wall correction factor (-) screens defined by different wire diameters, mesh sizes and porosi-
P pressure drop (Pa) ties which are placed on an optimal location at the gas exhaust of
Re, NRe Reynolds number (-) the switchgear. Obviously, as for the MV range, filters used in LV
T water temperature (°C) switchgears influence the main physical properties which charac-
v, w0 , U, Um measured water velocity or superficial water terize the gas at the exhaust, namely the pressure, the temperature
velocity (m s−1 ) and the velocity.
we effective or intersticial velocity (m s−1 ), we = The aim of the current work is to assess the pressure drops
w0 /ε and/or the friction factors of such filters in order to depict the
influence of filters formed by different stacked metal screens and
Greek symbols
to provide numerical values for modeling tools. In Section 2 the
ɛ porosity (-)
features of spherical particles and of the woven metal screens are
θ angle of the fluid flow (Kolodziej and Lojew-
given, and the experimental set-up is described. Section 3 is ded-
ska, 2009b)
icated to a brief review about the formulations of the Ergun ap-
μ water dynamic viscosity (Pa s)
proximation for spherical particles in order to validate the whole
ρ water density (kg m−3 )
of the experimental set-up by comparison with other published
τ tortuosity (-)
data. In Section 4 we report the pressure drop measurements ob-
ψ drag coefficient in the flow around approach
tained for woven metal screens. Experimental results are analyzed
(Innocentini et al., 20 0 0)
by means of modified Ergun law in order to assess specific param-
Subscripts eters such as the tortuosity, the angle between the flow inside the
D Darcy porous material and the flow direction induced by the porous ma-
F Fanning terial thickness. We give the pressure drop values and/or friction
k kinetic factors for the stacks of woven metal screens, these latters being
Wu for the three coefficients α , β and γ defined in the analyzed by means of different equations, one of which is specifi-
approximation of Wu et al. (2005) cally established for the studied screens.

2. Experimental set-up and procedure


component in the whole electric circuit. Due to the proximity of
the arc chamber walls (typically in polymer materials) the char- 2.1. Sphere beds features
acteristics and the properties (composition, electrical and thermal
conductivities, radiation, viscosity) of the arc plasma continuously Two sets of silica spheres provided by Sartorius and quoted
evolve due to the polymer wall ablation. They also depend on the B. Braun Glasperlen 0.25–0.30 mm and 0.25–0.50 mm – respectively
very high gradients in temperature and pressure of the arc cham- called dp = 275 μm and dp = 375 μm in the paper – are used to
ber. During the circuit breaker working some liquid (fused metal check and validate the experimental set-up on the one side, and
droplets) and gaseous (hot vapors resulting from ablated walls) to measure the pressure drop formulations dedicated to spheres
materials are produced. Thus filters have to be designed to avoid on the other side. The features of the spheres and the porosity of
or at least to significantly limit the flow of these fluids at the ex- the sphere beds (mean values and standard deviations) are given
haust duct of the circuit breaker as it is illustrated in Rochette et al. in Table 1(a). More details about the measurement and the defini-
(2015). One may easily understand that an arc plasma is a transient tion of the aspect ratio and compactness are given in Bussière et al.
phenomenon whose properties variations are very important: the (2006). The relative discrepancy between the porosities obtained
arc plasma chamber can have temperatures higher than 10 kK after for the two beds is ≈ 3% and has no specific meaning in regards to
few milliseconds, pressures up to some MPa (Rochette et al., 2015). experimental uncertainties. Experimental porosity values are 0.416
62 W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72

Table 1
Parameters of (a) the silica spherical particles and of (b) the woven metal mesh screens († , from (Woven Wire Cloth 2004, http://www.haverboecker.com March
2016)).

Particle diameter, dp (μm) Granulometric range (μm) Porosity, ɛ Shape factor (Bussière et al., 2006)
Aspect ratio Compactness

275 [250..300] 0.416 1.12 ± 0.04 0.93 ± 0.02


375 [250..500] 0.403 1.06 ± 0.04 1.01 ± 0.01

N° Mesh† Warp wire Weft wire Single screen Porosity† , ɛ Specific surface Diameter of single Hydraulic diameter,
diameter† , d1 (mm) diameter† , d2 (mm) thickness† (mm) area† , a (m−1 ) screen, D (mm) Dh from Eq. (1) (mm)

1 3.5 × 20 1.6 1.25 3.15 0.68 1272 40 2.14


2 5 × 32 1.25 0.9 2.59 0.67 1900 40 1.41
3 8 × 48 0.9 0.56 1.54 0.66 3464 40 0.762
4 11 × 56 0.71 0.47 1.44 0.64 4106 40 0.623
5 12 × 64 0.63 0.42 1.28 0.65 4222 40 0.616
6 21 × 85 0.45 0.3 0.92 0.61 7464 40 0.327
7 24 × 110 0.355 0.25 0.76 0.63 7760 40 0.325
8 30 × 150 0.224 0.18 0.51 0.65 9167 40 0.284

and 0.403 respectively for 275 μm and 375 μm. These values are
close to 0.395 – relative deviations are respectively ∼ 5% and ∼
2% – which show that the arrangement obtained after packing
is rather close to an orthorhombic one. The test cell diameter to
particle diameter ratios D/dp are respectively ∼145 and ∼107. The
granulometric analysis given in Fig. 1(a) and performed with a spe-
cific method described for example in Renard et al. (2002) shows
a larger granulometric interval for each set than the one quoted by
the supplier.

2.2. Woven metal screens features

A typical sketch of the woven metal screens used in the study


is shown in Fig. 1(b). The main features of these screens are given
in Table 1(b) for the eight different screens tested. The weft wires
are slightly thinner than the warp wires, the wire diameters are
of the millimetric domain ranging from 0.18 mm to 1.6 mm. Wo-
ven thickness varies from 0.51 mm to 3.15 mm which is linked to
the wire diameters used. The porosity ɛ is rather similar for each
metal screen and varies in the interval 0.61 to 0.68. Due to the
small variation, the porosity may not be considered as a relevant
parameter to distinguish each weave. On the contrary the specific
surface area a is a more relevant discriminating characteristic. For
woven screens numbered from 1 to 8, the specific area increases
from 1272 m−1 to 9167 m−1 . As a result the hydraulic diameter Dh
decreases respectively from 2.14 mm down to 0.284 mm, the defi-
nition being (Kolodziej et al., 2009):

Dh = (1)
a
Other formulations can be used respectively in the case of
spherical model (subscript S) (Wu et al., 2005) and in the case
of infinite cylinder model (subscript C) (Kolodziej and Lojew-
ska, 2009b):
1−ε
DS = 6 (2)
a
and
1−ε
DC = 4 (3)
a

2.3. Experimental set-up

The experimental set-up is displayed in Fig. 2(a) with the main Fig. 1. Porous media studied. (a) Granulometric distributions of the quoted 0.25–
components. The water flow is created by means of a hydraulic 0.30 mm (dp = 275 μm) and 0.25–0.50 mm (dp = 375 μm) silica spheres measured re-
pump from DAB (model KV40) with an operating temperature spectively with 265 and 384 particles. (b) Schematic diagram of the woven metal
range up to 110 °C, a maximum working pressure equal to 25 bars. mesh screens (known as single plain dutch weave from Woven Wire Cloth (2004),
also called plain dutch as quoted in Armour and Cannon (1968)).
The water flow is measured with a 8700 magnetic flowmeter from
W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72 63

cles already published in the literature. The second step consists in


assessing the permeability of various stacks of woven metal mesh
screens that are currently used for electrical safety purposes.

3. Experimental results for sphere beds

3.1. Flow regimes

Prior to define the permeability formulations one has to won-


der about the type of the flow regime in regard to the Reynolds
number. The experimental Reynolds number is given in Fig. 3(a) in
function of the flow channel diameter D, the particle diameter dp
and is defined by:
ρvD
ReD = (4)
μ
and
ρvd p
Red p = (5)
μ
For the two sphere beds the ReD values are very similar (vary-
ing from ∼103 to 104 ), the discrepancies are insignificant and
linked to experimental uncertainties (density and dynamic viscos-
ity dependence with temperature). On the contrary, the Red p val-
ues are slightly higher for 375 μm (from ∼101 to 102 ) compared to
275 μm (from ∼101 to 5 × 101 ). From Kaviany (1999), the flow can
be considered as inertial-flow regime: this means that one has a
steady nonlinear laminar flow characterized by 1–10 < Red p < 150
where the inertial force tends to dominate over the viscous forces.

3.2. Porous medium permeability formulation

Fig. 2. Schematic diagrams of (a) the experimental set-up and (b) the test cell with The pressure drop measurements with uncertainties are given
pressure transducers (Pi ).
in Fig. 3(a) for the two sphere beds. For a given experimental ve-
locity value, higher pressure drops are observed for 275 μm which
is coherent with the lower corresponding hydraulic diameter.
Rosemount (8705 flowtube equipped with a 8732 transmitter) al- The numerous studies dedicated to permeability assessments
lowing flow measurements from 0.01 m s−1 to 10 m s−1 . The incer- published in the literature can be roughly summarized from the
tainty is quoted at ± 0.5% for [0.3–10] m s−1 and ± 0.0015 m s−1 for type of porous medium point of view: sphere beds (Bussière et al.,
flow speed inferior to 0.3 m s−1 . The resolution is 0.1 L min−1 . The 2006; Macdonald et al., 1979; Eisfeld and Schnitzlein, 2001;
volume of the water tank is 500 L and enables to work in closed Pope et al., 2011; Choi et al., 2008), foams (metallic (Mahjoob and
flow system. The temperature of the water flow is continuously Vafai, 2008) or mineral (Moreira and Coury, 2004) or polymeric
measured by means of a K-type thermocouple. From this measure- (Schmid et al., 1999)), woven metal mesh screens (Wu et al., 2005;
ment the water density and the dynamic viscosity are deduced Belforte et al., 2009; Kolodziej et al., 2009; Kolodziej and Lojew-
from data quoted in Handbook of Chemistry and Physics (2012), ska, 20 09b; Nika, 20 08; Kolodziej and Lojewska, 2009a) are the
Huber et al. (2009), Kadoya et al. (1963) and Kaviany (1999). The main porous media quoted.
measurement is necessary to assess true values of viscosity and Whatever the porous medium, the starting point is the defini-
density on one hand, and to take into account the influence of the tion of the permeability given by the Darcy equation for an incom-
water heating due to the pump working on the other hand. pressible fluid (Rohsenow et al., 1998) (the pressure drop is lin-
The test cell is presented in Fig. 2(b) with the major features. ear with the flow velocity, viscous term) improved by the Forch-
The inner diameter is 40 mm. The test cell is equipped with four heimer equation (addition of the quadratic contribution, inertial
static pressure transducers Kistler K-line calibrated up to 2 MPa term) (Multiphase Flow Handbook 2006):
with output voltage set to 10 V ± 0.5%. The output voltage is mea-
P 1 1
sured with a 3014B Tektronix oscilloscope. Porous media or filters = μv + ρv2 = αμv + βρv2 (6)
L k1 k2
(woven metal mesh screens) can be placed at the entrance and at
the exit of the test cell. A dedicated void of length L is reserved where α is the Darcy coefficient (m−2 ) with α = 1/k1 and β is the
for dispersed porous media such as silica sphere beds. Two metal- Forchheimer coefficient (m−1 ) with β = 1/k2 . Eq. (6) describes the
lic screens of known mesh are used to hold particles when a dis- pressure drop per unit length of fluid during the flow through a
persed porous medium is studied. In the case of dispersed porous porous medium. It includes the viscous contribution – character-
medium, the experimental porosity is assessed from the measure- ized by the parameter k1 , also commonly known as the Darcyan
ment of the mass difference of the test cell before and after filling parameter, to take into account the linear dependency with the
and compacting. flow rate. It also depends on the inertial contribution, described
In an initial stage, the permeability of dispersed porous media by the parameter k2 , also commonly known as the non-Darcyan
is studied using the two quoted silica sphere beds. The objective parameter, to take into account the quadratic dependence with the
is first to validate the whole set-up using well-defined beds. Sec- flow rate, and assumes a constant value for the viscosity and the
ondly we check the permeability formulations for spherical parti- density of the flowing fluid.
64 W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72

Fig. 3. For sphere beds. (a) Pressure drops P/L (Pa/m) versus experimental velocity v for silica sphere beds with 275 μm and 375 μm particle diameters with best fit using
Eq. (7) and correlation coefficient. (b) Fanning friction factor versus Reynolds number from Eq. (8.2) with best fit.

Table 2
Various A and B values of Eq. (7) obtained for spherical particles. Data refering to Macdonald et al. (1979) correspond to the mean value and the standard deviation
(s.d.) when expressed. For quartz particles from Sartorius, dp = 275 μm corresponds to 0.25–0.30 mm and dp = 375 μm corresponds to 0.25–0.50 mm.

Ref. A B ɛ dp (μm) D/dp Red p Material Remarks about beads or


particles

Macdonald et al. (1979)a 132.7 1.291 0.366 to Quoted dp Red p /(1 − ε ) Glass Spherical and smooth glass
(±14% s.d.) (±16.5% 0.640 values report from ∼10−2 to particles, beads of narrow
s.d.) from ∼ 7.9 μm 104 size distribution
to 3.7 mm
Macdonald et al. (1979)b 321c 1.298 0.369 to Marble Spherical and smooth
(±11.5% 0.415 marble mixtures, uniform
s.d.) size or mixed in known
proportions
Bussière et al. (2006)d 273.5 10.54 0.38 275 ∼145 ∼2 to ∼50 Quartz Spherical and smooth
quartz particles
Bussière et al. (2006) d
241.3 2.077 0.41 375 ∼107 ∼8 to ∼65 Quartz Spherical and smooth
quartz particles
Bussière et al. (2006)d 409.6 0.454 0.37 250..315 ∼142 ∼8 to ∼110 Silica sand Silica sand particles with
aspect ratio ∼ 1.36
Bussière et al. (2006)d 317.6 0.464 0.36 355..400 ∼106 ∼10 to ∼142 Silica sand Silica sand particles with
aspect ratio ∼ 1.33
This worke 311.2 3.267 0.416 275 ∼145 ∼8 to ∼50 Quartz Spherical and smooth
quartz particles
This work e
279.9 2.417 0.403 375 ∼107 ∼10 to ∼95 Quartz Spherical and smooth
quartz particles
a
Reference therein Macdonald et al. (1979): Gupte AR, Dissertation, Karisruhe, 1970 and Rumpf H, Gupte AR, Chem. Ing. Tech. 43 (1971) 367
b
Reference therein Macdonald et al. (1979): Dudgeon CR, Houille Blanche, 21 (1966) 785
c
s.d. is not given because the uncertainty is too high
d
Working fluid is air at room temperature
e
Working fluid is water whose temperature is measured at each experiment

More information about the determination of k1 and k2 coeffi- 3.3. Expressions and results for sphere beds
cients in the case of air flow can be found in our previous work
(Bussière et al., 2006). We give in Table 2 these values expressed 3.3.1. Pressure drops
under the form of the coefficients A and B (see Section 3.3 for def- For the pressure drop we refer to the Ergun equation
inition) and the values obtained in the case of the current study (Ergun, 1952) and to the modified Ergun equation as quoted by
with the same method but with water flow. Macdonald et al. (1979) where the original numerical values are
W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72 65

Table 3
Comparison of k1 and k2 coefficients obtained for air flow (Bussière et al., 2006) and water flow with results from
Ergun (1952).

dp (μm) k1 (m2 ) k2 (m)

Air flow Water flow From Ergun (1952) Air flow Water flow From Ergun (1952)
−11 −11 −11 −6 −5
275 1.6 × 10 5.13 × 10 7.8 × 10 0.33 × 10 1.04 × 10 5.2 × 10−6
375 4.6 × 10−11 9.26 × 10−11 19 × 10−11 3.1 × 10−5 1.71 × 10−5 6.3 × 10−4
50 0 0 3.0 × 10−8 1.30 × 10−8 2.5 × 10−8 4.8 × 10−4 4.31 × 10−4 8.1 × 10−3

substituted by the coefficients A and B: valid whatever the type of flow. Some restrictions are given in Nika
(2008) for turbulent flows with A = 300 and B = 3.5 to take into ac-
P (1 − ε ) μ
2
1−ε ρ 2 count higher energy losses. Macdonald et al. (1979) have studied a
=A v+B 3 v (7)
L ε 3 D2p ε Dp wide range of experimental data – various materials and shapes
with A = 150 and B = 1.75 the original values of the two coeffi- such as glass spherical particles, cylindrical fibers, spherical mar-
cients given by Ergun (1952). Dp is the equivalent or effective di- bles either of uniform size or under the form of known mixtures
ameter of particles defined by using a sphere model, i.e. D p = DS = – in order to assess the two coefficients A and B of Eq. (8.2). Mean
6(1 − ε )/a from Eq. (2). In the case of a bed formed by identical values and standard deviations shown in Table 2 are calculated by
spherical particles then Dp is exactly the diameter of the spherical considering A and B as constant – even if a small dependence with
particle, thus D p = d p in Eq. (7) (Kaviany, 1999). ɛ is observed – for a given experimental configuration. Finally Mac-
Experimental and calculated pressure drops are given in Donald et al. recommend two sets of values: (A, B) = (180,1.8) for
Fig. 3(a) for 275 and 375 μm particles. For both particle diame- smooth particles and (A, B) = (180,4.0) for roughest particles as a
ters, approximations are obtained with satisfactory correlation co- good approximation for engineering purposes.
efficients and we obtain respectively (A; B) = (311.2;3.267) with Results are plotted in Fig. 3(b) with fitted approximations:
R2 = 0.9994 for 275 μm and (A; B) = (279.9;2.417) with R2 = 0.9999 (A; B) = (306.0;3.379) with R2 = 0.9987 for 275 μm and (A;
for 375 μm. Comparisons with published values obtained with B) = (280.2;2.413) with R2 = 0.9985 for 375 μm. Of course these val-
air flow and with formulations resulting from Ergun (1952) and ues are identical to those obtained in Section 3.3.1.
Rohsenow et al. (1998) are given in Tables 2 and 3. Values obtained with air as working fluid have to be considered
apart from other data corresponding to water as working fluid.
All the results are given in Table 3 in terms of k1 and k2 coeffi-
3.3.2. Friction factors
cients. The starting point of Eqs. (6)–(8.2) is the equation of Darcy
Kolodziej et al. (2012) highlighted that there are two main ap-
hence incremented by Forchheimer contribution with the quadratic
proaches to describe the pressure drop of a fluid flowing through
fluid velocity term depending on density. Eqs. (6)–(8.2) or other
stacked wire gauzes. In the approach flow through, the pressure
formulations (such as the semiheuristic momentum equation sug-
drop for the total bed length is given under the form of the Fan-
gested by Brinkman) have been established for incompressible
ning friction factor and Reynolds number is defined from the hy-
fluid (Rohsenow et al., 1998) which is not a valid assumption for
draulic diameter. The wire gauze is considered as a set or a bun-
air or other gases. For example Innocentini et al. (20 0 0) clearly
dle of given model tubes. In the second approach flow around, the
show the discrepancies between the results obtained by means of
pressure drop is given under the form of the drag coefficient ψ
the Forchheimer equation for incompressible fluids or the Forch-
calculated from the pressure drop for one gauze and from the frac-
heimer equation for compressible fluids.
tional opening of the gauze (different from porosity), the Reynolds
number being defined with the wire diameter. This latter is disre-
garded as the current work is concerned by stacks of woven metal 3.4. Conclusion for sphere beds
screens.
Eq. (7) can be expressed under the form of the Blake-type fric- In Fig. 4(a) and 4(b), A and B values obtained with water
tion factor following Ergun (1952) by: flow are plotted versus the tube-to-particle diameter ratio (D/dp )
and compared with the values quoted by Eisfeld and Schnitzlein
P D p ε 3 1−ε (2001) for which the D/dp ratio ranges from around 6 to 60. In
fk = = 150 + 1.75 (8.1)
L ρUm2 1 − ε NReD p Eisfeld and Schnitzlein (2001) additional results are given for D/dp
< 10 for which A values are significantly higher than 500. For
with the Reynolds number defined by: NReD = D p ρUm /μ, where
p the coefficient A, we notice that our results are consistent with
Um is the superficial fluid velocity measured at average pressure. higher quoted values (the highest value around 750 being dis-
The friction factor fk defined by (8.1) refers to the kinetic energy regarded) with D/dp ∼ 145 and 107 respectively for 275 μm and
losses, that is to say, the ratio of the total energy losses to the 375 μm. For the coefficient B, the value obtained for dp = 375 μm is
quadratic term representing the kinetic losses. Another formula- similar to published data except for dp = 275 μm where the value
tion was established referring to the ratio of the pressure loss to is significantly higher. Eisfeld and Schnitzlein (2001) has already
the viscous energy losses ( fv = 150 + 1.75 × NReD /(1 − ε )) but it is shown that an opposite trend is observed for B in comparison to A,
p
not used in the current work. Following the modified Ergun equa- namely a slight increase of B when increasing the D/dp ratio. Such
tion of Macdonald et al. (1979), an equivalent formulation is given a conclusion is reinforced by our measurements carried out with
by: D/dp ∼ 107 and 145.
To further the analysis, the influence of the wall proximity
ε3 1−ε
Fk  =A +B (8.2) is studied by means of the formulation given by Choi et al.
1−ε NRe  (2008) where the pressure drop is given by the relation:
ρV D
where Fk  1ε−ε = LP , NRe  = 0μ eq with Deq ≡ Dp and V0 ≡
3 Deq ε 3
ρV02 1−ε P (1 − ε )2 μM2 1 − ε ρ MCW 2
=A v+B 3 v (9)
w0 . The values for the two constants A and B are prescribed to be L ε3 D2p ε Dp
66 W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72

1000 4
From Eisfeld and Schnitzlein (2001)
Andersson (1963) Redp = 0.07 .. 1315
Bernard and Wilhelm (1950) Redp = 2 .. 2324
Ergun and Orning (1949) Redp = 0.4 .. 30
Jeschar (1964) Redp = 1 .. 16884
750 This work
3
dp = 275 µm water
dp = 375 µm water

500 2

B
A

250 1

0 0
0 25 50 75 100 125 150 0 25 50 75 100 125 150
D/dp D/dp
(a) (b)
6
6x10 1.00
dp = 275 µm
Choi et al. (2008)
with 311.2 and 3.267
6
5x10 Cw
0.99

6
4x10
0.98
ΔP/L (Pa/m)

6
3x10
Cw

0.97
6
2x10

6 0.96
1x10

0 0.95
0.0 0.1 0.2
-1
v (m.s )
(c)
Fig. 4. Comparison of the (a) A Andersson (1963), Ergun and Orning (1949), Jeschar (1964), Bernard and Wilhelm (1950) and (b) B coefficients obtained in the current work
for water with data published in Eisfeld and Schnitzlein (2001) (and references therein) for spherical particles. (c) Influence of the wall effect depicted by the formulation
given by Choi et al. (2008) Eq. (9) (CW is the wall correction factor).

where M and CW are two factors defined for taking into account experimental set-up is designed to perform realistic measurements
the effect of the proximity of the wall (see Choi et al., 2008) of pressure drops. Section 4 is dedicated to the study of pressure
for detailed explanations of the model and original reference). For drops and formulations in the case of woven metal screens used
dp = 275 μm, calculations provide M ≈ 1.008 while CW is plotted in for electrical safety purposes.
Fig. 4(c) from which we deduce that CW = 0.9750 +/- 0.0032 (s.d.)
for the velocity range we have considered, i.e. CW is very close to
unity for D/dp ∼ 145. Thus, our measurements can be considered 4. Experimental results for woven metal screens
as weakly dependent on the wall influence.
Experimental results and tests of various approximations dedi- Pressure drops due to woven metal screens are studied from
cated to spherical particle beds have shown that the whole of the the friction factor point of view. One of the major difficulties for
W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72 67

Table 4
Coefficients A1 and A2 deduced from Eqs. (10.1) and (10.2)
for the woven metal screens (30 measurement points for
each).

N° A1 A2 R2

1 – – –
2 – – –
3 583 0.613 0.93
4 637 0.153 0.96
5 396 0.572 0.97
6 422 0.976 0.98
7 653 0.268 0.98
8 – – –

combination of these two effects could explain the incoherents


trends obtained for the three stacks. Considering the screens 3 to
7, the best approximation is given by:

534
f = + 0.505 (10.3)
Fig. 5. Evolution of the Fanning friction factor f for stacks of woven metal screens ReDh
using the hydraulic diameter Dh . The curve corresponds to the best fit following
Eq. (10.2) (screens 1, 2 and 8 being disregarded) with A1 = 534 and A2 = 0.505. with a correlation coefficient R2 = 0.9248. Thus, we can conclude
that the Fanning friction factor, which is by definition the ratio of
the wall shear stress (τ W ) to the volumic density of kinetic energy
this kind of porous media lies in the definition of the characteris- (1/2ρ v2 ) (Rohsenow et al., 1998), can not be approximated by a
tic length depicting the fluid flow through the void parts of the relation like Eq. (10.2) with the Reynolds number defined from the
screens. In the following subsections, we test different formula- hydraulic diameter. Although Eq. (10.1) includes both the specific
tions or geometrical models, that is to say the hydraulic diameter, surface area and the porosity in the definition by means especially
the spherical diameter and the cylindrical diameter. As the litera- of the hydraulic diameter, results show that the approximation is
ture dealing with friction factor formulations is abundant, we aim not valid for the studied woven metal screens.
to select and test those dedicated to screens or stacks of screens
corresponding to nearly millimetric order wires screens. In the cur- 4.2. Fanning friction factor defined with the spherical diameter
rent work, the woven metal screens are built using wire diame-
ters (weft and warp wire diameters being different) ranging from The second formulation is established considering a spherical
0.18 mm to 1.6 mm and measurements of pressure drop are carried model to depict the fluid flow in the void space. Namely the spher-
out for stacks of these screens (typically five layers). ical diameter is used to write the friction factor and the Reynolds
number. This formulation has been widely used up to now in vari-
4.1. Fanning friction factor defined with the hydraulic diameter
ous types of studies. Measurements are given in terms of the Fan-
ning friction factor following the Ergun equation (Ergun, 1952):
As a starting point, the measurements are given in terms of the
Fanning friction factor following the Darcy–Weisbach equation as P D p ε 3
fk = (11.1)
recalled by Kolodziej and Lojewska, 2009a,b: L ρU 2 1 − ε
P 2ρw2e
= f (10.1) which is plotted versus the modified Reynolds number defined
L Dh D p ρU
by: 1Re
−ε = μ with the spherical diameter Dp defined above in
where the hydraulic diameter is defined by Dh = 4ε /a and the pro- −ε
Section 3.3.1. The equation for approximation is fk = A 1Re + B and
posed equation is:
we recall:
A1
f = + A2 (10.2) 1−ε
Re fk = 150 + 1.75 (11.2)
Re
with Re = ReDh . From Eq. (10.1) f is plotted in Fig. 5 as a func-
tion of the Reynolds number ReDh = ρ we Dh /μ and compared with for low Reynolds number (Ergun, 1952), while:
Eq. (10.2) to obtain the coefficients A1 and A2 given in Table 4
1−ε
for the studied metal mesh screens. Results are not displayed for fk = 300 + 3.5 (11.3)
Re
the screens 1, 2 and 8 because the trends observed for these
screens deviate significantly from Eq. (10.2) and incoherent val- for high Reynolds number (Nika, 2008). Experimental results and
ues are obtained for the two coefficients. We recall that in this approximations are plotted in Fig. 6(a) and (b) for the set of woven
study measurements are carried out for stacks of identical screens screens stacks. Approximation computed from Eq. (11.2) shows a
and not for only one screen with given and known characteristics. good agreement for Re/(1 − ε ) > 102 excepted for the screens 1, 2
The intermediate gap between two consecutive screens will intro- and 8, but it is far from experiments for Re/(1 − ε ) < 102 . Approx-
duce more or less deviations from these given characteristics and imation from Eq. (11.3) has to be disregarded both for the screens
the value of the hydraulic diameter can become doubtful due to and the Reynolds number range studied. As a result we conclude
variation of ɛ. Also it has been shown for example by Wu et al. that none of these approximations can be used with our screens
(2005) that metal screens of different woven types (plain square, stacks.
fill twill, fourdriner, plain dutch, twilled dutch) have different fric- In order to depict a wider range of Reynolds number on one
tion coefficients depending on the arrangement of the wires. The side, and to formulate a representative approximation for each
68 W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72

Fig. 7. Plot of Ergun-Bird equation following Eq. (12) with various values for C1 and
C2 coefficients. Are also given two approximations obtained from the measurements
only in the high Reynolds number domain (Re > 150): with coefficient C2 fixed at
3.5 and without any constraints.

fitted approximation being:


1−ε
 1 − ε 0.071
fk = 561 + 0.883 × (11.5)
Re Re
with the correlation coefficient R2 = 0.9225. The prediction bands
are plotted for ±30%. As a result we conclude that an approxima-
tion such as the one defined by Eq. (11.4) including the spheri-
cal diameter is not able to depict the pressure drop of the screens
studied in this work even if results can be considered as satisfac-
tory for screens numbered 3 to 7. From these two last sections we
also conclude that results obtained using Dh and D p = DS lead to
similar unsatisfactory trends for the studied screens.

4.3. Fanning friction factor defined with the cylindrical diameter

In the two last previous sections, we have shown that the ap-
proximations defined by means of the hydraulic and the spherical
diameters fail in the attempt to depict the pressure drops due to
the stacks of woven metal screens. In this section the following
Fig. 6. (a) Evolution of the Fanning friction factor fk for stacks of woven metal
points will be studied: the hypothesis of cylindrical diameters, the
screens using the spherical diameter Dp Eqs. (11.1)–(11.5). Measurements are com-
pared with the following approximations: the modified Fanning friction factors from definition of tortuosity, the effective velocity taking into account
Ergun (1952) for low (curve 1) and high (curve 2) Reynolds numbers; the empirical the influence of the tortuosity.
equation from Jones et al. (herein (Wu et al., 2005)) (curve 3) and the empirical As a first step before taking into account these latter points, we
equation from Wu (Wu et al., 2005) in the case of plain square type (curve 4). (b)
have to identify the weaknesses of the formulations. We start from
Empirical equation deduced for our measurements from Eq. (11.4) and spherical di-
ameter Dp following Wu et al. (2005) with prediction bands (−30% and + 30%).
the classical Ergun equations as cited by Bird (hence written EB),
Stewart, Lightfoot and given in Kolodziej and Lojewska (2009b):
P μw0 (1 − ε )2 ρ w20 1 − ε
= C1 + C (12)
type of metal screens on the other side, Wu et al. (2005) defined L D2p ε3 2
D p ε3
an empirical equation given by:
where Dp is the equivalent spherical diameter, with original values
1−ε
 1 − ε γW u C1 = 72 and C2 = 1.75 hence amended by C1 = 150 and C2 = 1.75,
fk = αW u + βW u (11.4) and by C1 = 300 and C2 = 3.5 for turbulent flows. These approx-
Re Re
imations are plotted in Fig. 7 in the case of the stack with the
where the coefficients α Wu , β Wu and γ Wu have to be determined screen numbered 7. We see that none of the EB equations with
from experimental pressure drops for each type of metal screens typical values for the coefficients C1 and C2 fit with the measure-
(plain square, plain dutch, full twill, twilled dutch types). The coeffi- ments. The two following approximations are performed only for
cient values obtained by Wu for plain square type are α Wu = 250, the points for which the modified Reynolds number is higher than
β Wu = 1.69 and γ Wu = 0.071, and the curve is shown in Fig. 6(a). 150: if C2 is set to 3.5 the resulting approximation shows a very
The agreement with our experimental data is better but still not poor correlation coefficient (R2 = 0.6597); if C1 and C2 are free of
satisfactory especially because of the trends for screens 1, 2 and 8. any constraints, then a rather acceptable correlation coefficient is
The fitting coefficients α Wu , β Wu and γ Wu have been obtained for obtained but the values of the two coefficients are far from classi-
all the experimental results disregarding the screens 1, 2 and 8, the cal values and they have no meaning (respectively 1092 and 4.8).
W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72 69

Table 5 Table 6
Coefficients C1 and C2 obtained for the stacks by means of Experimental values of the angle θ deduced from the application of
Eq. (12) but with the cylindrical diameter instead of the Eq. (13.1) following Kolodziej and Lojewska (2009b) and Eq. (14) from
spherical diameter. Kolodziej et al. (2012).

N° C1 C2 R2 N° τ θ from Eq. (13.1) θ from Eq. (14)


1 2474 1.42 0.58 1 1.160 84.0 74.7
2 1603 3.09 0.54 2 1.165 85.5 78.3
3 893 1.90 0.92 3 1.170 82.0 67.9
4 661 1.42 0.96 4 1.180 79.4 57.9
5 416 2.02 0.87 5 1.175 79.7 60.2
6 412 3.42 0.88 6 1.195 82.7 66.0
7 485 3.20 0.93 7 1.185 82.4 67.1
8 381 3.98 0.98 8 1.175 81.9 66.9

In a second stage, the spherical diameter is replaced in


Eq. (12) by the cylindrical diameter defined by DC = 4 1−a ε result- 0.079
ft = (13.4)
ing in the same equation with Dc equal to dw in Kolodziej et al. Re0e .25
(2012). The results for the constants C1 and C2 are given in Table 5.
These results are obtained disregarding the experimental points for where fapp and ft are respectively the friction factors for the lam-
low Reynolds number (<150) as previously done. Since the spheri- inar and turbulent domains, DC is used in place of dw the wire
cal case, results do not show reliable values for the two constants. diameter, as proposed in Kolodziej and Lojewska (2009b). The tor-
Plots are not shown but the fitted curves are very similar to those tuosity is calculated from τ = 1 + (1 − ε )/2 which corresponds to
obtained for example in Fig. 7 for the woven metal screen 7. a geometric model of a cubic cell with one sphere inside as re-
Whatever the formulation tested up to now, we observe that ferred in Harris et al. (2001). The parameter χ + is a dimension-
there are two main trends depending on the range of the Reynolds less quantity characterizing the length of the elementary chan-
number, namely low or high Reynolds number with a transition nel for the flow within a mesh of the screen; Ree is the effective
value around 150. These two trends have been already observed Reynolds number defined from the hydraulic diameter; fapp is de-
for example in Kolodziej and Lojewska (2009b) by Kolodziej et al. duced from Shah and London (1978) and ft is the Blasius formula
In the current experiments all the Reynolds numbers ReDh , ReD p for smooth pipes in the turbulent domain with coefficients given
and ReDc are lower than ∼ 4 × 103 and for most of the studied in Kolodziej and Lojewska (2009b), Kolodziej et al. (2012), Yan and
woven screens they are smaller to 103 for the velocity investigated Lin (1999) and Ouyang and Aziz (1996). The experimental pressure
range. So firstly, one has to pay great attention to the transient drops are fitted by means of Eq. (13.1), the fitting parameter be-
domain between low and high Reynolds numbers which should be ing θ whose values are given in Table 6 together with the tortu-
described more accurately, and secondly, the spherical diameter Dp osity deduced from the porosity of woven metal screens. The ex-
has to be substituted with the cylindrical diameter DC . perimental pressure drop is compared with the approximation in
We now discuss these two key points. For the depiction of the Fig. 8(a) for θ = 82.4° obtained for the woven screen 7, and for <
transient domain, the improvement suggested by Kolodziej et al. θ > = 82.4° the mean angle value obtained for the whole of the
refers to previous studies (Fowler and Hertel, 1940; Harris et al., studied woven screens. Satisfactory agreement is reported between
2001), concerned by porous media consisting of fibers from var- experiment and approximation for which discrepancies are inferior
ious materials. For such porous media, the path of the flow can to 10% at the most, except for both smaller and higher values of
not be approximated by the classical effective velocity given by measured velocities.
we = w0 /ε but one has to take into account θ – the angle between Other formulations have been tested for the turbulent friction
the direction of the flow inside the porous medium and the di- factor. In Ouyang and Aziz (1996) one finds an exhaustive sec-
rection given by the thickness of the crossed porous medium – tion dealing with the coefficients of the formulation first pub-
and the tortuosity. Following Kolodziej and Lojewska (2009b) and lished by Blasius. The relation is given under the form f = r Re−t
Harris et al. (2001) the effective flow length and the porous thick- with various values for the coefficients r and t deduced from
ness are linked by: Le = Lτ /cos(θ ), and the effective flow velocity the bibliography. In particular, the power law has been modi-
is defined by: we = (w0 /ε ) · (τ /cos(θ ) ). These improvements are fied with t ranging from 0.079 down to 0.003678, the explana-
of interest for the current study. Different stacks of woven screens tion from Ouyang and Aziz (1996) arguing the various Reynolds
were tested, where woven screens have different porosities and number domains and the very simple form of the approxima-
different diameters for the weaved wires. tion which may fail in representing real situations. Among the
At the end, the resulting pressure drop is the sum of the lami- power laws and considering Reynolds number range, we have
nar and turbulent contributions and the pressure drop is given by tested those with r = 0.316 and t = 0.25 from Fang et al. (2011) (for
(see Kolodziej and Lojewska (2009) for the full meaning and for- smooth pipes with Re < 2 × 104 ), and r = 0.02118 and t = 0.1461
mulation of each term) as: from Ouyang and Aziz (1996) (for smooth pipes up to Re ≤
2 × 106 ). The values obtained for θ show small differences ∼ 5% at
P ρ w20 1 − ε τ 3 the most. This result is consistent with two considerations. Firstly
= 4( fapp + ft ) (13.1)
L 2DC ε 3 cos3 (θ ) the Reynolds number range studied in the current work since one
has ReDh < 4 × 103 << 2 × 104 which appears as a high limit for the
with:
use of t = 0.25 (Fang et al., 2011). Secondly, for all the studied wo-
dw ρ w e Dh ven metal screens, the turbulent contribution in the total friction
χ+ = , Ree = (13.2)
Dh Ree μ factor is inferior to 10%.
The discussion about the influence of tortuosity is improved by
1.25 3.44
+ 16 − √ Kolodziej et al. (2012), concerning the effective flow length and
3.44 4χ + χ +
fapp Ree =  + (13.3) the effective flow velocity, where the formulations are respectively
χ + 1 + 0.0 0 021/χ +2 amended by: Le = L/cos(θ ) and we = w0 τ /ε . The resulting pressure
70 W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72

drop is then:
P ρ w20 1 − ε τ 2
= 4 ( f l + ft ) (14)
L 2DC ε 3 cos (θ )
where fl ≡fapp , τ is unchanged and we still use DC instead of dw as
in the original equation in Kolodziej et al. (2012). The results for θ
are given in Table 6 and the approximation for the woven screen
7 is shown Fig. 8(b) for θ = 67.1° The approximation deduced from
Eq. (14) is closer to the experimental points than it is with the use
of Eq. (13.1) but the trend is very similar.
One of the key point is the definition of the tortuosity used
in Eqs. (13.1) and (14). Many expressions can be found in the lit-
erature dealing with experimental formulations of the tortuosity
obtained for various materials and especially for woven screens.
Among them we first refer to Chinda and Brault (2012) for packed
cylindrical particles porous medium. A rather good agreement is
found between the depicted tortuosity model and the formula-
tion from Mauret and Renaud (1997) given by τ = 1 + 1.55 ln(1/ε )
called capillary model, under the condition that the porosity is not
too low, roughly higher than 0.5, depending on the geometry of the
cylindrical particles from the examination of the results. We have
tested this tortuosity formulation in Eq. (14), the tortuosity in-
creasing from 1.37 to 2.52 for the screens numbered 1 to 8. Results
(not detailed in the paper) show that it is not possible to fit our ex-
perimental pressure drops with some screens, which is mainly due
to the high value of tortuosity. We also refer to Armour and Can-
non (1968) and Fischer and Gerstmann (2013) for woven screens.
In Fischer and Gerstmann (2013), for woven screens of the dutch
twilled type (DTW) with 40 μm and 70 μm wire diameters, a Q tor-
tuosity factor of 1.3 is prescribed by means of formulations given
in Armour and Cannon (1968). In Armour and Cannon (1968) the
value Q = 1 is set for plain weaves and Q = 1.28 for dutch weaves.
The estimation deriving from Eq. (14) is applied in Fig. 8(c) with
Q = 1.28. Once again, the approximated trend is very similar to pre-
vious results.
In Fig. 9(a) we draw comparison with the experimental pres-
sure drops for the whole of the tested woven screens with the
calculated pressure drops obtained by means of Eq. (14) and τ =
1 + (1 − ε )/2 (results obtained with τ = 1 + 1.55 ln(1/ε ) are not
shown since they are very similar). We observe that there is a good
agreement for all the screens except for the woven screen 8. The
Fanning friction factors f = fl + ft are plotted in Fig. 9(b) for all
the woven screens together with the fitting curve (except woven
screen n°8) defined by f = 7.357 × Re−0.596 + 0.04351. This formu-
lation is chosen because the turbulent contribution is significantly
higher than the laminar contribution in the resulting Fanning fric-
tion factor f.

5. Conclusion

To improve the efficiency of LV circuit breakers and to com-


ply with IEC guidelines, researchers and engineers have to improve
the modeling tools used to simulate physical parameters and their
evolutions during electrical safety apparatus operation. One of the
key point studied in this work is the drag coefficients of woven
metal screens when submitted to the flow of a fluid. In fact these
screens are used as filters in LV circuit breakers and they play var-
ious important roles when an electric fault occurs. Firstly from the
thermal point of view they significantly reduce the temperature,
Fig. 8. (a) Experimental pressure drop for the woven metal screen 7 (●) with the
ionization degree, pressure, velocity, energy of the exhaust gases
approximation (Kolodziej and Lojewska, 2009b) from Eq. (13.1) calculated for the
fitted angle θ = 82.4° (red line) and for θ = 82.2° (blue line) the mean value calcu- consecutive to the arc plasma quenching. Secondly from the me-
lated for the whole set of screens. (b) The same as (a) but with the approximation chanical point of view, they tend to avoid the ejection of molten
from Eq. (14) (Kolodziej et al., 2012). (c) The same as (b) but with the tortuosity metal droplets resulting from the electrical contacts erosion and
value τ ≡ Q = 1.28. (For interpretation of the references to colour in this figure leg- they strongly damp the shock wave linked to the significant energy
end, the reader is referred to the web version of this article.)
release by the default current. To increase the efficiency of the LV
circuit breakers including such filters we have built an experimen-
W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72 71

Bernard, R.A., Wilhelm, R.H., 1950. Turbulent diffusion in fixed beds of packed
solids. Chem. Eng. Progress 46, 233–244 cited after [27].
Besnard, C., 2007. La simulation numérique pour la conception de cellules sécurisées
en cas d’arc interne. In: SEE (Ed.), Proceedings of the Third European Conference
on HV and MV Substation Equipment, S6-1. Lyon, p. 3.
Bussière, W., Rochette, D., Latchimy, T., Velleaud, G., André, P., 2006. Measurement
of Darcy and Forchheimer coefficients for silica sand beads. High Temp. Mater.
Processes 10, 55–78.
Chinda, P., Brault, P., 2012. Tortuosity model of a three-dimensional stacked cylin-
drical particles porous medium for using with the thin film solid oxide fuel cell
electrodes. TIJSAT 17 (1), 42–53.
Choi, Y.S., Kim, S.J., Kim, D., 2008. A semi-empirical correlation for pressure drop in
packed beds of spherical particles. Transp. Porous Med. 75, 133–149.
Eisfeld, B., Schnitzlein, K., 2001. The influence of confining walls on the pressure
drop in packed beds. Chem. Eng. Sci. 56, 4321–4329.
Ergun, S., 1952. Fluid flow through packed columns. Chem. Eng. Prog. 48 (2), 89–94.
Ergun, S., Orning, A.A., 1949. Fluid flow through randomly packed columns and flu-
idized beds. Ind. Eng. Chem. 41, 1179–1184 cited after [27].
T.R. Faber, A.A. Maulandi, M. Rival, D.R. Pearson, 2003 US Patent No. 0048586.
Fang, X., Xu, Y., Zhou, Z., 2011. New correlations of single-phase friction factor for
turbulent pipe flow and evaluation of existing single-phase friction factor cor-
relations. Nuclear Eng. Des. 241, 897–902.
Fischer, A., Gerstmann, J., 2013. Flow Resistance of Metallic Screens in liquid,
Gaseous and Cryogenic Flow. EUCASS, p. 12.
Fowler, J.L., Hertel, K.L., 1940. Flow of a gas through porous media. J. Appl. Phys. 11,
496–502.
Handbook of Chemistry and Physics (2012), 92nd Ed. CRC Press.
Harris, D.K., Cahela, D.R., Tatarchuk, B.J., 2001. Wet layup and sintering of metal-con-
taining microfibrous composites for chemical processing opportunities. Compos-
ites 32, 1117–1126.
http://www.haverboecker.com, 2016, March Documentation Haver & Boecker, In-
formation – Toiles métalliques (Terminologie, types de tissage, et formes de
mailles), P 17F 327 072002 1 Fe.
Huber, M.L., Perkins, R.A., Laesecke, A., Friend, D.G., Assael, M.J., Metaxa, I.N., Vo-
gel, E., Mares, R., Miyagawa, K., 2009. New international formulation for the
viscosity of H2 O. J. Phys. Chem. Ref. Data 38 (2), 101–125.
IEC 62271-200 High voltage switchgear and controlgear – AC metal-enclosed
switchgear and controlgear for rated voltages above 1 kV and up to and includ-
ing 52 kV, first ed. (2003)
Innocentini, M.D.M., Pardo, A.R.F., Pandolfelli, V.C., 20 0 0. Influence of air compress-
ibility on the permeability evaluation of refractory castables. J. Am. Ceram. Soc.
83 6, 1536–1538.
Jeschar, R., 1964. Druckverlust in Mehrkornschüttungen aus Kugeln. Archiv Eisen-
hüttenwesen 35, 91–108 cited after [27].
Kadoya, K., Matsunaga, N., Nagashima, A., 1963. Viscosity and thermal conductivity
of dry air in the gaseous phase. J. Phys. Chem. Ref. Data 14 4, 1985.
Kaviany, M., 1999. Principles of heat transfer in porous media. Mechanical Engineer-
ing Series, 2nd ed. Springer Verlag.
Kolodziej, A., Jaroszynski, M., Janus, B., Kleszcz, T., Lojewska, J., Lojewski, T., 2009.
Fig. 9. (a) Calculated versus experimental pressure drops for the whole of the wo- An experimental study of the pressure drop in fluid flows through wire gauzes.
ven screens. Approximation is deduced from Eq. (14) with τ = 1 + (1 − ε )/2, and Chem. Eng. Comm. 196, 932–949.
Kolodziej, A., Lojewska, J., 2009a. Flow resistance of wire gauzes. AIChE J. 55 (1),
the prediction bands are given for 90%. (b) Fanning friction factors f = f l + ft de-
264–267.
duced from the same approximation for the whole of the tested woven screens.
Kolodziej, A., Lojewska, J., 2009b. Experimental and modelling study on flow resis-
tance of wire gauzes. Chem. Eng. Process. 48, 816–822.
Kolodziej, A., Lojewska, J., Jaroszynski, M., Gancarczyk, A., Jodlowski, P., 2012. Heat
tal set-up designed to assess the pressure drops and drag coeffi- transfer and flow resistance for stacked wire gauzes: experiments and mod-
cients of woven metal screens or stacks of those. From the mea- elling. Int. J. Heat Fluid Flow 33, 101–108.
Macdonald, I.F., El-Sayed, M.S., Mow, K., Dullien, F.A.L., 1979. Flow through porous
surements given in terms of friction factors we have studied vari- media – The Ergun equation revisited. Ind. Eng. Chem. Fundam. 18 (3), 199–208.
ous models to define the correct equation by taking into account Mahjoob, S., Vafai, K., 2008. A synthesis of fluid and thermal transport models for
cylindrical diameters, angle between the direction of the flow in- metal foam heat exchangers. Int. J. Heat Mass Transfer 51, 3701–3711.
Mauret, E., Renaud, M., 1997. Transport Phenomena in Multi-Particle Systems II: Pro-
side the porous medium and the direction given by the thickness posed New Model Based on Flow around Submerged Objects for Sphere and
of the crossed porous medium, and specific contribution of lami- Fiber Beds Transition between the Capillary and Particulate Representations.
nar and turbulent domains. This equation is essential in LV circuit Chem. Eng. Sci. 52, 1819–1834.
Moreira, E.A., Coury, J.R., 2004. The influence of structural parameters on the per-
breakers modeling for new design developments. meability of ceramic foams. Braz. J. Chem. Eng. 21 (1), 23–33.
Morel, R., Rival, M., 1990 Eur. Patent No. 0437151B1.
Acknowledgment Multiphase Flow Handbook, (2006), ed. by C.T. Crowe, CRC, Part 10-1.
Nika, P., 2008 Convertisseurs thermoacoustiques – Moteurs et générateurs (in
french) Techniques de l’ingénieur Doc. BE 8061 1–25.
Authors would like to thank Mr. Michael Rips from Haver & Ouyang, L.B., Aziz, K., 1996. Steady-state gas flow in pipes. J. Pet. Sci. Eng. 14,
Boecker for having provided us with the parameters of the screens, 137–158.
Pope, K., Naterer, G.F., Wang, Z., 2011. Pressure drop of packed bed vertical flow for
and Mr. Jean-Paul Gonnet, Charles Besnard and Marc Rival from
multiphase hydrogen production. Int. J. Hydrogen Energy 36, 11338–11344.
Schneider Electric for financial support and helpful discussions. Renard, J.B., Worms, J.C., Lemaire, T., Hadamcik, E., Huret, N., 2002. Light scattering
by dust particles in microgravity : polarization and brightness imaging with the
References new version of the PROGRA2 instrument. Appl. Optics 41 (4), 609–618.
Rival, M., Kilindjian, C., Clery, Y., Bonente, S., 1999 US Patent No. 5889249.
Andersson, K.E.R., 1963. Pressure Drop in Packed Beds. Trans. of the Royal Inst. of Rochette, D., Clain, S., André, P., Bussière, W., Gentils, F., 2007. Two-dimensional
Technol., Stockholm Sweden, N°201, cited after [27]. modelling of internal arc effects in an enclosed MV cell provided with a pro-
Armour, J.C., Cannon, J.N., 1968. Fluid flow through woven screens. AIChE J. 14 (3), tection porous filter. J. Phys. D 40, 3137–3144.
415–420. Rochette, D., Clain, S., Bussière, W., André, P., Besnard, C., 2010. Porous filter op-
Belforte, G., Raparelli, T., Viktorov, V., Trivella, A., 2009. Metal woven wire cloth timization to improve the safety of the medium-voltage electrical installations
feeding system for gas bearings. Tribol. Int. 42, 600–608. during an internal arc fault. IEEE Trans. Power Delivery 25 (4), 2464–2470.
72 W. Bussière et al. / International Journal of Heat and Fluid Flow 65 (2017) 60–72

Rochette, D., Clain, S., Gentils, F., 2008. Numerical investigations on the pressure Schneider Electric–Square D, Arc flash protection with Masterpact NW and NT cir-
wave absorption and the gas cooling interacting in a porous filter, during an cuit breakers, Data Bulletin 0613DB0202R603 (2003).
internal arc faults in a medium-voltage cell. IEEE. Trans. Power Delivery 23 (1), Shah, R.K., London, A.L, 1978. Laminar Flow Forced Convection in Ducts. Academic
203–212. Press, NY p. 97 and 98.
Rochette, D., Clain, S., Gentils, F., Wild, J., Bussière, W., 2011. Numerical simulation Woven Wire Cloth, 2004. Reference Book, 4th edition Haver & Bocker.
of the porous filter properties for the internal arc mollifying effects. Elec. Power Wu, W.T., Liu, J.F., Li, W.J., Hsieh, W.H., 2005. Measurement and correlation of hy-
Syst. Res. 81, 66–73. draulic resistance of flow through woven metal screens. Int. J. Heat Mass Trans-
Rochette, D., Clain, S., Gonnet, J.P., Bussière, W., 2015. Numerical study of the impact fer 48, 3008–3017 Herein references: D.P. Jones, H. Krier, Gas Flow Resistance
of filters located in the exhaust duct of a low-voltage circuit breaker. IEEE Trans. Measurements Through Packed Beds at High Reynolds Numbers, J. Fluids Eng.
Comp. Packag. Manuf. Technol. 5 (1), 49–56. 105 2 (1983) 168-172.
Rohsenow, W.M., Hartnett, J.P., Cho, Y.I., 1998. Handbook of Heat Transfer. Mc- www.schneider-electric.com, (2007) Présentation de la norme CEI 62271-202 (in
Graw-Hill, p. 9. Chapters 4. French).
Schmid, R., Stuff, R., Klein, U.K.A., Jamjoon, F.A., Al-Suwaiyan, A., 1999. Low Yan, Y.Y., Lin, T.F, 1999. Condensation heat transfer and pressure drop of refrigerant
free-stream turbulence in test sections through packed beds and fibrous mats. R-134a in a smal pipe. Int. J. Heat Mass Transfer 42, 697–708.
Exp. Fluids 26, 451–459.

You might also like