You are on page 1of 11

Agricultural Water Management 126 (2013) 64–74

Contents lists available at SciVerse ScienceDirect

Agricultural Water Management


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

An experimental and analytical study to analyze hydraulic behavior of


nozzle-type underdrains in porous media filters
G. Arbat a,∗ , T. Pujol b , J. Puig-Bargués a , M. Duran-Ros a , L. Montoro b ,
J. Barragán c , F. Ramírez de Cartagena a
a
Department of Chemical and Agricultural Engineering and Technology, University of Girona, C. de Maria Aurèlia Capmany, 61, 17071 Girona, Spain
b
Department of Mechanical Engineering and Industrial Construction, University of Girona, C. de Maria Aurèlia Capmany, 61, 17071 Girona, Spain
c
Department of Agricultural and Forestry Engineering, University of Lleida, Av. de l’Alcalde Rovira Roure, 191, 25198 Lleida, Spain

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

Article history: Sand filters are commonly used to protect drip irrigation systems especially when large amounts of
Received 22 January 2013 organic contaminants are present. Moreover sand filters have the advantage of simplicity and large
Accepted 3 May 2013 capacities in comparison with other types of filters.
Available online 29 May 2013
The main objectives of this paper are firstly to develop an analytical equation to calculate the total
pressure drop in a sand filter taking into account the pressure drop in the underdrain and secondly to
Keywords:
validate this equation comparing its results with the obtained experimentally.
Microirrigation
An experimental study with a scaled commercial porous media filter was conducted with different
Head losses
Sand media filter
media bed depths and sand grain sizes. The results obtained with different media bed depths revealed
Packed bed that the Ergun equation underpredicted the pressure drop in the porous media filter.Subtracting the
Fluidization pressure drop caused by experiments with different sand heights shows that Ergun equation works well
Mathematical modeling for predicting the pressure drop in a region of the sand bed sufficiently far from the underdrain, but fails
Backwashing in the region immediately next to this element. To overcome this problem, a new model based on a set of
connected channels of the same diameter and progressive reduction of its number as the flow approaches
the nozzles has been developed. The results of the model were compared with the ones obtained in the
scaled sand filter as well as the ones obtained in commercial sand filters (Mesquita et al., 2012). The new
analytical equation improves the results of the Ergun equation to predict the pressure drop produced by
the entire sand bed by taking into account the effect of the underdrain (nozzle-type) and can be applied
to accurately predict pressure drop in the commercial sand filters commonly used in drip irrigation.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction Media filters used in agriculture are usually packed in a pres-


surized cylindrical tanks (with internal diameters ranging from 30
Water quality is a major concern in the management of drip to 122 cm), disposed vertically and using sand as a filtering media
irrigation systems as the solids contained in the irrigation water (Burt and Styles, 2007).
can clog the small orifices and flow paths of the emitters. Thus fil- Head losses in a media filter are produced by the friction loss
tration is essential for the successful operation of micro-irrigation due to the internal auxiliary elements of the filter container and
systems because it helps to prevent emitter clogging (Ayars et al., the resistance offered by the media (Testezlaf, 2008; Burt, 2010).
2007). The three common filter types used in micro-irrigation The prediction of the pressure drop produced by media filters in
systems are screen, disk and media filters. Media filters are usu- clean condition has a practical interest as the designer must select
ally the standard method for protecting micro-irrigation systems pump that overcomes the friction of a clean filter, plus an additional
(Trooien and Hills, 2007) and are specially recommended when loss due to accumulated dirt. When the total loss equals a pre-set
large amount of organic particles contaminants are present (Haman value, the filter will automatically backflush (Chang et al., 1999;
et al., 1994; Burt, 1996; Capra and Scicolone, 2004; Duran-Ros et al., Clark et al., 2007; Mesquita et al., 2012). However, few studies ana-
2009). lyze the effect of the auxiliary components of media filters on head
loss, which is related with water and energy consumption as well
as filter efficiency.
The backwashing is carried out reversing the direction of the
∗ Corresponding author. Tel.: +34 972 418 459; fax: +34 972 418 399. water flow in the filter tank. Thus the sand bed is fluidized drag-
E-mail addresses: gerard.arbat@udg.edu, gerard.arbat@gmail.com (G. Arbat). ging out the dirt and debris from the filtering media. Flow velocity

0378-3774/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.agwat.2013.05.004
G. Arbat et al. / Agricultural Water Management 126 (2013) 64–74 65

during backwashing should not be too high as it may produce the pressure drop in the sand filter (sand bed plus nozzles), and (iii) to
loss of sand particles from the filtering media. For this purpose it validate the model in comparison with experimentally measured
is interesting to know the minimum fluidization velocity, which pressure drop data obtained in a scaled filter as well as in a com-
can be computed from the procedure described in McCabe et al. mercial one.
(2001).
The Ergun equation has been extensively used to predict head 2. Materials and methods
losses in the filter media (Macdonald et al., 1979; McCabe et al.,
2001) since there is a general consensus on its accuracy. However, 2.1. Experimental setup
its application is strictly valid in the limit of an infinitely extended
packed bed (Nemec and Levec, 2005; Choi et al., 2008). The bound- A scaled sand filter was constructed based on a commercial sand
ary condition at the sides of the packed bed, known as the wall filter (Regaber, Parets del Vallès, Spain) of 500 mm internal diam-
effect, has been studied by numerous investigations. These stud- eter and with a filtration surface of 1963 cm2 . The dimensions of
ies demonstrate that when the bed diameter is not significantly the scaled filter were chosen such that the Reynolds number cor-
larger than the particle diameter the friction produced by the wall responding to the circular tank of the filter was similar to that of
of the tank is not negligible, invalidating the applicability of Ergun the commercial filter. Since the commercial filter had 12 nozzles
equation (Foumeny et al., 1993; Winterberg and Tsotsas, 2000; Choi and the scaled filter 1 nozzle, the scaling factor for all lengths was
et al., 2008). The boundary condition near the filter underdrain also approximately 1/121/3 . Therefore the volume flowing through each
violates the assumption of an infinitely extended packed bed made nozzle was the same in both filters. The dimension of the nozzle
by the Ergun equation. The capillary model based on the hydraulic and its passing area is described in detail in Section 2.3. The filter
radius concept, the physical model in which is based the Ergun had a 200 mm inner diameter and 757 mm total length. An inner
equation, assumes that porous media consists of a set of parallel plate to fit the underdrain element (nozzle) and to support the sand
identical channels (McCabe et al., 2001; Nemec and Levec, 2005). was welded at 140 mm above the bottom of the filter. During the
This simplified model of the flow within the porous media may trials the pressure was measured at both inlet and outlet of the
fail to represent the actual flow occurring near the underdrain ele- sand filter by means of a Leo 2 (Keller, Winterthur, Switzerland)
ments since the streamlines must converge near the passing area digital manometer (±0.07% accuracy). The inlet and outlet filter
through the slots of the underdrains. Indeed, different empirical pipes could be easily switched in order to operate at either in the
studies have shown that an important part of the total pressure filtration or in the backwash regimes. A centrifugal pump Prisma
drop in commercial sand filters is due to the underdrain system 20/4 M (ESPA, Banyoles, Spain) was used to suction the water from
(Burt, 2010; Mesquita et al., 2012). The application of computa- a tank and to provide the pressure to the filter. The water flow was
tional fluid dynamics (CFD) has been previously used to understand measured using a volume accumulator 405 S (Sensus, London, UK).
the flow under the walls of a packed bed (Reddy and Joshi, 2010; A screen filter (Regaber, Parets del Vallès, Spain), 165 mm inside
Riefler et al., 2012) and also in the nozzle-type underdrain of a sand diameter, 508 mm inlet and outlet pipes and 120 ␮m filtration level
filter (Arbat et al., 2011). In this work a nozzle is defined as the was installed between the water tank and the sand filter in order
underdrain element formed by a pod with multiple slots that allows to prevent the entrance of particles to the sand filter and to retain
water flow during filtration and backwashing regimes and prevents the sand during the backwashing of the sand filter. The scaled sand
losses of the filtering media. In commercial sand filters the nozzles filter and installation layout is depicted in Fig. 1.
are distributed at the bottom of the sand bed.
In the latter case the authors showed the importance of the rela- 2.2. Experimental procedure
tionship between the passing area of the nozzles and the perforated
plate in the resulting pressure losses in order to prevent high flow The characteristic curve of the sand filter, i.e., the relationship
velocities in the sand bed. These authors also showed that when between superficial velocity (V0 ) and the pressure difference from
this relationship decreases, most of the pressure drop is produced the filter inlet and outlet (Dp = pin − pout ) was obtained at the filtra-
in a tiny area inside the packed bed close to the underdrain. These tion and backwash regimes for different filter media configurations.
previous works reveal the importance of properly including the The same procedure was followed in each individual experiment.
underdrain system when developing models for predicting head Initially, we close the valve at the filter outlet until the inlet pressure
losses in packed beds. Nevertheless it must be noticed that the reached approximately 400 kPa. With this position of the valve, the
application of CFD software to simulate the pressure drop in a sand flow-rate corresponding to 120 s was calculated from the volume
filter has strong limitations since it requires an enormous compu- accumulator readings and the minimum and maximum pressures
tational force to discretizate the complex geometries of the sand at the filter inlet and outlet (pin and pout ), corresponding to the
filter. same period of time, were registered with the digital manometer.
At the knowledge of the authors there is not a published After that, the procedure was repeated by opening the outlet valve
analytical solution that can be applied to compute the head loss until the outlet pressure was reduced approximately 20 kPa. The
produced by the underdrain of a porous media filter. In this work, procedure was repeated several times until the outlet valve was
and based on the reduction of the effective filter diameter near the completely opened, and thus the sand filter characteristic curve
underdrain, similar to the behavior of the streamlines in a conduit could be obtained. Each experiment was repeated 3 times to ensure
as it approaches an orifice (Jeong and Choi, 2005), we develop repeatability of the data.
a new physical model of a porous media in order to predict the
pressure drop in the sand filter that takes into account the effect 2.3. Filter configurations
of the nozzle-type underdrain. For this region of the filter, the new
analytical model consists of bundle of inclined channels of the Two different sand particle sizes and three different sand bed
same equivalent diameter that are merging together, reducing the weights (or, equivalently, heights) have been tested in the scaled
number of available channels as the flow approaches the slots of sand. Filtration and backwashing regimes were tested (Table 1).
the nozzle. The commercial silica sand used in the experiments was sieved
The main objectives of the paper are: (i) to experimentally ana- in order to obtain a narrow range of sand sizes. This allowed
lyze the hydraulic behavior of the underdrains (nozzles) in a sand us to work with two different sand sizes (0.63–0.75 mm and
filter, (ii) to present a new analytical equation to calculate the total 0.75–0.84 mm) whose main characteristics are shown in Table 2.
66 G. Arbat et al. / Agricultural Water Management 126 (2013) 64–74

Fig. 1. (a) Scaled sand filter, and (b) installation layout at the filtration configuration.

Table 1 (Regaber, Parets del Vallès, Spain) that was taken as a reference and,
Different filter configurations tested in the experiments with the scaled sand filter.
therefore, correspond to the nominal column height of the scaled
Flow regime Sand size Sand weight Media bed filter. The tests with 1.547 kg of sand with the nozzle gave a media
(mm) (kg) deptha (mm) bed depth of 20 mm over the mid height of the nozzle, which repre-
Filtration 0.63–0.75 1.547a 20.0 sents a very short sand column in relationship with the experiments
Filtration 0.63–0.75 3.860 70.1 with 13.530 kg of sand. The experiments with the short sand col-
Filtration 0.63–0.75 13.530 279.5 umn were designed to analyze the pressure drop generated in a tiny
Backwashing – None –
area over the underdrain, and to compare their results with those
Backwashing 0.63–0.75 13.530 279.5
Filtration 0.75–0.84 3.860 70.4 obtained in the proposed analytical model described in Section 3.
Filtration 0.75–0.84 13.530 280.4 Additionally, experiments with an intermediate media bed depth
a
Sand bed depths over the mid height of the nozzle. corresponding to 3.86 kg of sand were carried out. The difference
of the experimental pressure losses obtained in the experiments
with 13.530 and 3.860 kg of sand would allow to verify the pressure
Soyer and Akgiray (2009) showed that the sphericity s values
losses calculated with the Ergun equation in a 0.21 m long sand col-
in ten different silica sands ranged from 0.71 to 0.79. Currently it is
umn without the influence of the underdrain system. Similarly, the
considered that silica sand sphericity ranges from 0.70 to 0.80 (U.S.
difference of the pressure losses obtained with 13.530 and 1.547 kg
Department of Defense, 2003). In this study, and according with the
of sand would serve to know the pressure losses in a sand column
results showed in the literature, the sphericity for the 0.63–0.75 and
0.26 m long.
0.75–0.84 mm sand particle sizes was assumed to be 0.75.
The underdrain configuration was a nozzle of identical charac-
2.4. Fluidization velocities at the backflow regime
teristics than that of the commercial sand filter (Regaber, Parets
del Vallès, Spain) assembled on the orifice at the middle of the
The minimum fluidization velocity (V̄0M ) in the experiments
plate with 23 mm diameter that separates the water bottom cham-
with 13.530 and 3.860 kg of sand (sizes from 0.63 to 0.75 mm) was
ber from the porous media (Fig. 2a and b). The nozzle had 45
calculated from Eq. (1), according to McCabe et al. (2001):
slots 0.45 mm wide and 30 mm long with a total passing area of
2
607.5 mm2 . The passing area through the orifice was 415 mm2 . 150V̄0M (1 − εM ) 1.75V̄0M 1
+ = g(p − ) (1)
The width of the slots was measured using feeler gauges ranging S2 Dp2 ε3M S Dp ε3 M
0.05 mm between two consecutive sizes. The gauge of 0.45 mm was
the wider that passed through the slots and there were no differ- where  is the absolute fluid viscosity, s is the sphericity of the
ences of width between them. As the scaled filter has only one sand particles, Dp the particle diameter, g is the acceleration of grav-
nozzle and 200 mm internal diameter, each nozzle servers an area ity,  is the fluid density, p the density of the sand particles and
of 314 cm2 . εM the porosity at incipient fluidization, which was taken as the
The experiments with 13.530 kg of sand represent approxi- fixed bed porosity of the sand. In fluidization studies it is commonly
mately the 1/12 of the volume of sand in the commercial sand filter accepted that εM can be taken equal to the fixed bed porosity (e.g.,
Torrez Irigoyen and Giner, 2011). At this respect Formisani et al.
(1998) stated that the determination of εM is cumbersome and the
Table 2
Main characteristics of the two silica sand sizes used in the study. procedure may not provide a value significantly different from the
fixed bed porosity.
Grain sizes Average sand Bulk density Particle density Porosity (%)
According to McCabe et al. (2001) the porosity of the expanded
(mm) diameter (mm) (kg m−3 ) (kg m−3 )
bed (εE ) can be calculated as:
0.63–0.75 0.690 1470 (16) 2556 (0) 42.50 (0.00) L 
0.75–0.84 0.795 1462 (39) 2507 (6) 41.67 (1.44) M
εE = 1 − (1 − εM ) (2)
The number in brackets is the standard deviation of three different samples.
LE
G. Arbat et al. / Agricultural Water Management 126 (2013) 64–74 67

Fig. 2. (a) Front view of the nozzle, (b) top view of the nozzle and (c) regions I and II in the sand bed.

where LM and εM are the bed height and porosity at incipient flu- successfully predicts the energy losses as a function of the volumet-
idization respectively and LE is the expanded bed height. ric flow rate (or, equivalently, the mean flow velocity) of the scaled
It was assumed that the maximum sand bed expansion would filter detailed in the previous section and, also, of a commercial
correspond to the filter length, as in this case the sand would reach filter thoroughly studied by Mesquita et al. (2012).
the outlet of the filter. Under these conditions the expanded bed Our analytical expression is based on a series pressure drop rep-
height (LE ) was equal to the filter height and εE was calculated from resentation from point 1 located at the inlet manometer to point
Eq. (2). The fluid velocity under this condition (V̄M ) can be calculated 2 located at the exit manometer (Fig. 1). Thus, the measured total
from the empirical relationship proposed by Lewis et al. (1949): pressure loss p12 follows:
 ε m V̄0 p12 = ps + pns (4)
E
= (3)
εM V̄M where ps and pns are the pressure losses through the porous
where the exponent m was estimated from the Reynolds number media and the non-porous media, respectively.
at the minimum fluidization velocity using the graphical method We divide pns into three main contributions
proposed in McCabe et al. (2001), which is based on the correlation
pns = pwi + po + pwe (5)
given by Leva (1959).
The fluid velocity V̄M was compared with the ranges of veloci- where pwi is the pressure drop from point 1 to the top of the sand
ties during the backflow tests in order to verify if there were the column (water inlet region), po is the pressure drop through the
conditions to produce the outflow of sand. It must be noticed that nozzle orifice and pwe is the pressure drop from the exit of the
during each backflow experiment the sand that leaves the filter was nozzle orifice to point 2 (water exit region). Note that these terms
intercepted and collected by a screen filter (Fig. 1) and then dried may include losses from auxiliary elements (e.g., backflush valves).
and weighted. On the other hand, the pressure loss through the sand bed ps
is divided into two terms
3. Nozzle type underdrain analytical model
ps = psI + psII (6)

In this section we develop an analytical model for predicting where psI,II correspond to the pressure losses in regions I and II
the total pressure loss in sand filters that employ nozzles as the shown in Fig. 2c. In region I, the flow is uniform through the sand
underdrain system. In the next section, we show that this model bed and the pressure drop follows the Ergun equation. In region II,
68 G. Arbat et al. / Agricultural Water Management 126 (2013) 64–74

of being formed by inter-connected cylindrical channels of equiv-


alent diameter Deq through which water flows (Fig. 3a). Analytical
expressions for all the terms shown in Eqs. (4)–(6) are derived in
Appendix.

4. Results and discussion

4.1. Zone of influence of the nozzle

The pressure drop in a sand bed column 0.21 m long filled with
sand with size ranging from 0.63 to 0.75 mm was calculated as the
difference between the pressure drop measured with 13.530 kg and
that with 3.860 kg of the indicated sand sizes (Table 1). Similarly,
it was calculated the pressure drop for the same sand bed height
with 0.75–0.84 mm sand size. The pressure drop in a sand bed col-
umn 0.26 m long filled with 0.63–0.75 mm sand size was calculated
doing the difference between the pressure drop measured with
13.530 kg and that with 1.547 kg. Since experimental data were
not obtained at regular intervals of superficial velocity, the previ-
ous differences were obtained after fitting a quadratic polynomial
equation to the pressure drop as a function of the superficial veloc-
ity. In all the cases the goodness of fit between the measurements
and adjusted values was very high (R2 > 0.996).
Thus, doing the difference of the pressure drop between two
independent experiments, and at the same flow rate, the pressure
losses in the auxiliary elements of the filter, as well as the ones
Fig. 3. Idealization of the pore space in the sand bed: (a) parallel channels in the
region I and (b) channel network in region II. produced in the lower part of the sand bed close to the nozzle cancel
out. Therefore the pressure drop can be compared with the one
predicted with Ergun equation, which is applicable to predict the
however, the flow is not uniform due to the influence of the nozzle. pressure losses in the sand bed (McCabe et al., 2001).
Since the drain opening area of the nozzle is smaller than the filter Fig. 4 shows that the pressure drop predicted with the Ergun
cross-sectional area, the fluid velocity within the sand increases as equation follows a similar trend than that obtained in the experi-
it approaches the nozzle. This leads to large values of pressure losses ments although it slightly over-predicts the pressure drop for the
in a very small zone close to the nozzle. In this region, the Ergun three cases. In both experimental and Ergun’s predictions, the lower
equation does not provide reasonable predictions. This effect was pressure drop was produced by the sand column of 0.21 m height
already noted by means of detailed CFD simulations of a commer- and the bigger grain size (0.75–0.84 mm), followed by the sand col-
cial filter in Arbat et al. (2011). Here, we develop a new expression umn of the same height filled with grain sizes from 0.63 to 0.75 mm.
for determining the contribution of the psII term based on a gen- The greatest pressure drop was produced by the longer sand bed
eralization of the idealized representation of the porous media (L = 0.26 m) filled with 0.63–0.75 mm sand size. As generally admit-
employed in McCabe et al. (2001). The porous media is assumed ted Ergun equation is suitable to predict pressure drop in infinitely

Fig. 4. Experimental and Ergun’s predictions of the pressure drop in the sand bed (p, kPa) for different superficial velocities (V0 , m h−1 ). The 0.21 m sand bed height (L)
labeled as “measured” comes from the subtraction of measured data of pressure losses for two amounts of sand volumes in order to cancel out the effects of auxiliary elements
and only retain the effect of the media bed depth. The same applies for the sand bed height (L) of 0.26 m. The direction of the flow is from the top of the porous media to the
nozzle.
G. Arbat et al. / Agricultural Water Management 126 (2013) 64–74 69

Fig. 5. Comparison of the experimental pressure losses at different flow rates with the ones predicted using the proposed model and Ergun equation. (a) Results corresponding
to the experiment with 13.530 kg of 0.63–0.75 mm sand diameter. (b) Results corresponding to the experiment with 3.860 kg of the same diameters. In both figures the direction
of the flow is from the top of the porous media to the nozzle.

extended packed bed, especially when the porosity range is rather is experimentally confirmed that the length of region II is less than
narrow (0.35 < ε < 0.55), the bed is made up of similar sized particles the 2.5 mm over the top of the nozzle.
and the flow rates are moderate (Nemec and Levec, 2005), which
was the case of the sand filter experiments presented here. 4.2. Scaled filter
Applying the model described in Section 3, the sand bed is
responsible of most of the pressure losses, roughly 90%, of which The model has been applied to simulate pressure losses at the
approximately 40% of the total pressure losses are produced in the filtration regime in a scaled filter that includes a single nozzle.
region ranging from the top of the sand bed to a short distance When Ergun equation was applied to predict the total head losses
from the nozzle (region I) and 60% in a very tiny region at the bot- in the sand filter filled with 13.530 kg of sand with size in the range
tom of the sand column, close to the nozzle slots (region II). The 0.63–0.75 mm, the pressure drop compared with the experimental
extension of region II, where most of the pressure losses are pro- data was clearly underpredicted (dotted line in Fig. 5a). Therefore,
duced, is initially unknown but the experiment with 1.547 kg of we include the effect of accessories to the Ergun equation in order
sand, where the sand column was as short as 2.5 mm over the top to take into account the pressure drop produced in the water inlet
of the nozzle, would confirm that it is restricted to a very tiny area (pwi ), orifice (po ) and water exit (pwe ), as described in Sec-
at the bottom of the sand column. Indeed, the experimental results tion 3. The results, however, still underpredicted the experimental
corresponding to a 0.26 m sand column (Fig. 4), obtained by sub- values (see the dashed line in Fig. 5a).
tracting the measured values of the 13.530 kg case with those of the When pressure drop was predicted with the proposed model
1.547 kg of sand, do not show any abrupt increase of the pressure (solid line) it was assumed that the region with a uniform flow
drop compared with the one obtained with a 0.21 m sand column behavior in the porous media (region I) ranges from the top of the
and follows the good agreement with Ergun equation. Therefore it sand bed to the top of the nozzle (Lu = 0.262 m). Thus, region II in
70 G. Arbat et al. / Agricultural Water Management 126 (2013) 64–74

the sand bed is limited to the distance between the top of the noz-
zle to the center of the nozzle slot (Ln − Lu = 0.0175 m). Note that
the length of the non-uniform flow is limited to the center of the
nozzle slot because this is the mean length of the streamlines from
the ending of region II to the nozzle exit (it must be noticed that
the slots are arranged vertically and the water can cross them at
different heights, but the mean travel length taking into account all
the possible pathways will correspond to 0.0175 m). The assump-
tion that the ending of region I reach the top of the nozzle leads to
accept a non-uniform behavior in a region very close to the noz-
zle only. This assumption was confirmed in the previous section
when comparing Ergun equation for different media bed depths
with experimental data.
The pressure drop predicted with the proposed model was very
close to the experimental data (closed circles), especially when
the superficial velocity was greater than 70 m h−1 (Fig. 5a). As
stated above, this figure also shows the results for a simplified case
Fig. 6. Results predicted by the Ergun equation and our analytical model for the
(dashed line) where Ergun equation plus accessories were take into filter studied in Mesquita et al. (2012). For three different sand bed depths, lines
account but the nozzle effect was neglected, so region II in the sand correspond to the experimental values of the pressure drop as a function of the
bed was not taken into consideration and, therefore, the pressure superficial velocity (Mesquita et al., 2012), open symbols represent the values pre-
dicted by our model and open symbols with a cross correspond to the prediction of
drop through the porous media was exclusively modeled by Ergun
the Ergun equation when using a sand bed height equal to the whole sand depth.
equation. This case cannot explain the behavior of data, especially
at higher values of the volumetric flow rate, where the proposed
model clearly improves the pressure drop prediction. section we also detected this phenomenon but we carefully level
Fig. 5b shows the case for 3.860 kg of sand and the results off the surface at the end of each experiment. This hydrodynamic
are similar to those obtained with the 13.530 kg of sand. Thus, effect is beyond the scope of the analytical model introduced in
the proposed model predicts well the experimental pressure drop Section 3, so we only use the G1 grain size data of Mesquita et al.
for superficial velocities greater than 70 m h−1 . The Ergun equa- (2012).
tion itself underpredicts the experimental pressure drop for all the Results predicted by our analytical model for the filter studied in
range of superficial velocities. The pressure drop obtained adding Mesquita et al. (2012) are shown in Fig. 6. Since we ignore the details
the effect of the accessories to the ones obtained with Ergun equa- of the experimental set up employed in Mesquita et al. (2012) (i.e.,
tion is close to the experimental results for the lower range of exact location of the manometers, lengths and diameters of inlet
superficial velocities, from 54 to 72 m h−1 . and exit pipes, etc.), the pns term in Eq. (4) is not taken into con-
Similar results (not shown in the figures) were obtained in all sideration. This implies that our model would, in principle, predict
the other tests: 1.547 kg of sand with size ranging from 0.63 to a slightly higher pressure drop than that shown in Fig. 6.
0.75 mm, 3.860 kg of sand with size ranging from 0.75 to 0.84 mm The values of the parameters required for evaluating the ps
and 13.530 kg of sand with size ranging from 0.75 to 0.84 mm terms in Eq. (6) were extracted from the published data in Mesquita
(Table 1). et al. (2012) (with the volumetric flow rate divided by the number of
nozzles since ps makes reference to a single underdrain element).
4.3. Commercial filter Only two additional data not found in Mesquita et al. (2012) were
needed: the sphericity of the sand and the height of region II. The
In a very recent study, Mesquita et al. (2012) have analyzed the former is chosen equal to 0.75 in agreement with Section 2.3. The
effects of the sand particle size and sand bed depth on the pressure latter assumes a height equal to 20.6 mm, which, in consistency
losses through three different commercial filters. These three filters with the methodology employed in the scaled filter, corresponds
had different underdrain systems: manifold (F1 type), disk-type (F2 to the vertical distance from the upper part of the nozzle where the
type) and nozzle (F3 type). The latter fits the type of sand filters here slot begins to the center of gravity of the slot open area.
analyzed. For the F3 type they experimentally obtained the head Fig. 6 shows a remarkable agreement between our analytical
losses as a function of the volumetric flow rate for three different prediction and the experimental data, especially at higher sand bed
types of grain sizes (G1–G3) each one used in three different bed depths. In contrast, the Ergun equation applied to the whole sand
depths (H1–H3). Additional technical details of this F3 type filter bed height gives substantially lower values than the observed ones
can be found in Mesquita et al. (2012). since ignores the influence of region II. This confirms the validity of
In order to determine the zone of influence of the nozzle, we our analytical approximation for predicting the pressure losses in
performed the same analysis as that carried out in Section 4.1 but commercial filters with a nozzle type underdrain system.
applied to the data shown in Mesquita et al. (2012). For each grain
type, we plotted the difference of the head losses for two different 4.4. Conditions at the backflow regime
bed depths (Hi and Hj ) as a function of the superficial velocity. As
already discussed in Section 4.1, we expected that this difference As shown in Fig. 7 the pressure losses without sand and with
would match the value predicted by the Ergun equation when using different sand bed weights (3.860 and 13.530 kg) were minimal
a sand bed height equal to the difference of sand bed heights Hi − Hj . since at the backflow regime the sand bed was fluidized and
However, only the data for the grain size labeled as G1 in Mesquita therefore the amount of sand within the filter has little influence
et al. (2012) showed a reasonable behavior, since plots for other in the pressure losses.
grain types did not show a monotonic increase as a function of the The minimum fluidization velocity (V̄0M ), obtained from Eq.
superficial velocity. Such unexpected behavior may be caused by (1), was 12.2 m h−1 that was overcome in all the backflow experi-
the accumulation effect of sand in some regions, already detected ments. The difference in the mean velocities was minimal for the
in Mesquita et al. (2012), which may vary the effective sand bed experiments without sand and the ones with 3.860 and 13.530 kg
height. In few tests of the scaled filter analyzed in the previous of sand, being their average values and standard deviations of:
G. Arbat et al. / Agricultural Water Management 126 (2013) 64–74 71

Fig. 7. Pressure drop (p) for different velocities at the backflow mode during the different trials without sand and with 3.86 kg and 13.530 kg of sand (0.63–0.75 mm).

95.4 ± 22.7 m h−1 , 94.7 ± 20.5 m h−1 and 99.0 ± 18.7 m h−1 respec- by commercial sand filters using nozzle-type underdrains, which
tively. Therefore the mean velocities as well as the minimum are currently used in the filtration units used in microirrigation
velocity during the trials were well above the minimum fluidization systems.
velocity. As the sand particles were suspended the pressure losses
for the experiments without sand and for the experiments with
3.860 and 13.530 kg of sand were very similar. Acknowledgements
The porosities (εE ) at the maximum bed expanded conditions
(i.e., the case where the expanded sand bed reached the outlet of The authors would like to express their gratitude to the Spanish
the filter) for 13.530 and 3.860 kg of sand, calculated from Eq. (2) Ministry of Education and Science for its financial support of this
were respectively 0.7253 and 0.9167 and the corresponding veloc- study through grants CGL2009-09342/BTE, CGL2008-02832/BTE,
ities, calculated from Eq. (3), were 83.52 m h−1 and 194.40 m h−1 . FIS-2009-13050, to the Autonomous Government of Catalonia for
The velocities during the experiment with 13.530 kg of sand were grant 2009-SGR-374. The authors would also like to thank Sergi
greater than the one computed with Eq. (3), while during the exper- Saus, Jordi Vicens, Regaber and the Municipality of Celrà for their
iment with 3.860 kg of sand the velocity was always less than help in carrying out this investigation.
194.40 m h−1 (Fig. 7). Therefore in the experiment with 13.530 kg of
sand it would be expected that part of the sand bed would be drag
from the filter to the outlet. This was confirmed at the end of the Appendix A. [{(Appendix)}]
three replications of the experiment since 560 g of sand were recov-
ered in the screen filter. On the other hand, the screen filter was Here we derive the analytical expression for estimating the pres-
completely free of sand after the three replications of the experi- sure difference measured by the manometer at the inlet (point 1)
ment with 3.860 kg, as it was expected. At this respect it can be con- and that at the outlet (point 2) of the filter (Fig. 1). Eq. (4) evalu-
cluded that Eqs. (2) and (3), applied as described in Section 2.4 can ates this total pressure drop p12 in terms of sand plus non-sand
be applied to predict the maximum allowed velocity at the backflow contributions ps + pns whose analytical expressions are detailed
regime (V̄M ) in order to prevent sand outflow from the filter. below.
For all these regions, we take into consideration both friction
losses for a length of pipe L and local losses due to sudden expan-
5. Conclusions sions and contractions.
Pressure losses due to friction effects, major losses (pf ), are
The Ergun equation with a shape factor (sphericity) is well expressed in terms of the Darcy–Weisbach equation (White, 2011):
suited to predict the pressure drop produced by an infinitely
long packed bed. Nevertheless, the idealization of parallel chan-
nels, which is the physical model that supports Ergun equation, L  2 L 8
pf = f v = f 5 2 Q2 (A.1)
does not match the hydraulic behavior in the lower part of D2 D 
a filter since there exists an underdrain element (nozzle). The
experiments carried out confirmed that the flow through the where v is the mean flow velocity computed as v = Q/(D2 /4), D is
porous media is not uniform near the nozzle where stream- the pipe diameter along the length L and f is the friction factor. The
lines converge toward its slots. In the idealized new analytical latter is a function of the pipe relative roughness (here assumed
model here presented, this effect is implemented by a reduc- smooth) and of the Reynolds number:
tion of the number of channels available as well as a reduction
of the effective filter diameter. The pressure drop predicted with vD
Re = (A.2)
the proposed model clearly improves Ergun’s predictions, since 
remarkably fits experimental data, especially at relatively high val-
ues of the superficial velocity. The analytical model presented in where  and  are the fluid density and absolute viscosity, respec-
this paper can be applied to predict the pressure drop produced tively.
72 G. Arbat et al. / Agricultural Water Management 126 (2013) 64–74

On the other hand, local (or minor) losses (pk ) are obtained loss occurring in the water region inside the filter, pwe2 corre-
from (White, 2011): sponding to the local (minor) loss at the contraction zone where the
filter ends and the exit pipe starts and pwe3 corresponding to the
 2 8
pK = K v = K 2 4 Q2 (A.3) friction (major) loss occurring within the exit pipe until reaching
2  D
the position of the exit manometer. The last term pwe3 is equal
where K is the minor loss coefficient. to pwi1 Eq. (A.4) since the inlet and exit pipes are equal. Simi-
larly to pwi3 , pwe1 follows Eq. (A.4) with DF0 instead of d and
A.1. Non porous media L = 0.141 m. On the other hand, the minor loss pwe2 related to the
sudden contraction reads (White, 2011)
From Eq. (5), the pressure drop in the region with no sand pns 
is expressed in terms of three contributions: pwi , po and pwe . d2 8
pwe2 = 0.42 1− 2
Q2 (A.7)
DF0  2 d4
A.1.1. pwi – water inlet
Three contributions to the pressure drop from the
A.2. Porous media
inlet manometer to the top of the sand bed pwi
(=pwi1 + pwi2 + pwi3 ) are assumed. The first one pwi1
Based on McCabe et al. (2001), we assume that water in the
corresponds to the major losses due to friction effects in the inlet
porous media flows through n non-connected cylindrical channels
pipe of diameter d = 25.4 mm. Since the volumetric flow rate is
of length z and equivalent diameter Deq . Two geometrical con-
of the order of Q ≈ 10−3 m3 s−1 , the Reynolds number (computed
using Eq. (A.2)) at the inlet conduit is Re ≈ 5 × 104 . This means that  to this2 model.
ditions apply  First, the water volume inside the
channels Vc = nzDeq /4 must be equal to the water volume in the
the flow is turbulent and, since, 4000 < Re < 105 , the friction factor f  
may be expressed in terms of the Blausius formula f = 0.316 Re−1/4 . porous media εVF where ε is the sand porosity and VF = zDF2 /4
Therefore, Eq. (A.1) for this particular case reads is the volume of the filter occupied by sand. At a given vertical slab
of thickness dz, the condition Vc = εVF reads
pwi1 = 0.241Ld−19/4 1/4 3/4 Q 7/4 (A.4) 2
Deq DF2
ndz = ε dz (A.8)
where L = 95 mm is the length of the inlet pipe from the manometer 4 4
to the entry into the filter.
The second geometrical condition equals the wetted surface
The second contribution to pwi corresponds to the sudden
area of the n channels Sc (=Deq nz) with the sand surface area Ss .
expansion from the inlet pipe of diameter d = 25.4 mm to the fil-
Note that Ss = rp Vs with rp = Sp /Vp the surface to volume ratio of a
ter of diameter DF0 = 200 mm. Following White (2011), the minor
2
2 sand particle and Vs (=(1 − ε)VF ) the volume of sand. Since sand
loss coefficient reads K = 1 − d2 /DF0 that substituted into Eq. particles are not totally spherical, rp = Sp /Vp = 6/(Dp ) with Dp the
(A.3) gives particle diameter. At a given vertical slab of thickness dz, this second
 2 geometrical condition reads
d2 8
pwi2 = 1− Q2 (A.5) DF2
2
DF0  2 d4 6
Deq ndz = (1 − ε) dz (A.9)
Dp 4
Finally, the third contribution to pwi , namely pwi3 , takes into
account the friction loss of the water-only region inside the filter From Eq. (A.8), the number of channels n satisfies
above the porous media. Since the Reynolds number for the water-  2
DF
only region inside the filter is ReF = 6366 for Q ≈ 10−3 m3 s−1 , pwi3 n=ε (A.10)
follows Eq. (A.4) with DF0 instead of d and L equal to the height of Deq
water within the filter above the sand bed. We point out that, for and, from Eq. (A.9), the equivalent diameter is:
simplicity, we also use Eq. (A.1) for Q ≈ 0.5 × 10−3 m3 s−1 although
in this case Re < 4000. 2 ε
Deq = Dp (A.11)
3 (1 − ε)
A.1.2. po – orifice The Reynolds number (Eq. (A.2)) in the channels is
After the sand, the water flows into the nozzle and goes Rec = vc Deq / where the channel mean fluid velocity follows
through the orifice of diameter Do = 23 mm and length h = 10 mm. vc = Qc /(Deq2 /4) assuming a total volumetric flow rate Q equally
Since h/DO > 0.125, the orifice may be considered as a short tube divided into the n channels (Qc = Q/n). This implies that vc = v/ε
(Jankowski et al., 2008) and the pressure drop po through this where v is the mean fluid velocity in sand-free conditions (i.e.,
element may be obtained as a series representation of major and v = Q/(D2 /4)).
minor losses (Jankowski et al., 2008) With the above, the Reynolds number in the channels Rec =
  2/3(1 − ε)−1 Dp DF−1 ReF . Since Dp /DF  1, Rec  ReF and the flow
h 8Q 4 1 − ˇ4 8Q 4
po = f + (A.6) regime within the sand bed is clearly laminar. Therefore, the pres-
4
Do 2 Do Cd2 2 sure drop due to friction effects pf in a channel follows Poiseuille
equation:
where for the major loss, f is the friction factor, and f h/Do equals
to 64 h/(ReDo ) + 36/Re + 1.8 Re(Re − 8)/(256 + Re2 ), being Re the vc  4Q
pf = 32 dz = 32 dz (A.12)
Reynolds number in the orifice (Jankowski et al., 2008), and for the 2
Deq 4 n
Deq
minor loss, ˇ = Do /DF0 = 0.115 and a discharge coefficient Cd = 1.6
(White, 2011). that from Eqs. (A.10) and (A.11) reads,

 4Q (1 − ε)2
A.1.3. pwe – water exit pf = 72 dz (A.13)
2 Dp2 DF2 ε3
Finally, the pressure drop from the exit of the orifice to the
manometer located at the exit pipe pwe is divided into three Minor losses pm are also taken into account by assuming that
main contributions: pwe1 corresponding to the friction (major) these n channels through which water flows in the porous media
G. Arbat et al. / Agricultural Water Management 126 (2013) 64–74 73

are not straight but made by m sudden expansions and contrac- expansion (Ka ≈ 1.42; White, 2011). However, this discrepancy is of
tions due to the sand shape. We assume that these minor losses minor importance in region I, where the flow is laminar and minor
are found at regular intervals equal to the sand particle effective losses computed with Eq. (A.17) accounts only for a 10% of psI .
diameter Dp , so for a channel of length z, m = z/(Dp ) and, at a slab
of thickness dz, Eq. (A.3) reads A.2.2. Region II
In region II, psII is also the sum of major plus minor losses. Also
 2 dz  16Q 2
dz
pm = Ka v = Ka (A.14) as in region I, the major losses pfsII are obtained after integrating
2 c Dp 4 n2 D
2 2 Deq p
Eq. (A.12) (or, equivalently, Eq. (A.13)) from z = Lu (the end of region
where Ka is the minor loss coefficient for sudden expansion and I) to z = Ln (the end of region II located at the nozzle entry). However,
contraction. as discussed above, the number of channels n available in region II
Eq. (A.14) expressed in terms of the total volumetric flow rate Q decreases as the flow approaches the slots of the nozzle so n is a
is function of z. Therefore, the major losses pfsII from Eq. (A.12) are

Ln
 16Q 2 dz Q 1
pm = Ka (A.15) pfsII = 128 dz (A.18)
2 2 D4 ε2 Dp 4
Deq n
F Lu
The flow through the porous media is not expected to be We point out that from Eq. (A.10), a decrease in n is equiva-
uniform, especially near the nozzle where streamlines converge lent to reduce the effective filter diameter DF . For simplicity, we
toward its slots. In our idealized model this effect implies a reduc- assume a linear decrease of the equivalent filter diameter DF on z
tion of the number of channels n available or, equivalently from with DF = DF0 at z = Lu and DF = DF1 at z = Ln . Thus, substituting Eq.
Eq. (A.10), a reduction on the effective filter diameter DF (Fig. 3). (A.10) into Eq. (A.18) and integrating leads to
Indeed, such a reduction in the effective filter diameter is simi-
lar to the behavior observed for the streamlines in a conduit as it  4Q (1 − ε)2 DF0
pfsII = 72 (Ln − Lu ) (A.19)
approaches an orifice (e.g., Jeong and Choi, 2005), where recircula- 2 Dp2 2
DF0 ε3 DF1
tion zones, not involved in the main stream, arise attached to the
inner corners. Therefore, we divide the sand bed into two regions. which, per unit length, is nothing but Eq. (A.16) multiplied by
Region I ranges from the top of the sand bed (at z = 0) to a distance DF0 /DF1 . For the nozzle of the scaled filter, the effective filter
very close to the nozzle (at z = Lu ) and is characterized by a uniform diameter DF1 at z = Ln corresponds to the equivalent diame-
flow with DF constant (=0.2 m). Region II ranges from the end of ter for the open area of the nozzle Sn (=607.5 × 10−6 m) being
region I (at z = Lu ) to the end of the sand column at the nozzle slots DF1 = (4Sn /)1/2 = 0.028 m, so DF0 /DF1 ≈ 7.2.
with non-uniform flow (i.e., with DF variable). Minor losses pmsII are obtained after integrating Eq. (A.15)
Following Eq. (6), the total pressure drop in the porous media from z = Lu to z = Ln ,
(ps ) is the sum of the pressure losses in region I (psI ), plus those

Ln
 16Q 2 1
in region II (psII ), being calculated as follows. In order that the pmsII = Kb 4 D
dz (A.20)
2 2 Deq p Lu
n2
model can be applied to make predictions, the length of regions I
and II must be previously known. For this purpose, and consistently where Kb is the minor loss coefficient for a T junction.
with the experimental results obtained with different media bed Substituting Eq. (A.10) into Eq. (A.20) and assuming a linear
depths, as previously exposed in Section 4.1, the boundary of region decrease for the effective filter diameter as in Eq. (A.19), Eq. (A.20)
I was limited at the top of the nozzle. gives
 2

A.2.1. Region I  16Q 2 (Ln − Lu ) 1 DF0 DF0 DF0
pmsII = Kb + +1 (A.21)
In region I, psI is the sum of major pfsI plus minor losses 2 2 D4 ε2 Dp 3 DF1 2
DF1 DF1
F0
pmsI . The major losses term pfsI is obtained after integrating Eq.
that reduces to Eq. (A.17) for DF0 = DF1 , as it should.
(A.13) from z = 0 to z = Lu , giving
In addition to major losses pfsII and minor losses pmsII , we
 4Q (1 − ε)2 have also a third contribution to the pressure drop psII3 in region
pfsI = 72 Lu (A.16) II that consists of local losses related to the connections needed
2 Dp2 DF2 ε3
to reduce the number of water channels from n0 at z = Lu (i.e., at
where  is the tortuosity that accounts for the fact that channels DF = DF0 ) to n1 at the nozzle slot z = Ln (i.e., at DF = DF1 ). For simplicity,
are not totally straight. For a value of  = 150/72 ≈ 2, Eq. (A.16) we assume that connections occur by pairs at q regular intervals,
corresponds to the Blake–Kozeny equation for describing the pres- so n0 /n1 = 2q . From the values above, q ≈ 6 and, from Eq. (A.3),
sure losses of a flow in a porous media at low Reynolds numbers
(MacDonald et al., 1991).  16Q 2 q 1
psII3 = Kb (A.22)
Minor losses pmsI follow by integrating Eq. (A.13) from z = 0 to 2 ε2 2 i=1 D4
Fi
z = Lu ,
with DFi the value of the effective filter diameter at level
 16Q 2 Lu zi = Lu + (Ln − Lu )i/q for i = 1, 2, . . ., q and Kb is on the order of 0.5
pmsI = Ka (A.17)
2 2 D4 ε2 Dp (White, 2011).
F
The previous derivation has applied numerical values corre-
Note that Eq. (A.17) for Ka = 3.5(1 − ε)/ε is nothing but the sponding to the scaled filter. It is straightforward to obtain it with
Burke–Plummer equation for predicting the pressure drop in a data of any commercial filter with known geometrical slots, as it has
porous media at high Reynolds numbers (McCabe et al., 2001). been done in order to compare with experimental data in Section
Indeed, Eqs. (A.16) and (A.17) with  = 150/72 and Ka = 3.5(1 − ε)/ε 4.3.
correspond to the Ergun equation for estimating the pressure losses
in porous media (Macdonald et al., 1979).
References
For values corresponding to the types of sand here used,
Ka = 3.5(1 − ε)/ε ≈ 4.7, which is approximately three times larger Arbat, G., Pujol, T., Puig-Bargués, J., Duran-Ros, M., Barragán, J., Montoro, L., Ramírez
than the maximum value expected for a sudden contraction and de Cartagena, F., 2011. Using computational fluid dynamics to predict head
74 G. Arbat et al. / Agricultural Water Management 126 (2013) 64–74

losses in the auxiliary elements of microirrigation sand filter. Transactions of Jeong, J.T., Choi, S.R., 2005. Axisymmetric stokes flow through a circular orifice in a
the ASABE 54 (4), 1367–1376. tube. Physics of Fluids 17 (5), 1–5.
Ayars, J.E., Bucks, D.A., Lamm, F.R., Nakayama, F.S., 2007. Introduction. In: Lamm, F.R., Leva, M., 1959. Fluidization. McGraw-Hill, New York.
Ayars, J.E., Nakayama, F.S. (Eds.), Microirrigation for Crop Production (Design, Lewis, W.K., Gilliland, E.R., Bauer, W.C., 1949. Industrial and Engineering Chemistry
Operation and Management). Elsevier, Amsterdam, pp. 1–26. 41 (6), 1104–1117.
Burt, C.M., 1996. Media tank filtration for drip and microirrigation. International MacDonald, M.J., Cu, C.F., Guilloit, P.P., Ng, K.M., 1991. A generalized
Water and Irrigation 16 (2), 1–4. Blake–Kozeny equation for multisized spherical particles. AIChE Journal 37 (10),
Burt, C.M., Styles, S.W., 2007. Filtration (solids removal). In: Burt, C.M., Styles, 1583–1588.
S.W. (Eds.), Drip and Microirrigation Design and Management for Tress, Vines McCabe, W.L., Schmith, J.C., Harriot, P., 2001. Unit Operations of Chemical Engineer-
and Field Crops Practice Plus Theory. ITRC, BRAE and CalPoly, San Luis Obispo, ing. McGraw-Hill, New York.
California, pp. 175–222. Macdonald, I.F., ElSayed, M.S., Mow, K., Dullien, F.A.L., 1979. Flow through porous
Burt, C.M., 2010. Hydraulics of Commercial Sand Media Filter Tanks Used for Agricul- media: the Ergun equation revisited. Industrial and Engineering Chemistry Fun-
tural Drip Irrigation. ITRC Report No. R 10001. Irrigation Training and Research damentals 18 (3), 199–208.
Center, San Luis Obispo, California. Mesquita, M., Testezlaf, R., Ramirez, J.C.S., 2012. The effect of media bed characteris-
Capra, A., Scicolone, B., 2004. Emitter and filter tests for wastewater reuse by drip tics and internal auxiliary elements on sand filter head loss. Agricultural Water
irrigation. Agricultural Water Management 68 (2), 135–149. Management 115, 178–185.
Chang, M., Trussell, R.R., Guzman, V., Martinez, J., Delaney, C., 1999. Laboratory Nemec, D., Levec, J., 2005. Flow through packed bed reactors: 1. Single-phase flow.
studies on the clean bed head loss of filter media. Aqua 48 (4), 137–145. Chemical Engineering Science 60 (24), 6947–6957.
Choi, Y.S., Kim, S.J., Kim, D., 2008. A semi-empirical correlation for pressure Reddy, R.K., Joshi, J.B., 2010. CFD modeling of pressure drop and drag coefficient in
drop in packed beds of spherical particles. Transport in Porous Media 75 (2), fixed beds: wall effects. Particuology 8 (1), 37–43.
133–149. Riefler, N., Heiland, M., Räbiger, N., Fritsching, U., 2012. Pressure loss and wall shear
Clark, G.A., Haman, D.Z., Prochaska, J.F., Yitayew, M., 2007. General system design stress in flow through confined sphere packings. Chemical Engineering Science
principles. In: Lamm, F.R., Ayars, J.E., Nakayama, F.S. (Eds.), Microirrigation for 69 (1), 129–137.
Crop Production (Design, Operation And Management). Elsevier, Amsterdam, Soyer, E., Akgiray, A., 2009. A new simple equation for the prediction of filter expan-
pp. 161–220. sion during backwashing. Journal of Water Supply: Research and Technology –
Duran-Ros, M., Puig-Bargués, J., Arbat, G., Barragán, J., Ramírez de Cartagena, F., AQUA 58 (5), 336–345.
2009. Effect of filter, emitter and location on clogging when using effluents. Testezlaf, R., 2008. Filtros de areia aplicados à irrigaç ão localizada: teoria e prática.
Agricultural Water Management 96 (1), 67–79. Revista Engenharia Agrícola 28 (3), 604–613.
Formisani, B., Girimonte, R., Mancuso, L., 1998. Analysis of the fluidisation pro- Torrez Irigoyen, R.M., Giner, S.A., 2011. Fluidisation velocities during
cess of particle beds at high temperature. Chemical Engineering Science 53 (5), processing of whole soybean snack. Journal of Food Engineering 107 (1),
951–961. 90–98.
Foumeny, E.A., Benyahia, F., Castro, J.A.A., Moallemi, H.A., Roshani, S., 1993. Correla- Trooien, T.P., Hills, D.J., 2007. Application of biological effluent. In: Lamm, F.R.,
tions of pressure drop in packed beds taking into account the effect of confining Ayars, J.E., Nakayama, F.S. (Eds.), Microirrigation for Crop Production (Design,
wall. International Journal of Heat and Mass Transfer 36 (2), 536–540. Operation and Management). Elsevier, Amsterdam, pp. 329–356.
Haman, D.Z., Smajstrla, A.G., Zazueta, F.S., 1994. Media Filters for Trickle Irrigation U.S. Department of Defense, 2003. Unified Facilities Criteria: Army filtration of liq-
in Florida. AE-57, Agricultural and Biological Engineering Department, Florida uids, UFC 3-280-04, Available on line at: http://www.wbdg.org/ccb/DOD/UFC/
Cooperative Extension Service, Institute of food and Agricultural Sciences, Uni- ufc 3 280 04.pdf (29.05.12).
versity of Florida, Gainesville, FL. White, F.M., 2011. Viscous Flow in Ducts, Fluid Mechanics, 7th ed. McGraw-Hill,
Jankowski, T.A., Schmierer, E.N., Prenger, F.C., Ashworth, S.P., 2008. A series pressure Boston, MA, pp. 355–394.
drop representation for flow through orifice tubes. Journal of Fluids Engineering Winterberg, M., Tsotsas, E., 2000. Impact of tube-to-particle-diameter ratio on pres-
130 (5), 1–7. sure drop in packed beds. AIChE Journal 46 (5), 1084–1088.

You might also like