Sunteți pe pagina 1din 8

Atmospheric Research 102 (2011) 335–342

Contents lists available at SciVerse ScienceDirect

Atmospheric Research
journal homepage: www.elsevier.com/locate/atmos

Aerodynamic collection efficiency of fog water collectors


Juan de Dios Rivera ⁎
Centro del Desierto de Atacama (Dep. Ing. Mecanica y Metalurgica), Pontificia Universidad Católica de Chile, Chile

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

Article history: Fog water collectors (FWC) can provide water to arid zones with persistent advection and oro-
Received 27 May 2011 graphic fog. A key feature of any FWC is the mesh used to capture fog droplets. Two relevant
Received in revised form 11 August 2011 mesh characteristics are its shade coefficient and the characteristics of the fibers used to weave
Accepted 17 August 2011 or knit the mesh. This paper develops a simple superposition model to analyze the effect of
these factors on the Aerodynamic Collection Efficiency (ACE) of FWCs. Due to the simplicity of
Keywords: the model it cannot be directly applied to actual FWC meshes, and serve only for guidance on
Fog water collector the order of magnitude of the optimum shade coefficient and the corresponding ACE. The model
Efficiency shows that there is a maximum ACE of the order of 20–24.5% for shade coefficients between 0.5
Fog
and 0.6, for the particular mesh simulated. Aerodynamic collection efficiency can be increased
Water
by making the FWC concave and improving the aerodynamics of the mesh fibers.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction South Africa (Olivier, 2002; Olivier and de Rautenbach, 2002;


Louw et al., 1998), Oman (Abdul-Wahab et al., 2010), western
The lack of fresh water resources is a crucial economic, social Mediterranean basin (Estrela et al., 2008), and Namibia
and political problem around the world that is growing as more (Shanyengana et al., 2002). In summary, fog water is a prom-
fresh water supplies are contaminated or depleted (Rijsberman, ising new source of fresh water and there is active research
2006; Smakhtin et al., 2004; Schemenauer and Joe, 1989). Fog in this subject in several countries of the world.
Water Collectors (FWC) are helping to solve this problem by pro- Most fog water collectors consist of a large screen held by
viding new fresh water resources in certain arid areas around the poles or a frame, perpendicular to the wind driven fog
world, for human or cattle consumption and also for irriga- (Fig. 1). Over time, many different designs have been used
tion purposes (Abdul-Wahab and Lea, 2008; Schemenauer (Schemenauer et al., 1988; Schemenauer and Cereceda, 1994;
and Cereceda, 1991; Schemenauer et al., 1988). Suitable per- Gischler, 1991), but at the present time the screen type collec-
sistent fog in many arid zones close to the ocean is of the ad- tor is the most common for producing significant amounts of
vection type, as clouds formed over the water are pushed by water for human use and, therefore, we focus this paper on
predominant winds towards the continent where they turn this type. The screen is typically made of two layers of Raschel
into fog when intercepted by high lands; although some- mesh, one layer of mosquito mesh, or other suitable media.
times orographic fog also contributes to fog water collection The FWC cannot collect all the liquid water contained in the
(Cereceda et al., 2008b). Several researchers have studied the fog. Indeed, due to the obstruction of wind produced by the col-
characteristics of these fogs and their potential as a water source lector, part of the flow is diverted and passes around the mesh.
for human use. In particular, they have studied areas like the pa- Additionally, from all the droplets of the fog that would go
cific coast of northern South America (Cereceda et al., 2008a, through the collector, only a fraction is actually captured. The
2008b, 2002; Larrain et al., 2002), the Canary Islands (Marzol, rest of the droplets never hit the mesh fibers, because some
2008; Marzol, 2002), Morocco (Marzol and Sánchez, 2008), flow through the openings of the mesh, or bounce back into
the airstream. Captured droplets coalesce and drain by gravity
⁎ V. Mackenna 4860-Macul, Santiago, Chile. Tel.: + 56 2 354 5886;
to the lower part of the collector where the water is received
fax: + 56 2 354 5828. by a gutter and transported to a tank. However, in this last pro-
E-mail address: jrivera@ing.pcu.cl. cess water can be re-entrained into the airflow or may drip out

0169-8095/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.atmosres.2011.08.005
336 J.D. Rivera / Atmospheric Research 102 (2011) 335–342

size distribution and mesh characteristics (Schemenauer and


Nomenclature
Joe, 1989; Schemenauer and Cereceda, 1991). Here we will
focus on one mesh characteristic that is the free flow area
Aop mesh openings area, m 2
ratio. The free flow area ratio, f, is defined as the mesh openings
A total mesh area, m 2
area, Aop, divided by the total screen area, A.
b mesh length, m
C0 mesh pressure drop coefficient
Aop
Cd collector drag coefficient f ¼ ð1Þ
A
f free flow area ratio
h mesh height, m
Many papers on fog collectors use the concept of shade
LWC Liquid water content, kg/m 3
coefficient, s, which is equal to one minus the free flow area
p pressure, kPa
ratio.
Re Reynolds number
s Shade coefficient s ¼ 1−f ð2Þ
v Velocity, m/s
ẇ ″ Water flow rate, kg/s/m 2
The shade coefficient represents the part of the collector's
area capable of capturing droplets.
We define water collection efficiency, ηcoll, as the ratio be-
Greek tween the water flow rate actually collected in the gutter per
δ fibers diameter, m ″
unit screen area, ẇcoll , and the liquid water that flows in a
η efficiency stream tube with a unit cross sectional area, unperturbed by
the collector. The latter corresponds to the liquid water flux of
the moving fog, and can be calculated as the product of unper-
Sub indices turbed wind velocity, v0, times the liquid water content, LWC.
capt capture Therefore, the collection efficiency can be expressed by
coll collection
dr draining ″
ẇcoll
op openings ηcoll ¼ : ð3Þ
v0 ⋅LWC
AC aerodynamic collection
0 unperturbed by FWC This collection efficiency can also be expressed as the
product of each one of the efficiencies of the already men-
tioned processes involved in fog water collection. As men-
of the gutter from mesh slack, wrinkles or folds. For each one of tioned before, since the FWC is an obstacle for the flow of
these processes we can define a particular efficiency, and alto- wind driven fog, a portion of the unperturbed flow passes
gether they lead to a global efficiency for the collection of fog around, and only a fraction of the fog actually reaches the col-
water. lector. We call this fraction “fog interception efficiency.” On
Water collection yields depends on many factors, the most the other hand, only part of the collector surface can catch
important being wind velocity, fog liquid water content, droplet droplets, and this corresponds to the shade coefficient s. For

Fig. 1. Typical fog water collector in Atacama Desert; a double layer of Raschel mesh, held in place by two poles, intercepts wind driven fog. The poles are supported by
steel guy cables anchored to the ground; in the lower edge of the mesh is the gutter that receives the water drained from the mesh. Photo credit: Pilar Cereceda.
J.D. Rivera / Atmospheric Research 102 (2011) 335–342 337

convenience we consider the product of the shade coefficient are not well characterized in terms of their shade coefficient
and the interception efficiency, and name it “aerodynamic and pressure drop. Indeed, Schemenauer and Joe (1989) as-
collection efficiency,” ηAC, which represents the portion of sume a shade coefficient between 0.50 and 0.76 for the dou-
droplets in the unperturbed fog that would collide with the ble layer of Raschel mesh they used in their experiment.
mesh. The aerodynamic collection efficiency (ACE) corre- There are at least three reasons for the uncertainty in s,
sponds to the maximum fraction of the unperturbed fog first, the two layers overlap in an unpredictable way, second,
that can be captured by the FWC. However, not all the drop- the mesh stretches upon installation reducing the width of
lets that would collide with the mesh are captured since the ribbons, and third, there is no guaranty that the value
some of them are deviated from the collision trajectory, fol- given by the mesh maker is true. The first two reasons have
lowing the airstreams that flow around the fibers; additional- been commented previously (Schemenauer and Joe, 1989).
ly, some of the droplets that hit the mesh may bounce back Bresci (2002) seems to be the only one that actually mea-
into the airflow. Consequently, we define capture efficiency, sured s, obtaining a value of 0.78 for a mesh used in Peru
ηcapt, as the fraction of droplets in a collision trajectory that and 0.83 for a mesh used in Canary Island. Moreover, some
are actually captured by the mesh. Finally, draining efficien- authors do not give the shade coefficient of all the meshes
cy, ηdr, is the fraction of the water captured by the mesh they use (Abdul-Wahab et al., 2010). Finally, there is no cer-
that actually reaches the gutter; some of the captured water tainty that published airflow resistance correlations apply to
is lost by re-entrainment and spill. In summary, there are all types of meshes.
losses due to fog passing around the FWC, fog passing All the efficiencies on Eq. (4) can be determined experi-
through the openings of the mesh, droplets bouncing back mentally with appropriate measurements and calculations.
into the airflow, re-entrainment of water flowing down the To determine the collection efficiency, ηcoll, it is necessary to
mesh, and water spilled before getting into the gutter. There- measure the unperturbed wind velocity and LWC, the total
fore, the collection efficiency can be calculated as water flow out of the gutter and the collector surface area,
and use Eq. (3). The ACE, ηAC, can be calculated by dividing
ηcoll ¼ ηAC ηcapt ηdr : ð4Þ the measured wind velocity very close to the mesh by the
measured unperturbed wind velocity, and multiplying it by
One problem presented by the proper selection of the mesh the measured shade coefficient. The capture efficiency, ηcapt,
is to find its best shade coefficient. Indeed, if the shade coeffi- can be determined by dividing the mesh efficiency measured
cient is too small, a reduced fraction of the fog droplets will in the manner of Schemenauer and Joe (1989) by the shade
hit the mesh and, on the other hand, if it is too large, most of coefficient. Finally, the draining efficiency, ηdr, can be calcu-
the fog carrying wind will pass around due to high resistance lated by difference using Eq. (3).
to the flow through it. It is apparent that there must be an opti-
mum shade coefficient for which a maximum water collection 2. Model for aerodynamic collection efficiency
would be obtained. This optimum is described by the maximum
ACE, assuming that changes in s would not affect the other two The flow field around a fog water collector is quite complex
efficiencies. because of the presence of the ground, which produces a large
Schemenauer and Joe (1989) had already discussed how boundary layer, the effect of the structure that supports the
mesh characteristics affect collection efficiency, but centered mesh, the partial flow of air through the mesh, and the recircu-
their discussion into ribbon width. They showed that, for the lation produced in the wake of the FWC (Bresci, 2002). Howev-
same shade coefficient, thinner ribbons increase collection er, here we will not consider those effects and will consider a
efficiency. They measured directly the product of the capture two-dimensional flow with the presence of only the mesh.
efficiency and the shade coefficient, ηcapt·s, which they This corresponds to a FWC installed very far above the ground,
named as mesh efficiency. Additionally, by comparing the outside the ground boundary layer. This simplified model will
product of the liquid water flux in the unperturbed fog and not be able to accurately predict the yield of an actual FWC,
the mesh efficiency with the water flow actually collected, but allows to demonstrate that there exists an optimum
they showed that only part of the fog reaches the collector shade coefficient and to find an order of magnitude of the max-
and the rest passes around it. This second effect is the focus imum achievable efficiency.
of our present work. The fraction of the upwind undisturbed flow that passes
In this paper we develop a simple model to show that there through the mesh depends on the balance between the large
exists an optimum shade coefficient for maximum ACE, and to scale drag on the mesh and the pressure drop of the airflow
find an order of magnitude for it. Recognizing that the real through the mesh. In other words, the difference in pressure
problem is far more complicated than our simplified model, between the windward and leeward side of the FWC has to
we think that it suffices for the time being to demonstrate match the pressure drop across the mesh. We propose to repre-
that there is an optimum and to show its order of magnitude. sent the real case, in which part of the flow passes through the
It is worthwhile to emphasize that the results of this model mesh and the rest is diverted around it, as the superposition of
cannot be directly applied to actual FWC meshes, and serve 1) a flow that passes around a solid screen and 2) one that is
only for guidance on the order of magnitude of the optimum forced to pass only through the mesh. Fig. 2 shows this super-
shade coefficient and the corresponding ACE. The real values position scheme.
for these two variables can be found by experimental measure- The principle of superposition, which has been used in
ments or numerical simulations. many areas of science and technology, states that the solution
Another important reason that prevents the direct appli- to a complex linear problem can be obtained as the summation
cation of this model to actual collectors is that meshes used of the solution to simpler problems in which the complex one
338 J.D. Rivera / Atmospheric Research 102 (2011) 335–342

v0

A0
=

A1
Mesh

v2
v1 +

b c
Fig. 2. Superposition scheme; a) represents the real case, b) represents a mesh where all the flow is forced to pass through it, and c) represents a non-permeable
screen or plate.

can be decomposed. Superposition has been used, for example, with C0 the pressure drop coefficient for the mesh. The pressure
to model heat transfer problems (Kays and Crawford, 1980; difference across the screen in case c) is
Arpaci, 1966), and complex flows in blood vessels (Cebral
ρv22
et al., 2003), porous soil (Simunek et al., 2003) and in aquifers Δp ¼ Cd ð8Þ
2
(Reilly et al., 1984; Bair and Roadcap, 1992). Superposition is
valid for potential flows, which is the case in front of the with Cd the drag coefficient for a non-porous plate. Finally, the
mesh or screen (White, 1986). In the wake there is recircula- ACE defined previously is
tion and, therefore, superposition is not strictly applicable.
However, since we seek a simple, approximate result, we will A0
ηAC ¼ s ð9Þ
assume superposition is approximately valid in this case. The A1
error in this assumption will affect the value of the drag and
pressure drop coefficients, which will give an error in the opti- and substituting Eq. (6) we get
mum free flow area ratio and the value of the maximum ACE,
but it does not affect the existence of a maximum. The error v1
ηAC ¼ s: ð10Þ
in the drag coefficient seems to be more important since we v0
found drag coefficient correlations only for non-permeable
screens. Now it only remains to express v1/v0 in terms of known var-
According to this superposition scheme, the velocity of iables. Since the pressure difference across the non-porous
the unperturbed wind, v0, equals the addition of the veloci- screen b) has to be equal to the pressure difference across the
ties of case b) and c) shown in Fig. 2. mesh c), from Eqs. (8) and (7) we get

2 2
v0 ¼ v1 þ v2 ð5Þ C0 v1 ¼ Cd v2 : ð11Þ

Assuming constant density, we can write continuity equa- Substituting v2 from Eq. (5) into (11) and arranging the
tion as terms, we obtain the quadratic
 2
v0 A0 ¼ v1 A1 ð6Þ v0 v C
−2 0 þ 1− 0 ¼ 0: ð12Þ
v1 v1 Cd
where A0 is the unperturbed area of the stream tube that passes
through the mesh and A1 is the mesh area (see Fig. 2a). Addi-
The solution to this quadratic is
tionally, the pressure difference across the mesh in case b) is
sffiffiffiffiffi
ρv 2 v0 C0
Δp ¼ C0 1 ð7Þ ¼ 1F ð13Þ
2 v1 Cd
J.D. Rivera / Atmospheric Research 102 (2011) 335–342 339

in which we have to consider only the positive sign since v0 The pressure drop coefficient for silk mesh equals the one
has to be greater than v1. Substituting this into Eq. (10) we fi- for wire mesh multiplied by a constant
nally get the result for the ACE.
  2 
s 1
ηAC ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi ð14Þ C0 ¼ 1:62 1:3ð1−f Þ þ −1
1 þ C0 =Cd  f
 s 2 
Notice that the drag coefficient Cd in this model corresponds ¼ 1:62 1:3s þ : ð16Þ
1−s
to a non-permeable screen and, therefore, is independent of
the shade coefficient, s. However, the pressure loss coefficient
C0 depends on s. Notice also that since both Cd and C0 are inde- These two equations are valid for Reynolds number based
pendent of wind velocity (for typical wind velocities, as will be on the mesh fiber diameter, Reδ, greater than 400. Since most
demonstrated in next section), LWC, and droplets size distribu- of the mesh materials used in FWC is plastic, we will consider
tion, the ACE depends only on the shade coefficient, the collec- only to the last equation, assuming that plastic fibers are
tor geometry and the mesh aerodynamics characteristic. more similar to silk than to wire. For Reδ smaller than 400,
Idel'cik gives a correction factor kRe that multiplies Eq. (16)
3. Evaluation of the aerodynamic collection efficiency shown by Table 3.
Looking at the data on Table 1, considering that the largest
In order to obtain values for the drag and pressure drop co- volume of collected water is obtained at the highest veloci-
efficient we have to characterize the flow conditions, for which ties, higher than around 4 m/s, and considering free area ra-
we need some typical data for the application we are interested tios greater than 0.3, we estimate Reδ larger than 100. From
in. For this purpose we take the data given by Schemenauer Table 3 we see that for the expected Reδ the correction is of
and Joe (1989) that is summarized in Table 1. The mesh dimen- the order of 1%. Therefore, we neglect the low Reynolds num-
sions h and b are shown in Fig. 3. ber correction for this simplified analysis.
Drag coefficients for two- and three-dimensional objects We evaluated Eq. (14) with the drag coefficients given by
have been extensively studied and are presented in most fluid Table 2 and pressure drop coefficient given by Eq. (16), and
mechanics textbooks. For fully developed turbulent flow, plotted it in Fig. 4 for a range of shade coefficient. We decided
ReN 10 4, the drag coefficient of sharp edged objects is indepen- to use the pressure drop coefficient for a silk mesh assuming
dent on Reynolds number (White, 1986; Fox and McDonald, it is closer to the typical polyethylene Raschel mesh commonly
1985). Considering the smaller dimension and the velocity used in FWC (Schemenauer and Cereceda, 1991; Schemenauer
range given in Table 1, we calculated a Reynolds number of et al., 2005). Indeed, since the Raschel mesh is made of polyeth-
over 5.6·105. Notice that even for the smallest collector, a Stan- ylene ribbons 1.0 mm wide and 0.1 mm thick, approximately,
dard Fog Collector (SFC) (Schemenauer and Cereceda, 1994), we expect it to have a larger pressure drop coefficient than a
with dimensions of one by one meter, the Reynolds number wire mesh where the fibers are smooth cylinders.
is over 1.4·105, which means fully turbulent flow. White
(1986) gives Cd for rectangular plates of different aspect ratios 4. Discussion
b/h (see Fig. 3) and Fox and McDonald (1985) give Cd for a C-
shaped screen; both are shown in Table 2. Fig. 4 shows that the ACE has a maximum that varies
Upon inspection of Eq. (14), it is apparent that the ACE in- according to the shape of the FWC. The highest ACE is
creases for increasing drag coefficient. Looking at the figures obtained by the concave FWC. Flat collectors have a lower
on Table 2, this implies that a large aspect ratio is better for a ACE that decreases as the aspect ratio decreases. In summary,
flat FWC, and that a concave FWC should have a higher ACE the highest efficiency is obtained by a concave collector, and
than a flat one. These are important conclusions for designing for flat ones, the larger the aspect ratio, the better. We cannot
better fog collectors. tell the effect of the aspect ratio on concave FWC because we
It is more difficult to get the pressure drop coefficient, C0, could not find for this shape the drag coefficient as a function
because it depends on the free flow area ratio f, and the char- of aspect ratio. The maximum ACE is obtained for a shade co-
acteristics of the mesh fibers and their knit or weave. The efficient of 0.57 and 0.54, for a concave collector and a flat
most complete compilation of correlations for C0 is the one one with aspect ratio of five, respectively. However, this effi-
made by Idel'cik (1960). He makes the difference between ciency changes by less than 0.005 points in the range of shade
wire and silk meshes. The basic correlation is for wire mesh: coefficient from 0.5 to 0.6, for all cases.
Typical large FWC have an aspect ratio of approximately 3
 2  s 2
1 (Schemenauer and Joe, 1989; Schemenauer et al., 2005). There-
C0 ¼ 1:3ð1−f Þ þ −1 ¼ 1:3s þ ð15Þ
f 1−s fore, the maximum efficiency of a flat FWC should be in the
order of 20%, for a mesh with a pressure drop coefficient
given by Eq. (16). This figure is consistent with the measure-
Table 1
Typical data for fog water collectors with Raschel mesh. ments of Schemenauer and Joe (1989). On the other hand,
the SFC that has an aspect ratio of one has an ACE only 0.5%
Variable Value Units smaller than a large one, with aspect ratio of five (Schemenauer
Mesh dimensions, h × b 4 × 12 m and Cereceda, 1994), because the drag coefficient does not
Unperturbed wind velocity, v0 2–6.5 m/s change significantly in this range (see Table 2). In summary,
Air temperature 10 C according to this model, aspect ratio is important only for
Mesh fiber width, δ 1.0 mm
values above five.
340 J.D. Rivera / Atmospheric Research 102 (2011) 335–342

Table 3
Correction factor for Reδ less than 400.
Pole b Re kRe

40 1.16
80 1.05
120 1.01
300 1.00
350 1.01
400 1.01
Wind blown fog
Mesh Idel'cik, 1960, p. 308.

on how the fibers of the two layers overlap. These figures cor-
respond to the range our model predicts for maximum ACE,
assuming that Idel'cik correlation correctly represents the
Ground pressure drop across a double layer of Raschel mesh.
Besides the measurements of mesh and overall collection
efficiency (ηcoll) performed by Schemenauer and Joe (1989),
we could not find other reports of efficiency measurements,
to compare with our theoretical results. This is probably be-
Fig. 3. Sketch of a fog water collector indicating the main dimensions. cause of the difficulties of measuring the LWC of the fog simul-
taneously with the water yield of the FWC. Schemenauer and
Joe (1989) found that the overall efficiency (ηcoll, see Eq. (3))
It is interesting to notice that making the FWC concave of their collector was of the order of 20%, and argue that both
could increase the ACE from 20% for a flat FWC with aspect the capture efficiency and draining efficiency should be very
ratio of five up to 24.5% (see Fig. 4). However, a concave high. Therefore, the ACE of that particular FWC should be a
shape could decrease the draining efficiency since it will few percent points higher than 20%, which agree quite well
probably augment the water spilled on the underside of the with our theoretical results for a concave mesh with a shade co-
concave mesh. It is worth noting that FWC without a frame, efficient of about 0.7.
like the ones recommended by FogQuest (Schemenauer Shanyengana et al. (2003) compared meshes with different
et al., 2005) naturally have a concave shape, which probably shade coefficient, material and pattern and, even though they
gives them a better ACE. did not measure efficiency as such, their results can be explained
This model shows no size scale effect, as long as the Reyn- by the trend indicated by our model. They compared the water
olds number of the FWC is above 10 4, which is true even for collection rate from collectors made with local shade nets of dif-
the SFC. This agrees with the comparison between a SFC ferent shade coefficients with the rate obtained with the 35%
(1 m 2 mesh) and a large fog collector (48 m 2 mesh) reported shade coefficient Raschel mesh used in the SFC (Schemenauer
by Schemenauer and Cereceda (1994), which had a very sim- and Cereceda, 1994). They found a very small difference be-
ilar collection rate. tween the local, 40% shade coefficient, Indoor Aluminet mesh
It is very common for FWC to use a double layer of Raschel and the standard 35% shade coefficient Raschel mesh, both in
mesh with 35% shade coefficient (Schemenauer et al., 2005; double layer. They also found that the collection rate decreased
Schemenauer and Cereceda, 1994). For this arrangement, with increasing shade coefficient. Indeed, the rate of the 60% and
Schemenauer and Joe (1989) estimate a total shade coeffi- 90% shade coefficient mesh was 77% and 38%, respectively, of
cient of less than 0.76 and probably close to 0.5, depending the rate corresponding to the reference 40% shade coefficient
mesh. Since in all cases the mesh was installed as a double
layer, which means that the actual shade coefficient is higher
than the ones mentioned above, this result seem to agree with
Table 2
Drag coefficients for rectangular plates and a C-section with the open side
Fig. 4 that shows a significant decrease of the ACE for very
facing the flow. high shade coefficients. The authors correctly argue that the de-
creasing collection rate with increasing shade coefficient is due
Geometry Aspect ratio (b/h) Drag coefficient (Cd)
to the diversion of the airflow around the FWC, which is the cen-
b 1 1.18 tral part of our model.
5 1.2
Another interesting data to compare with this simplified
10 1.3
20 1.5 model is the results of Computerized Fluid Dynamics (CFD)
h
(1) ∞ 2.0 simulations made by de la Jara (2011). He used a commercial
software (COMSOL Multyphysics 3.5a) to simulate a flat FWC
with a mesh consisting of many vertical, parallel, 1 mm-diameter
2.3 cylindrical fibers, with shade coefficient ranging from 0.1 to 0.9.
He found a maximum ACE of 38% for s=0.6. This higher value
(2)
can be explained by the lower resistance to airflow of the parallel
(1) White, 1986, p. 420. cylindrical fibers, compared to a knitted or woven mesh. This
(2) Fox and McDonald, 1985, p. 460. implies a reduced C0, for which the effect on ACE is discussed
J.D. Rivera / Atmospheric Research 102 (2011) 335–342 341

In order to further explore the potential of manipulating the


aerodynamics of the mesh we plot in Fig. 6 the maximum ACE
as a function of the drag coefficient, Cd, for wire and silk mesh.
The latter has a pressure drop coefficient, C0, larger than the
former by a factor of 1.62. In summary, it seems conceivable
to design a mesh with a better aerodynamic shape such as to
lower its resistance to air flow, having a small pressure drop co-
efficient and, consequently, a high ACE. Nevertheless, it is ap-
parent that the ACE and, hence, the total efficiency of a FWC
is unlikely to exceed around 30%. Additionally, we have to
keep in mind that other factors, like cost, durability and avail-
ability are very important.
Fig. 4. Aerodynamic collection efficiency as a function of shade coefficient for
FWC shaped as flat plates of different aspect ratio and as a concave surface. A brief comment about cost is necessary at this point. It is
obvious that the per cent gain in efficiency of a better mesh
has to be greater than the per cent increase in cost of the sys-
below. This CFD simulation also shows that the actual ACE is tem. However, this is an analysis that has to be made in each
larger than the one predicted by the present model because particular case, since costs may vary significantly from place
the flow lines are not perpendicular to the mesh, as shown to place. Nevertheless, it is worth noting that the cost of the
by Fig. 2, and this increases the effective shade coefficient. In- mesh represents a small percentage of the total cost of the in-
deed, simulations of de la Jara show that the ACE may increase stallation, less than 10% of the materials alone (Osess, 2002)
from 26% at 90° incidence angle to 50% at 20° incidence angle, and probably less than 5% considering labor and transportation.
for s=0.3. The varying incidence angle of the flow lines was Therefore, even a large increase in cost for a more efficient
already mentioned by Schemenauer and Joe (1989) as a factor mesh will cause a relatively small increase in the total cost of
that may change the mesh efficiency, since they measured it at the FWC.
the center of the collector where the incidence angle is close to
90°. 5. Concluding remarks
Eq. (14) shows that efficiency increases for decreasing C0/Cd
ratio, which means that if we could shape FWCs in such a way Our simple model has shown that there is an optimum
that lowers this ratio we could get higher water yields. In order shade coefficient for the mesh of a FWC for maximum ACE.
to explore the benefits of this possibility, in Fig. 5 we plot the The order of magnitude of this optimum shade coefficient is
ACE as a function of C0/Cd with the shade coefficient, s, as a pa- 0.5 to 0.6 for the specific mesh represented by the pressure
rameter. It is apparent that the efficiency can increase signifi- drop correlation of Idel'cik (1960) and assuming correct the su-
cantly for C0/Cd ratios lower than 2. It does not seem possible perposition scheme employed. These conclusions do not con-
to get a Cd larger than 2.3 (see data in White, 1986, and Fox sider possible changes either in the droplet capture efficiency
and McDonald, 1985), but C0 could be reduced improving the or the draining efficiency that will also affect the FWC yield.
aerodynamics of the mesh. Indeed, looking at Idel'cik correla- The two largest uncertainties in this model are, precisely, the
tions, we can see that C0 for silk mesh is larger than for wire actual pressure drop through the real mesh and the corre-
mesh by a factor of 1.62 (Eqs. (15) and (16)). This is probably sponding drag coefficient. Nevertheless, our results agree in
due to the smoother cylindrical shape of the wire in the mesh the order of magnitude with the only data on measured effi-
compared to the silk fibers made from many filaments. This ciency (Schemenauer and Joe, 1989). Additionally, the exis-
gives a maximum ACE of about 28% for wire mesh compared tence of an optimum shade coefficient was also demonstrated
to a maximum of 24.5% for a silk one. Therefore, it is conceiv- by de la Jara (2011), by a numerical simulation of a very simple
able to think that with a better mesh it is possible to gain mesh. In order to get more precise figures, it is necessary to im-
some improvement on the FWC's efficiency. plement numerical models capable of including all the relevant

0,30
collection efficiency
Max. aerodynamic

0,25

0,20

0,15
wire
0,10
silk
0,05

0,00
1 1,5 2 2,5
Drag coefficient
Fig. 5. Aerodynamic collection efficiency as a function of C0/Cd (plot of
Eq. (14)), showing significant increase for C0/Cd less than 2. Notice that in Fig. 6. Maximum aerodynamic collection efficiency as a function of the drag
order to change C0/Cd for a constant s, the aerodynamics of the mesh has coefficient, Cd, for wire and silk mesh; silk mesh has a pressure drop coeffi-
to be modified. cient, C0, larger than the one for wire mesh by a factor of 1.62.
342 J.D. Rivera / Atmospheric Research 102 (2011) 335–342

complexities of the flow and the mesh, and validate them with Estrela, M.J., Valiente, J.A., Corell, D., Millán, M.M., 2008. Fog collection in the
western Mediterranean basin (Valencia region, Spain). Atmos. Res. 87,
experimental data. However, to make it a useful tool it is also 324–337.
necessary to characterize all the different kinds of meshes Fox, R.W., McDonald, A.T., 1985. Introduction to fluid mechanics. John Wiley
used in FWCs. It is worthwhile to explore the possibility of de- and Sons, New York.
Gischler, C., 1991. The Missing Link in a Production Chain, Vertical obstacles
signing new meshes with reduced pressure drop especially for to catch Camanchaca. ROSTLAC-UNESCO, Montevideo – Uruguay. ISBN
this particular application. 92-9089-019-7.
It is clear from this model that non-permeable screens, like Idel'cik, I.E., 1960. Memento des pertes de charge. Eyrolles, Paris, translation
from the Russian, original title, Spravotchnik po guidravlitcheskim
plastic or metal sheets do not work for FWC. This fact has also soprotivleniam.
been demonstrated experimentally (Abdul-Wahab et al., Kays, W.M., Crawford, M.E., 1980. Convective heat and mass transfer, 2nd Ed.
2010). It also seems reasonable that meshes with cylindrical fil- McGraw Hill Book Company, New York.
Larrain, H., Velásquez, P., Cereceda, P., Espejo, R., Pinto, R., Osses, P., Scheme-
aments should work better than meshes made of ribbons, like
nauer, R.S., 2002. Fog measurements at the site “Falda Verde” north of
the typical Raschel mesh, because of less resistance to air Chañaral compared with other fog stations of Chile. Atmos. Res. 64,
flow. In summary, it seems feasible to design a better mesh 273–284.
for FWCs, which by increasing the amount of water collected Louw, C., van Heerden, J., Olivier, J., 1998. The South African fog-water collec-
tion experiment: meteorological features associated with water collec-
by square meter could decrease the total cost per liter of tion along the eastern escarpment of South Africa. Water SA 24,
water, if its cost is not significantly increased. 269–280.
Marzol, M.V., 2002. Fog water collection in a rural park in t he Canary Islands
(Spain). Atmos. Res. 64, 239–250.
Acknowledgments Marzol, M.V., 2008. Temporal characteristics and fog water collection during
summer in Tenerife (Canary Islands, Spain). Atmos. Res. 87, 352–361.
Marzol, M.V., Sánchez, J.L., 2008. Fog water harvesting in Ifni, Morocco. An
I would like to thank all the members of Acquaniebla team assessment of potential and demand. Erde 139, 97–119.
for their support and discussion while developing this work. I Olivier, J., 2002. Fog-water harvesting along the West Coast of South Africa: a
also want to acknowledge Dr. Richard LeBoeuf's help in editing feasibility study. Water SA 28, 349–360.
Olivier, J., de Rautenbach, C.J., 2002. The implementation of fog water collec-
this document and Dr. R. S. Schemenauer for his very helpful
tion systems in South Africa. Atmos. Res. 64, 227–238.
advice. Osess, F., 2002. Proyectos de atrapanieblas, metodología de evaluación social
y caso de estudio. Seminario para optar al titulo de Ingeniero Comercial,
mencion economia, Universidad de Chile.
References Reilly, T.E., Franke, O.L., Bennett, G.D., 1984. The principle of superposition
and its application in ground-water hydraulics. U.S. Geological Survey,
Abdul-Wahab, S.A., Lea, V., 2008. Reviewing fog water collection worldwide Series Number 84–459.
and in Oman. Int. J. Environ. Stud. 65, 485–498. Rijsberman, F.R., 2006. Water scarcity: fact or fiction? Agric. Water Manag.
Abdul-Wahab, S.A., Al-Damkhi, A.M., Al-Hinai, H., Al-Najar, K.A., Al-Kalbani, 80, 5–22.
M.S., 2010. Total fog and rainwater collection in the Dhofar region of Schemenauer, R.S., Cereceda, P., 1991. Fog-water collection in arid coastal lo-
the Sultanate of Oman during the monsoon season. Water Int. 35, cations. Ambio 20, 303–308.
100–109. Schemenauer, R.S., Cereceda, P., 1994. A proposed standard fog collector for
Arpaci, V.S., 1966. Conduction heat transfer. Addison Wesley Publishing Co. use in high-elevation regions. J. Appl. Meteorol. 33, 1313–1322.
Bair, E.S., Roadcap, G.S., 1992. Comparison of flow models used to delineate Schemenauer, R.S., Joe, P.I., 1989. The collection efficiency of a massive fog
capture zones of wells: 1. leaky-confined fractured-carbonate aquifer. collector. Atmos. Res. 24, 53–69.
Ground Water 30, 199–211. Schemenauer, R.S., Fuenzalida, H., Cereceda, P., 1988. A neglected water re-
Bresci, E., 2002. Wake characterization downstream of a fog collector. Atmos. source: the Camanchaca of South America. Bull. Am. Meteorol. Soc. 69,
Res. 64, 217–225. 138–147.
Cebral, J.R., Castro, M.A., Soto, O., Löner, R., Alperin, N., 2003. Blood-flow Schemenauer, R.S., Cereceda, P., Osses, P., 2005. Fog Water Collection Manu-
models of the circle of Willis from magnetic resonance data. J. Eng. al. FogQuest, Ontario, Canada. www.fogquest.org.
Math. 47, 369–386. Shanyengana, E.S., Henschel, J.R., Seely, M.K., Sanderson, R.D., 2002. Explor-
Cereceda, P., Osses, P., Larrain, H., Farías, M., Lagos, M., Pinto, R., Scheme- ing fog as a supplementary water source in Namibia. Atmos. Res. 64,
nauer, R.S., 2002. Advective, orographic and radiation fog in the Tara- 251–259.
pacá region. Chile. Atmos. Res. 64, 261–271. Shanyengana, E.S., Sanderson, R.D., Seely, M.K., Schemenauer, R.S., 2003.
Cereceda, P., Larrain, H., Osses, P., Farías, M., Egaña, I., 2008a. The climate of Testing greenhouse shade nets in collection of fog for water supply.
the coast and fog zone in the Tarapacá Region, Atacama Desert. Chile. J. Water Supply Res. Technol. AQUA 52 (3), 237–241 (London, UK).
Atmos. Res. 87, 301–311. Simunek, J., Jarvis, N.J., van Genuchten, M.Th., Gärdenäs, A., 2003. Review
Cereceda, P., Larrain, H., Osses, P., Farías, M., Egaña, I., 2008b. The spatial and and comparison of models for describing non-equilibrium and preferen-
temporal variability of fog and its relation to fog oasis in the Atacama tial flow and transport in the vadose zone. J. Hydrol. 272, 14–35.
Desert. Chile. Atmos. Res. 87, 312–323. Smakhtin, V., Revenga, C., Döll, P., 2004. A pilot global assessment of environ-
de la Jara, E., 2011. Análisis numérico de una malla simple para atrapanie- mental water requirements and scarcity. Water Int. 29, 307–317.
blas, M. Sc. Thesis, Pontificia Universidad Católica de Chile, Santiago, White, F.M., 1986. Fluid Mechanics, second ed. McGraw-Hill Book Company,
Chile. New York.

S-ar putea să vă placă și