You are on page 1of 9

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

net/publication/265206393

An experimental and analytical study to calculate pressure drop in sand filters


taking into account the effect of the auxiliary elements

Conference Paper · July 2014

CITATIONS READS

0 2,182

7 authors, including:

G. Arbat Jaume Puig-Bargués


Universitat de Girona Universitat de Girona
82 PUBLICATIONS   1,310 CITATIONS    92 PUBLICATIONS   1,066 CITATIONS   

SEE PROFILE SEE PROFILE

Miquel Duran-Ros Toni Pujol


Universitat de Girona Universitat de Girona
52 PUBLICATIONS   677 CITATIONS    93 PUBLICATIONS   1,424 CITATIONS   

SEE PROFILE SEE PROFILE

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

Adoption of the deficit irrigation strategies to improve the water use efficiency of under-greenhouse and open-field vegetable crops View project

Drip irrigation design View project

All content following this page was uploaded by Jaume Puig-Bargués on 01 September 2014.

The user has requested enhancement of the downloaded file.


Ref: C0405

An experimental and analytical study to calculate pressure


drop in sand filters taking into account the effect of the
auxiliary elements

Gerard Arbat, Jaume Puig-Bargués, Miquel Duran-Ros and Francisco Ramírez de


Cartagena, Department of Chemical and Agricultural Engineering and Technology,
University of Girona, C. de Maria Aurèlia Capmany, 61, 17071 Girona, Catalonia, Spain
Toni Pujol and Lino Montoro, Department of Mechanical Engineering and Industrial
Construction, University of Girona, C. de Maria Aurèlia Capmany, 61, 17071 Girona,
Catalonia, Spain
Javier Barragán, Department of Agricultural and Forestry Engineering, University of Lleida,
Av. de l'Alcalde Rovira Roure, 191, 25198 Lleida, Catalonia, Spain

Abstract

Sand filters are frequently used in micro-irrigation especially when water contains large
amounts of organic contaminants. This type of filter has the advantage of its simplicity and
that the main filtration mechanism is based on depth filtration, giving an additional removal
capacity in comparison with screen or disc filters, which essentially work by surface filtration.
As the prediction of the head losses is of practical interest for the design of the micro-
irrigation systems, the main objective of this paper is to present and validate an analytical
model to compute the head losses produced by a sand filter, which takes into account the
effect of the filter underdrain.
It has been shown that the classical Ergun equation works well for predicting the pressure
drop within a sand bed when it is sufficiently far from the underdrain, but it fails in the region
immediately next to this element. To overcome this problem, a model based on a set of
connected channels of the same diameter and progressive reduction of its number as the
flow approaches the nozzles is presented.
An experimental study with a scaled commercial sand filter was conducted with two different
media bed weights (3.86 and 13.53 kg) and two sand grain size ranges (0.63-0.75 and 0.75-
0.84 mm) in order to compare the reliability of the proposed analytical model.
The experimental results were compared with those yielded by the proposed analytical
model, as well as with the Ergun equation. Statistical indexes show that the proposed model
improved the results of the Ergun equation. Thus, the root mean square error (RMSE)
comparing the results of the proposed model with experimental data ranged from 2.85 to
7.16 kPa, while when the results of Ergun equation were compared with the experimental
data the RMSE ranged from 11.02 to 17.94 kPa.
It can be concluded that the new analytical model 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 the pressure
drop in sand filters.

Keywords: drip irrigation; head losses; sand media filter; packed bed; mathematical
modeling

Proceedings International Conference of Agricultural Engineering, Zurich, 06-10.07.2014 – www.eurageng.eu 1/8


1. Introduction

Water quality is a major concern in the management of drip irrigation systems as the solids
contained in the irrigation water can clog the emitters. Thus, filtration is essential for the
successful operation of micro-irrigation systems. Media filters are commonly used for
protecting micro-irrigation systems and are specially recommended when large amount of
organic contaminants are present (Duran-Ros, Puig-Bargués, Arbat, Barragán, & Ramírez
de Cartagena, 2009)
Few studies have analyzed the effect of the auxiliary components of media filters on the
head loss, which is related with water and energy consumption as well as filter efficiency
(Burt, 2010; Mesquita, Testezlaf, & Ramírez, 2012).
The Ergun equation has been extensively used to predict head losses in the filter media
(Macdonald, El-Sayed, Mow, & Dullien, 1979) since there is a general consensus on its
accuracy. However, its application is strictly valid in the limit of an infinitely extended packed
bed (Nemec & Levec, 2005). The boundary condition at the sides of the packed bed, known
as the wall effect, has been studied by numerous investigations. These studies demonstrate
that when the bed diameter is not significantly larger than the particle diameter the friction
produced by the wall of the tank is not negligible, invalidating the applicability of Ergun
equation (Foumeny, Benyahia, Castro, Moallemi, & Roshani, 1993). The boundary condition
near the filter underdrain also violates the assumption of an infinitely extended packed bed
made by the Ergun equation. The capillary model based on the hydraulic radius concept, the
physical model in which is based the Ergun equation, assumes that porous media consists
of a set of parallel identical channels (Nemec & Levec, 2005). This simplified model of the
flow within the porous media may fail to represent the actual flow occurring near the
underdrain elements since the streamlines must converge near the passing area through the
slots of the underdrains. Indeed, different empirical studies have shown that an important
part of the total pressure drop in commercial sand filters is due to the underdrain system
(Burt, 2010; Mesquita et al., 2012). From a previous computational study Arbat et al. (2011)
showed that a great part of the head losses in a commercial sand filter are produced in a tiny
area inside the packed bed close to the underdrain. Their results reveal the importance of
properly including the underdrain system when developing models for predicting head losses
in packed beds. Nevertheless, it must be noticed that the application of computational fluid
dynamic (CFD) software to simulate the pressure drop in a sand filter has strong limitations
since it requires an enormous computational force to discretizate the complex geometries of
the sand filter.
Arbat et al. (2013) developed an analytical solution that can be applied to compute the head
losses produced by the underdrain of a porous media filter. The main objective of this paper
is to quantify the degree of agreement, using statistics of comparison, of the pressure drop
predicted with the analytical model presented in Arbat et al. (2013) as well as with Ergun’s
equation with experimental results obtained in a commercial scaled sand filter in order to
analyze the convenience to use the new equation to predict head losses in sand filter with
nozzle-type underdrains.

2. Materials and methods

2.1. Experimental procedures

A scaled sand filter was constructed based on a commercial sand filter (Regaber, Parets del
Vallès, Spain) of 500 mm internal diameter and with a filtration surface of 1963 cm2. The
dimensions of the scaled filter were chosen such that the Reynolds number corresponding to
the circular tank of the filter was similar to that of the commercial filter. Since the commercial
filter had 12 nozzles and the scaled filter 1 nozzle, the scaling factor for all lengths was
approximately 1/12 1/3. Therefore the volume flowing through each nozzle was the same in

Proceedings International Conference of Agricultural Engineering, Zurich, 06-10.07.2014 – www.eurageng.eu 2/8


both filters. A detailed description of the experimental filter and experiment can be found in
Arbat et al. (2013).
The relationship between superficial velocity (V0) and the pressure difference from the filter
inlet and outlet (Dp=pin-pout) was obtained at the filtration regime for different filter media
configurations. Each experiment was repeated 3 times to ensure repeatability of the data.
Two different sand particle size ranges (0.63-0.75 mm and 0.75-0.84 mm) and three different
sand bed weights (or, equivalently, heights) were tested at the filtration regime (Table 1).
The main characteristics of the sands are shown in Table 2.These tests were the same
showed by Arbat et al. (2013).
The underdrain configuration was a nozzle of identical characteristics than that of the
commercial sand filter (Regaber, Parets del Vallès, Spain). The nozzle had 45 slots 0.45 mm
wide and 30 mm long with a total passing area of 607.5 mm2. The passing area through the
orifice was 415 mm2. As the scaled filter has only one nozzle and 200 mm internal diameter,
each nozzle served an area of 314 cm2.
The experiments with 13.530 kg of sand represent approximately the 1/12 of the volume of
sand in the commercial sand filter that was taken as a reference and, therefore, correspond
to the nominal column height of the scaled filter. The tests with 1.547 kg of sand with the
nozzle gave a media bed depth of 20 mm over the mid height of the nozzle, which
represents a very short sand column in relationship with the experiments with 13.530 kg of
sand. The experiments with the short sand column were designed to analyze the pressure
drop generated in a tiny area over the underdrain, and to compare their results with those
obtained in the analytical model. Additionally, experiments with an intermediate media bed
depth corresponding to 3.860 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 verifying the pressure losses calculated with the Ergun equation in a 0.21 m
long sand column without the influence of the underdrain system. Similarly, the difference of
the pressure losses obtained with 13.530 and 1.547 kg of sand would serve to know the
pressure losses in a sand column 0.26 m long.

2.2. Nozzle type underdrain analytical model

The model presented in Arbat et al. (2013) is based on a series pressure drop
representation from point 1 located at the inlet of the filter to point 2 located at the exit filter.
The different regions of the filter are shown in Fig. 1. Thus, the measured total pressure loss
∆p12 follows:
∆p12 = ∆ps + ∆pns (4)
where ∆ps and ∆pns are the pressure losses through the porous media and the non-porous
media, respectively.
∆pns can be divided into three main contributions
∆pns = ∆pwi +∆po+∆pwe (5)
where ∆pwi is the pressure drop from point 1 to the top of the sand column (water inlet
region), ∆po is the pressure drop through the nozzle orifice and ∆pwe is the pressure drop
from the exit of the nozzle orifice to point 2 (water exit region). Note that these terms may
include losses from auxiliary elements (e.g., backflush valves).
On the other hand, the pressure loss through the sand bed ∆ps is divided into two terms
∆ps = ∆psI + ∆psII (6)
where ∆psI,II correspond to the pressure losses in regions I and II shown in Fig. 1 In region I,
the flow is uniform through the sand bed and the pressure drop follows the Ergun equation.
In region II, however, the flow is not uniform due to the influence of the nozzle. Since the
drain opening area of the nozzle is smaller than the filter cross-sectional area, the fluid
velocity within the sand increases as it approaches the nozzle. This leads to large values of
pressure losses in a very small zone close to the nozzle. In this region, the Ergun equation
does not provide reasonable predictions. This effect was already noted by means of detailed

Proceedings International Conference of Agricultural Engineering, Zurich, 06-10.07.2014 – www.eurageng.eu 3/8


CFD simulations of a commercial filter in Arbat et al. (2011). A new expression for
determining the contribution of the ∆psII term was developed by Arbat et al. (2013) based on
a generalization of the idealized representation of the porous media. The porous media is
assumed of being formed by inter-connected cylindrical channels of equivalent diameter Deq
through which water flows. Analytical expressions for all the terms shown in equations (4-6)
can be seen in Arbat et al. (2013). In the present work we compare the results obtained with
these equations with the ones obtained in the experimental tests described in section 2.1,
and also with the ones obtained with the classical Ergun’s equation.

2.3. Statistics of comparison between the head loss determined experimentally and
predicted by the models

In order to determine the agreement of the analytical model, the root mean square error,
RMSE and the Willmott’s index of agreement, d (Willmott, 1982) were computed for the
pressure drop at the different velocity regimes during the experimental tests.
The statistics were computed as follows:
n

 P  O 
2
i i
RMSE  i 1
(7)
n
n 2

 Pi  Oi 
d 1 i 1
2
(8)
n
 _ _

  Pi  O  Oi  O 
i 1  
being Oi , the pressure drop obtained from the experimental tests at a particular superficial
velocity, Pi the predicted pressure drop at the same regime, O the average pressure drop at
the different tested regimes. Therefore, the units of the RMSE are kPa, d has no dimensions
and its upper limit is 1, which indicates a perfect agreement between the analytical solution
and the experimental tests.

3. Results and Discussion

3.1. Proportion of head losses in the different parts of the filter

The pressure drop in a sand bed column 0.21 m long filled with sand 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 column 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. Using this procedure, the pressure losses in the auxiliary elements of the
fitler, as well as those produced in the sand bed closest to the nozzle cancelled out.
Therefore, the pressure drop can be compared with the one predicted with Ergun equation.
Figure 2 shows that the pressure drop predicted with the Ergun equation follows a similar
trend than that obtained in the experiments although it slightly over-predicts the pressure
drop for the three cases. In both experimental and Ergun's predictions, the lower pressure
drop was produced by the sand column of 0.21 m height and the bigger grain size (0.75-0.84
mm), followed by the sand column of the same height filled with grain sizes from 0.63-0.75

Proceedings International Conference of Agricultural Engineering, Zurich, 06-10.07.2014 – www.eurageng.eu 4/8


mm. The greatest pressure drop was produced by the longer sand bed (L=0.26 m) filled with
0.63-0.75 mm sand size. As generally admitted, Ergun equation is suitable to predict
pressure drop in infinitely extended packed bed, especially when the porosity range is rather
narrow (0.35<ε<0.55), the bed is made up of similar sized particles and the flow rates are
moderate (Nemec & Levec, 2005), which was the case of the sand filter experiments
presented here.
Applying the model developed by Arbat et al. (2013), the sand bed was responsible of most
of the pressure losses, roughly 90%. Figure 3 shows that in the experiments with 13.530 kg
of sand approximately the 10% of the total pressure losses were produced in the region
ranging from the top of the sand bed to a short distance from the nozzle (Region I) and 70%
in a very tiny region at the bottom of the sand column, close to the nozzle slots (Region II). In
the experiments with 1.547 and 3.860 kg of sand, approximately 35 and 50% of the total
pressure drop was produced in regions I and II respectively.
The extension of region II, where most of the pressure losses are produced, was initially
unknown but the experiment with 1.54 kg of sand, where the sand column was as short as
2.5 mm over the top of the nozzle, would confirm that it is restricted to a very tiny area at the
bottom of the sand column. Indeed, the experimental results corresponding to a 0.26 m sand
column (Fig. 2), obtained by subtracting the measured values of the 13.530 kg case with
those of the 1.547 kg of sand, did not show any abrupt increase of the pressure drop
compared with the one obtained with a 0.21 m sand column and followed the good
agreement with Ergun equation. Therefore it was experimentally confirmed that the length of
region II is less than the 2.5 mm over the top of the nozzle.

3.2. Statistics of comparison of the predictions of the different models with the
experimental ones

The degree of agreement of the analytical model and Ergun’s equation compared to the
experimental results was evaluated using the Willmott’s index of agreement (d) and the
RMSE (Table 3). The results show that when the proposed analytical model, that takes the
effect of the underdrain in the flow inside the porous medium (Region II), the difference
between the model prediction and the experimental results was reduced greatly in all the
cases compared with the results obtained with Ergun’s equation. It must be noticed that
when Ergun’s equation was used to compute the pressure drop, the pressure losses
produced by the auxiliary elements were taken into account and computed in the same way
than in the proposed analytical solution. The results of the Willmott’s index of agreement (d)
also indicate an improvement of the predictions of the proposed model, being this index
greater than 0.90 in all the comparisons with the proposed model while it ranged 0.60 to 0.80
when Ergun’s equation was used. Therefore, the proposed model showed a closer
prediction of the head losses than that of the Ergun’s equation in the entire set of
experimental tests.

4. Conclusions

The Ergun equation is well suited to predict the pressure drop produced by an infinitely long
packed bed. Nevertheless, the idealization of parallel channels, which is the physical model
that supports Ergun equation, does not match the hydraulic behavior in the lower part of a
filter since there exists an underdrain element (nozzle). The experiments carried out
confirmed that the flow through the porous media is not uniform near the nozzle where
streamlines converge towards its slots. In the idealized analytical model presented in Arbat
et al. (2013), this effect was implemented by a reduction of the number of channels available
as well as a reduction of the effective filter diameter. The statistics of comparison showed
that the proposed model clearly improves Ergun’s predictions, since remarkably fits
experimental data. The analytical model presented in this paper could be applied to predict

Proceedings International Conference of Agricultural Engineering, Zurich, 06-10.07.2014 – www.eurageng.eu 5/8


the pressure drop produced by commercial sand filters using nozzle-type underdrains, which
are currently used in the filtration units used in micro-irrigation systems.

5. Acknowledgements

The authors would like to express their gratitude to the Spanish Ministry of Economy and
Competitiveness for its financial support of this study through grants CGL2012-31180, and
FIS-2012-31307, to the Autonomous Government of Catalonia for grant 2009-SGR-374. The
authors would also like to thank Sergi Saus, Jordi Vicens, and Regaber for their help in
carrying out this investigation.

Figure 1: Sand bed regions considered to compute pressure drop in the model of Arbat et al. (2013).

Figure 2: Experimental and Ergun's predictions of the pressure drop in the sand bed (∆p, kPa) for
different superficial velocities (m h-1).

Proceedings International Conference of Agricultural Engineering, Zurich, 06-10.07.2014 – www.eurageng.eu 6/8


Figure 3: Percentage of the total pressure drop produced in the different regions considered to
compute pressure drop in the model of Arbat et al. (2013)

Table 1: Different filter configurations tested in the experiments with the scaled sand filter.

Sand size Sand weight Media bed

(mm) (kg) depth* (mm)

0.63-0.75 1.547* 20.0

0.63-0.75 3.860 70.1

0.63-0.75 13.530 279.5

0.75-0.84 3.860 70.4

0.75-0.84 13.530 280.4

* Sand bed depths over the mid height of the nozzle.

Table 2: Main characteristics of the two silica sand sizes used in the study.

Grain size Average sand Bulk density Particle density Porosity

ranges diameter (mm) (kg m -3) (kg m-3) (%)

(mm)

0.63-0.75 0.690 1470 (16) 2556 (0) 42.50 (0.00)

0.75-0.84 0.795 1462 (39) 2507 (6) 41.67 (1.44)

The number in brackets is the standard deviation of three different samples.

Proceedings International Conference of Agricultural Engineering, Zurich, 06-10.07.2014 – www.eurageng.eu 7/8


Table 3: Statistics of comparison of proposed analytical model with the experimental results, and of

Ergun equation with the experimental results.

RMSE Willmott’s index of agreement

(kPa) (d), (adimensional)


Grain sizes Sand weight
Proposed
(mm) (kg) Ergun’s
Proposed Ergun’s
analytical
analytical
equation*
solution equation*
solution

0.63-0.75 1.547*
2.80 16.60 0.98 0.59
0.63-0.75 3.860
2.85 17.94 0.98 0.59
0.63-0.75 13.530
4.11 14.58 0.93 0.73
0.75-0.84 3.860
7.16 11.50 0.90 0.69
0.75-0.84 13.530
4.03 11.02 0.98 0.79
(*) When Ergun’s equation was applied the pressure losses in the auxiliary elements were computed in the same way than in

the proposed analytical solution.

6. References

Arbat, G., Pujol, T., Puig-Bargués, J., Duran-Ros, M., Barragán, J., Montoro, L., & Ramírez de
Cartagena, F. (2011). Using computational fluid dynamics to predict head losses in the
auxiliary elements of a microirrigation sand filter. Transactions of the ASABE, 54(4), 1367-
1376.
Arbat, G., Pujol, T., Puig-Bargués, J., Duran-Ros, M., Montoro, L., Barragán, J., & Ramírez de
Cartagena, F. (2013). An experimental and analytical study to analyze hydraulic behavior of
nozzle-type underdrains in porous media filters. Agricultural Water Management, 126, 64-74.
Burt, C. M. (2010). Hydraulics of commercial sand media filters tanks used for agricultural drip
irrigation: criteria for energy efficiency. San Luis Obispo: Irrigation Training and Research
Center.
Duran-Ros, M., Puig-Bargués, J., Arbat, G., Barragán, J., & Ramírez de Cartagena, F. (2009). Effect
of filter, emitter and location on clogging when using effluents. Agricultural Water
Management, 96(1), 67-79.
Foumeny, E., Benyahia, F., Castro, J., Moallemi, H., & Roshani, S. (1993). Correlations of pressure
drop in packed beds taking into account the effect of confining wall. International Journal of
Heat and Mass Transfer, 36(2), 536-540.
Macdonald, I., El-Sayed, M., Mow, K., & Dullien, F. (1979). Flow through porous media-the Ergun
equation revisited. Industrial & Engineering Chemistry Fundamentals, 18(3), 199-208.
Mesquita, M., Testezlaf, R., & Ramírez, J. (2012). The effect of media bed characteristics and internal
auxiliary elements on sand filter head loss. Agricultural Water Management, 115, 178-185.
Nemec, D., & Levec, J. (2005). Flow through packed bed reactors: 1. Single-phase flow. Chemical
Engineering Science, 60(24), 6947-6957.

Proceedings International Conference of Agricultural Engineering, Zurich, 06-10.07.2014 – www.eurageng.eu 8/8

View publication stats

You might also like