Вы находитесь на странице: 1из 14

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

Available at www.sciencedirect.com

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

Removal of bacterial fecal indicators, coliphages and enteric


adenoviruses from waters with high fecal pollution by slow
sand filtration
Rosalie Bauer a, Halim Dizer b, Ingeborg Graeber a, Karl-Heinz Rosenwinkel c,
Juan M. Lopez-Pila a,*
a

Umweltbundesamt (Federal Environmental Agency), Corrensplatz 1, 14195 Berlin, Germany


Helios Klinikum Berlin Buch, Clinic for Physical Medicine and Early Rehabilitation, Schwanebecker Chaussee 50, 13125 Berlin, Germany
c
Institute for Water Quality and Waste Management, Leibniz University of Hannover, Welfengarten 1, 30167 Hannover, Germany
b

article info

abstract

Article history:

The aim of the present study was to estimate the performance of slow sand filtration (SSF)

Received 22 January 2010

facilities, including the time needed for reaching stabilization (maturation), operated with

Received in revised form

surface water bearing high fecal contamination, representing realistic conditions of rivers

3 August 2010

in many emerging countries. Surface water spiked with wastewater was infiltrated at

Accepted 25 August 2010

different pore water velocities (PWV) and samples were collected at different migration

Available online 28 September 2010

distances. The samples were analyzed for phages and to a lesser extent for fecal bacteria
and enteric adenoviruses. At the PWV of 50 cm/d, at which somatic phages showed highest

Keywords:

removal, their mean log10 removal after 90 cm migration was 3.2. No substantial differ-

Slow sand filtration

ences of removal rates were observed at PWVs between 100 and 900 cm/d (2.3 log10 mean

River bank filtration

removal). The log10 mean removal of somatic phages was less than the observed for fecal

Coliphages

bacteria and tended more towards that of enteric adenoviruses This makes somatic phages

Enteric adenoviruses

a potentially better process indicator than Escherichia coli for the removal of viruses in SSF.
We conclude that SSF, and by inference in larger scale river bank filtration (RBF), is an
excellent option as a component in multi-barrier systems for drinking water treatment also
in areas where the sources of raw water are considerably fecally polluted, as often found in
many emerging countries.
2010 Elsevier Ltd. All rights reserved.

1.

Introduction

Worldwide contamination of surface waters with enteric


viruses is of considerable health concern, since insufficient
removal of these pathogens during drinking water treatment
might lead to viral infection and gastroenteritis (WHO,
2006a; Ashbolt, 2004; Van-Heerden et al., 2004; Glass et al.,
2001). In emerging countries the problem of gastroenteritis
poses a burden particularly on the very young, since the

concomitant diarrhea causes acute loss of water which,


untreated, might quickly derange in life-threatening illness.
There the production of healthy drinking water poses
a bigger problem than elsewhere: as surface waters often
receive untreated or poorly treated wastewater, the
concentration of viral and other pathogens might be particularly high, creating an extraordinary challenge for drinking
water treatment (Moe and Rheingans, 2006; Ashbolt, 2004;
Gadgil, 1998). This situation will probably get worse, taking

* Corresponding author. Tel.: 49 30 7978 947; fax: 49 30 8903 1822.


E-mail address: mlopezpila@yahoo.de (J.M. Lopez-Pila).
0043-1354/$ e see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.watres.2010.08.047

into account the water shortages expected in the future due


to climate change, population increase, and to the water
demand associated with industrial development.
Soil filtration systems, as a rule are very efficient and
provide water of very good microbiological quality (Ellis, 1985).
Because of the capacity to reduce a variety of pathogens
(Hijnen et al., 2004) and the low costs and easy operation, slow
sand filtration (SSF) and river bank filtration (RBF), a water
treatment procedure using a similar principle (Ray et al., 2003),
seem especially well suited for the drinking water treatment
in emerging countries (Grunert et al., 2008; Shamrukh and
Abdel-Wahab, 2008) including small communities.
A large number of studies on the removal of viruses
through soil filtration have been conducted in the past (for
reviews see Schijven, 2001, and Jin and Flury, 2002). Studies
with fecally polluted waters are of special interest, as many
surface waters in emerging countries are highly polluted and
the composition of the infiltrate greatly influences the filtration performance of sand filters (Schijven and Hassanizadeh,
2000; Ray et al., 2002; Goldschneider et al., 2007). A high
organic load or surfactants might increase the breakthrough
of pathogens (Schijven, 2001). The results mentioned in the
above reviews indicate that soil filtration can very efficiently
reduce bacteriophages and viruses from fecally polluted
waters under very different working conditions. The studies
do not allow, however, general conclusions regarding maturation time, PWV, and filtration distance.
In order to make an estimate of the removal of viruses in
plants fed with highly polluted surface water, we have infiltrated wastewater-spiked surface water in two facilities, which
represent a medium scale between laboratory experiments and
applied sand filtration in the environment. Here we have
studied the removal of somatic phages, K13-phages, enteric
adenoviruses, and, in one of the test facilities, fecal bacteria.
The dosage of wastewater was adjusted and was chosen in
such a way as to reach a somatic phages concentration of
approximately 104/100 ml, as this was a concentration
observed in very polluted river waters (Grunert et al., 2008).
The study of the influence of the PWV on the removal of
microorganisms in RBF is hampered by the difficulty of
adjusting the PWV under different field conditions. We have
therefore studied such influence in an SSF-facility (Fig. 1A
and B) that incorporates some properties of RBF (even if not
all), including also partial lateral infiltration and tangential
flow of the surface water. This seemed important, as especially
the tangential flow of the water might influence the formation
of the schmutzdecke, known to play an important role in the
removal of microorganisms (Ellis, 1985). Other purposes of this
study were to estimate the time such plants need to reach
stability (maturation), and the influence of the (PWV) on the
removal efficiency for the microorganisms studied.
Because of the similar size and structure of coliphages and
many human viral pathogens (Grabow, 2001), they are
frequently used as process indicators in sand filtration
processes (Schijven and Hassanizadeh, 2000). In order to
evaluate the indicator function of phages, we have compared
the removal of somatic phages, K13-phages, and enteric
adenoviruses during SSF. Human enteric adenoviruses,
worldwide responsible for a number of gastroenteric diseases
(WHO, 2006a), are shed in high numbers with the feces and are

Fig. 1 e Test facilities for simulating SSF and some


hydraulic aspects of RBF. A: Test facility used for
conducting the experiments shown in Figs. 2 and 3. B: Test
facilities for simulating SSF and some hydraulic aspects of
RBF. Delta conductivity (real conductivity minus 950 mSi/
cm, the basis conductivity of the infiltrated water) at the
different sampling points after infiltrating water with
a conductivity of ca. 1,200 mS. The numbers in the right
corner of the figure represent the depth of the sampling
points in cm. The calculated values of the PWV at 30, 60,
90, 120 and 150 cm depth were respectively: 226.2 (68.5%);
238.0 (72.1%); 206 (62.6%); 129.1 (39.1%); and 89.4 (27.1%)
cm/d. The values in parenthesis represent the PWV as % of
the adjusted theoretical value of 330 cm/d. The sharp
slender peak represents the salinity at the outlet of the
drain. C: Test facility used for experiment of Table 2.

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

found in relative high numbers in polluted waters (Grunert


et al., 2008). Unpublished results of the European research
project Virobathe, (Anonymous, 2010) revealed that the
concentration on Escherichia coli in bathing waters significantly
correlated with the number of samples positive for human
enteric adenoviruses, suggesting that adenoviruses are
reasonable indicators of human fecal pollution. The development of inexpensive and manageable concentration methods
for viruses made this goal come closer (Calgua et al., 2008).

2.

Materials and methods

2.1.
Test facilities for SSF and for simulating some
features of RBF
The sand chosen for filling the filters of Fig. 1A and C was free
of clay material. It had a D10 coefficient of 0.269, i.e. 10% of the
grains had a diameter below and 90% above 0.269 cm. The
ratio between the D60 and D10 was 2.42; the Kf-value,
a measure for the water permeability, was 1  103 m s1. The
porosity was 0.32. Therefore, the pore water velocity, i.e. the
water velocity within the filter body, was equal to the velocity
above the filter divided by 0.32. All these values are typical of
coarse sand.
The unit shown in Fig. 1A, in which one part of the
experiments was carried out, consists of a pond
(3.35 m  6.85 m area) filled with sand to a height of 1.80 m
above a 20 cm layer of gravel. It is divided in three sections;
a section of 2.55 m height, 1.80 m length, 3.35 m width, and
a horizontal surface of 6.03 m2; a second section with a slanted
surface, a height between 2.55 and 1.30 m, a length of 325 m,
and a width of 3.35 m; its surface was 11.67 m2 in the slanted
part and 10.89 m2 in its horizontal projection (footprint);
a third section of 1.30 m height, 1.80 m length, 3.35 m width,
and a surface of 6.03 m2. The water was pumped into the top
of the pond through an inlet (inflow) and the level of water
was kept constant at approximately 30 cm above the highest
part of the sand surface by means of an overflow situated at
the opposite side of the inflow. The pore water velocity was
regulated with a pump (not shown) interconnected with the
drain pipe conducting the filtrate (backflow) towards a reservoir (not shown). In this reservoir, in addition to this backflow,
surface water from a groundwater lake was pooled with 1%
raw wastewater from a wastewater treatment plant and the
overflow from the filtration pond. The dimensions of this
reservoir were identical to that of the filtration pond and
allowed a good mixing of surface water with wastewater
before feeding the water to the filtration pond by a submerged
pump. Five sampling pipes at different depths (30 cm, 60 cm,
90 cm, 120 cm, 150 cm), and a distance of 0.5 m from one of the
small sides of the facility, carrying steel faucets at their
opposite ends outside the filtration pond, allowed sampling of
the water. Sampling was performed according to DIN EN ISO
19458 (2006). The pipe draining the water out of the filter
was embedded in the gravel in the basis of the pond.
The test facility was designed to reproduce some hydrogeological aspects of RBF. The filtration flow should have
a vertical as well as a horizontal component and the flowing
water should be kept in a constant current tangential to the

441

filter surface. To achieve this, the shape of the filter was


designed as shown in Fig. 1A, with an upper and a lower
horizontal filter surface and a transitional slanted surface
connecting both. Consequently, the filter resistance in the
upper horizontal part, being higher than in the lower part,
allowed a filtration rate smaller than in the lower surface, the
filtration rate in the slanted part being intermediate.
The thickness of the arrows in Fig. 1A attempts to symbolize
the strength of the infiltration rate entering the different parts
of the facility. This situation generated a slight current of the
supernatant along the slanted part and kept the water in
gentle movement without having to resort to a stirring device
which would have also stirred the sand, compromised the
results in the superficially located sampling pipes, and
possibly disturbed the formation of the schmutzdecke. Obviously, this construction did not generate a homogeneous, easy
to calculate piston-like water infiltration into the filter surface.
In order to estimate the true, effective PWV (PWVeff) above
the different sampling pipes a tracer experiment was carried
out. Sufficient NaCl was added to the water in the water
reservoir feeding the test facility to raise the conductivity in its
water to approximately 1200 mS/cm, and the salinity was
distributed uniformly by a stirring pump. The inflow pump to
the test facility was started and the drain valve opened to
allow a water velocity of 100 cm/d above the filter, corresponding to a theoretical PWV over the horizontal projection
of the filter surface of 320 cm/d (accounting for the porosity of
0.32 of the sand). Every 15 min samples of all sampling pipes
were taken and their conductivity measured. Fig. 1B shows the
results. The PWVeff above the sampling pipes was calculated
from the time period at which 50% of the maximal conductivity was attained. The calculated values at 30, 60, 90, 120 and
150 cm depth were respectively: 226.2 (68.5%); 238.0 (72.1%);
206 (62.6%); 129.1 (39.1%); and 89.4 (27.1%) cm/d. The values in
parenthesis represent the PWV as % of the adjusted theoretical value of 330 cm/d determined by the flow of the drain. As
expected, the PWVeff above all sampling pipes were lower
than the calculated for the horizontal projection of the filter
surface. Consistent with these findings, the PWVeff at the
drain level, strongly influenced by the filtration rate in the
lower horizontal part, was 360 cm/d, higher than the theoretical PWV (sharp peak in the Fig. 1B). The PWVeff of the
sampling sites at 30, 60 and 90 cm were similar, averaging
67.8% of the overall PWV, but at the two deeper sites at 120 and
150 cm the PWVeff were much lower.
The second filtration unit was slightly conical in section
(Fig. 1C). This unit allowed sampling only at 90 cm height. Its
surface was 90 m2 at the top and 49 m2 at the bottom of the
sand. Surface water was mixed with 1% wastewater in the
mixing pond of 6 m3 volume (Fig. 1C), where it was continuously homogenized with a rotating pump. The water flowed
by gravity through a series of weirs towards the filter surface,
where it ponded up to approximately 20 cm height. From here,
the water infiltrated vertically into the filter body. The bottom
of the filter consisted of a layer of 20 cm gravel in which a net
of perforated pipes was embedded to collect the filtrate in
a main drain and to carry it further towards the sewer system.
A faucet in the main drain allowed sampling under controlled
conditions. The PWV was controlled with a pump at the outlet
of the main drain.

442

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

2.2.
Estimation of the arithmetic mean values of the
microorganism concentrations

DIN EN ISO 7899-2 (2000), which is based on MPN with 96 wells


plates. The detection limit of the first method was 7, of the
second one 1 cfu/100 ml. For the analysis of the somatic
coliphages the USEPA method 1602 (2001), with the E. coli
strain WG5 (ATCC700078) as indicator strain was used. The
WG5 strain has a high sensitivity for somatic phages and is
highly suitable for analyzing wastewater, since its resistance
against the antibiotic nalidixic acid allows the suppression of
most bacteria present in wastewater (Havelaar and
Hogeboom, 1983). Instead of the agar formula stated in
USPA1602, the double concentrated modified Scholtens-SoftAgar (DIN EN ISO 10705-2, 2002) was mixed with the sample
before plating, as we had observed that the last formula
enhanced sensitivity for detecting phages. For detecting the
K13-phages, the same procedure was used, except that the E. coli
strain K13 was the indicator. This strain is less sensitive for
somatic phages than the ATCC700078, but has some sensitivity
to F phages (unpublished results of HD and JMLP). We chose
the K13 strain because the broader spectrum of phages detected

Many samples at the 90 and 120 cm depths resulted in nondetects. Using the simplifying assumption that the actual phage
concentrations were constant over time at each PWV, arithmetic
means encompassing the entire experiment were calculated by
dividing all recorded PFU or CFU for each particular PWV and
depth by the total volume taken for the assays. The values
obtained were expressed in the tables as PFU (or CFU)/100 ml. In
the tables, the initial phage concentration above the filters is
expressed as Co, the remaining concentration after filtration as C.

2.3.

Detection of E. coli, IE, and coliphages

E. coli (EC) was analyzed according to DIN EN ISO 9308-3 (1998).


The detection limit was 7 MPN/100 ml. The detection of
intestinal enterococci (IE) was carried out according to DIN EN
ISO 7899-1 (1999), which is based on membrane filtration, or
1,E+05

Somatic Phages
1,E+04

pfu/100 ml

1,E+03

1,E+02

1,E+01

1,E+00
50 cm/day

50 cm/day

100 cm/day

200 cm/day

400 cm/day

1,E-01
1,E+05
K13-Phages
1,E+04

pfu/100 ml

1,E+03

1,E+02

1,E+01

1,E+00

1,E-01
0
5
Days

10

15

20

25

30

35

40

45

50

55

60

65

70

Fig. 2 e Removal of somatic and K13-coliphages, during sand filtration at pore water velocities between 50 and 400 cm/d and
at different filtration depths. A mixture of surface water and wastewater was infiltrated in the facility of Fig. 1A; water
samples were analyzed from inflow (A) and after migration distances of 30 cm (-), 60 cm (:), and 90 cm (3). The
concentrations of the microorganisms are shown at various migration times. Table 1 is a companion to this figure. The
detection limit was 1 pfu/100 ml (lowest gridline in the figure); symbols below this line represent values below detection
limit. The experiment was conducted from October 11th till December, 17th 2004; the temperature of the infiltrate during
this time was between 4 and 9  C.

Table 1 e Arithmetic mean values (C [ organisms/100 ml), removal (Llog(C/Co)), and removala rate (RR) at the different PWVs and depths, of the microorganisms in the
experiment of Fig. 2.
PWV

Arithmetic Means
and Removal at 50 cm/day
(day 1e10)

Arithmetic Means and


Removal at 50 cm/day
(day 11e24)

Arithmetic Means and


Removal at 100 cm/day
(day 25e37)

Arithmetic Means and


Removal at 200 cm/day
(day 38e59)

Arithmetic Means and


Removal at 400 cm/day
(day 60e67)

log(C/Co)

RR

log(C/Co)

RR

log(C/Co)

RR

log(C/Co)

RR

log(C/Co)

RR

0
30
60
90
0
30
60
90

24,471
2.275
613
218
3510
656
48
11

0.00
1.03
1.60
2.05
0.00
0.73
1.86
2.50

e
3.44
1.90
1.50
e
2.43
3.79
2.13

9320
262
87
6
1760
36
12
1

0.00
1.55
2.03
3.19
0.00
1.69
2.17
3.25

e
5.17
1.60
3.87
e
5.63
1.59
3.60

8355
720
336
47
3716
205
46
2

0.00
1.06
1.40
2.25
0.00
1.26
1.91
3.27

e
3.55
1.10
2.85
e
4.19
2.16
4.54

4530
314
146
20
1557
35
16
1

0.00
1.16
1.49
2.36
0.00
1.65
1.99
3.19

e
3.86
1.11
2.88
e
5.49
1.13
4.01

5520
212
149
23
1150
28
17
1

0.00
1.42
1.57
2.38
0.00
1.61
1.83
3.06

e
4.72
0.51
2.70
e
5.38
0.72
4.10

Somatic
phages

K13-phages

a Removal Rate (RR) removal between two adjacent depths, d1 and d2, expressed as log(Cd2/Cd1)*m1. Cd concentration at the depth d.

Table 1a e Arithmetic mean values (C [ organisms/100 ml), removal (Llog(C/Co)), and removala rate(RR) at the different PWVs and depths, of the microorganisms in
a replica of the experiment of Fig. 2.
Velocity

Somatic phages

K13-phages

Arithmetic Means and


Removal at 50 cm/day
(day 1e10)

Arithmetic Means and


Removal at 50 cm/day
(day 11e38)

Arithmetic Means and


Removal at 100 cm/day
(day 39e51)

Arithmetic Means and


Removal at 200 cm/day
(day 52e73)

Arithmetic Means and


Removal at 400 cm/day
(day 74e88)

Depth

log(C/Co)

RR

log(C/Co)

RR

log(C/Co)

RR

log(C/Co)

RR

log(C/Co)

RR

0 cm
30 cm
60 cm
90 cm
0 cm
30 cm
60 cm
90 cm

7868
500
44
6.1
2822
112
10
<0.1

0.00
1.20
2.25
3.11
0.00
1.40
2.45
e

e
3.99
3.52
2.86
e
4.67
3.50
e

8071
163
15
4.0
1671
43
2.1
<0.1

0.00
1.69
2.73
3.30
0.00
1.59
2.90
e

e
5.65
3.45
1.91
e
5.30
4.37
e

8077
39
21
5.2
1390.8
6.1
2.6
0.3

0.00
2.32
2.59
3.19
0.00
2.36
2.73
3.67

e
7.72
0.90
2.02
e
7.86
1.23
3.13

4186
174
27
0.5
822
32.0
1.9
<0.1

0.00
1.38
2.19
3.92
0.00
1.41
2.64
e

e
4.60
2.70
5.77
e
4.70
4.09
e

12243
190
30
17
917
13
1.5
<0.1

0.00
1.81
2.61
2.86
0.00
1.85
2.79
e

e
6.03
2.67
0.82
e
6.16
3.13
e

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

Depth(Cm)

a Removal Rate (RR) removal between two adjacent depths, d1 and d2, expressed as log(Cd2/Cd1)*m1. Cd concentration at the depth d.

443

444

During a period of 67 days, from October 11th till December,


17th 2004, the SSF-facility shown in Fig. 1A was continuously

0.31
0.38
0.46
0.08
0.15
0.23
0.62
0.77
0.92
0.15
0.30
0.45
1.24
1.54
1.84
0.30
0.60
0.90
0.94
0.94
0.94

ponded In the
Total
water sand residence
time

0.19
0.23
0.28

0.47
0.47
0.47

ponded In the
Total
water sand residence
time

0.09
0.12
0.14

0.23
0.23
0.23

ponded In the
Total
water sand residence
time

0.05
0.06
0.07

Inactivation
(log10)
Residence time (days)
at PWV 400 cm/d
Inactivation
(log10)
Residence time (days)
at PWV 200 cm/d
Residence time
(days) at PWV 100 cm/d

Inactivation
(log10)

0.37
0.46
0.55
2.48
3.08
3.68
0.6
1.2
1.8
1.88
1.88
1.88

3.1.
Removal of somatic and K13-phages by slow sand
filtration at different depths and PWVs in the facility of
Fig. 1A

30
60
90

Results

ponded In the
Total
water sand residence
time

3.

Inactivation
(log10)

To sediment the virus, 2 l samples were centrifuged at


15.000 rpm for 10 h using the A-621-Sorvall Rotor (Thermo
Scientific, Germany). The supernatants were discarded; the
sediments containing the viruses were resuspended in
approximately 15 ml glycine buffer (0.25 M, pH 9.5) and pooled.
To separate the viruses further from impurities this suspension
was vortexed, and centrifuged 10 min at 7.000 rpm using the
Sorvall HS-4 rotor (Thermo Scientific, Germany). The supernatants were spun down in 65 ml polycarbonate tubes at
40,000 rpm for 3 h using the Ti 45-Beckmann rotor (Beckman
Coulter, Germany). The supernatants were discarded and the
sediments containing the viruses were dissolved in 0.15 ml PBS
buffer and stored at 20  C for further analyses. For isolation of
the viral nucleic acids from the concentrated sediment, the
QIAamp Viral Mini Kit was used according to the instruction of
the manufacturer. The final volumes of the extracted nucleic
acids were 100 ml. The real-time (RT) PCR-reactions were carried
out with an ABI PRISM 7000 sequence detection system
(Applied Biosystems, Germany). Enteric Adenoviruses genomes
(AdV-Genomes) were detected using a TaqMan probe Ad:ACDEF
(50 -(6FAM)-CCG GGC TCA GGT ACT CCG AGG CGT CCT-TAMRA30 ) and degenerated primers AdHexup 50 -CWT ACA TGC ACA
TCK CSG G-30 (forward), AdHexdo 50 -CRC GGG CRA AYT GCA
CCA G-30 according to Hernroth et al. (2002). The RT-PCR mixture
contained 10e20 ml DNA, 0.75 ml each forward and reverse
primer (20 mM), 0.75 ml TaqMan probe (10 mM), 12.5 ml 2xTaqMan
Universal Mastermix (Applied Biosystems, Germany) and
nuclease free water in a reaction mix of 15e25 ml. Amplification
was performed 2 min at 50  C, 10 min at 95  C followed by 45
cycles (15 s 95  C, 1 min 60  C). Standard curves were generated
by using serial dilutions (range, 1 to 106) of known amounts of
linearized pBR322 plasmids containing the entire hexon region
of Ad41 (Hernroth et al., 2002). Each sample was processed in
triplicate together with a positive (plasmid) and a no-template
control. The detection limit was 10 genome/100 ml of the original sample. Determination of the Chemical Oxygen Demand
(COD) and of Total Suspended Solids (SS) was carried out
according to DIN 38414-9 (1986) and DIN 38409-1 (1987),
respectively.

Residence time
(days)
at PWV 50 cm/d

2.4.
Detection of adenovirus genomes using real-time
polymerase chain reaction, of Chemical Oxygen Demand
(COD), and of Total Suspended Solids (SS)

Depth (cm)

might have made the results obtained with K13 more representative for the mixture of human-pathogenic viruses present
in fecally polluted waters. For the detection of both, the somatic
and the K13-phages, the sample volume was 0.1 and 1 ml for
the infiltrate, and 100 ml for the exfiltrates. Accordingly, the
detection limit was 1 pfu/ml and 1 pfu/100 ml respectively.

Table 2 e Residence time (days) of the infiltrated water and inactivation of the phages at the different PWV and depths (T50 of the phages [ 2 days), in the experiments of
Tables 1 and 1a.

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

445

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

Table 3 e COD and SS in the infiltrate and the filtrate at 90 cm during the experiment of Fig. 2 at different times after starting
infiltration. Samples were taken every 2e3 days.
Infiltrate

Number of samples
COD (mg/l)
Suspended Solids (mg/l)

Exfiltrate at 90 cm

Days 0e67

Days 0e2

Days 3e10

Days 11e24

Days 25e37

Days 38e59

Days 60e67

22
3.8  1.7
1.5  0.97

2
2.6  1.9
<0.2

3
4.4  0.8
<0.2

4
2.6  1.0
<0.2

3
2.25  0.7
<0.2

7
1.9  0.24
<0.2

2
2.7.0  0.5
<0.2

infiltrated with surface water spiked with 1% wastewater. The


temperature of the infiltrate during this time ranged between 4
and 9  C. The course of the experiment is shown in Fig. 2.
During the first 24 days the PWV was adjusted to 50 cm/d followed by a period of 13 days at 100 cm/d, 22 days at 200 cm/d,
and 8 days at 400 cm/d. The removal of phages improved
during the first 25 days, presumably as a consequence of
maturation of the filter. The increase of the PWV from 50 to
100 cm/d reduced the removal, but further stepping up of the
PWV to 200 and 400 cm/d had little effect on it. For a better
comparison of the removal of the phages Table 1, companion
to Fig. 2, summarizes the results quantitatively, showing the
arithmetic mean values, the removal, and the removal rate
(RR) for each microorganism at each velocity and depth. On the
simple basis of presumptive counts the observed removal of
the phages was highest at 50 cm/d (excluding the first 10 days),
with log10 removals for somatic phages of 1.55, 2.03, and 3.19
at the 30, 60 and 90 cm depths respectively. Substantial
differences in observed phage removal at 100, 200, and
400 cm/d were not observed. For the K13-phages, at 50 cm/d
(days 11e24), the log10 removals (excluding the first 10 days)
were 1.69, 2.17, and 3.25 at the 30, 60 and 90 cm depths
respectively. The removals at the other PWVs were similar. A
second experiment with identical design was carried out in the
same facility during the period from 30 June to 24 October 2005,
during which the temperature of the infiltrate ranged between
11 and 25  C. In this experiment the infiltration periods were 48
and 15 days at PWVs of 50 and 400 cm/d respectively. Phage
removals during this second experiment were consistent with
those presented in Fig. 2; however the observed mean reductions at each of the filtration distances were generally higher
during the second experiment (Table 1a).
In order to estimate the contribution that viral inactivation
might have in the total removal of both experiments, Table 2
shows the residence times of the infiltrate while being above

the filter in the ponded water, inside the filter, and as the sum
of both periods. In a former study (Dizer et al., 2005) the halftime (T50) for both, the somatic and the K13-phages, had been
found to be approximately 2 days under the same conditions
as used here, i.e. in 1% wastewater running in an artificial
outdoor channel. Using this data, one obtains the inactivation
stated in the Table 2. As can be observed comparing Table 2
with Tables 1 and 1a, inactivation contributes to a relatively
small extent to the total removal. If, e.g., filtration of somatic
phages at 400 cm and 90 cm depth in Table 1 is considered,
0.07 out of 2.38 log10 total removal might be ascribed to inactivation. At 50 cm/d (days 1e24) and 90 cm depth, the contribution due to inactivation was 0.55 log10, from a total removal
of 3.19 log10. It becomes evident that the bulk of the removal
takes place due to filtration processes other than inactivation.
The oxygen concentration in the infiltrated water fluctuated
between 8 and 12 mg/l during both experiments, and was always
below 0.1 mg/l at all sampling points from two days on after the
start of the experiments. During the experiment of Table 1 the
COD of the infiltrate averaged 3.88 mg/l, the SS 1.5 mg/l. These
parameters were determined also in the 90 cm exfiltrate.
Whereas the SS were drastically reduced, the reduction of the
COD, if any, was moderate (Table 3). No COD and SS were
determined in the experiment of Table 1a. The residence times
and the inactivation during them are shown in Table 2, the COD
and SS concentrations were similar to the ones in Table 3.

3.2.
Removal of somatic and K13-phages, and of E. coli,
intestinal enterococci and adenovirus genomes by slow sand
filtration at 90 cm depth and 900 cm/d PWV in the facility of
Fig. 1C
Since the removal rates observed at PWVs of 100e400 cm/d did
not greatly differ from each other, a further experiment was
conducted at a PWV of 900 cm/d. Because the facility of Fig. 1A

Table 4 e Arithmetic means (organisms or genomes/100 ml) and removal (Llog(C/Co)) values of somatic and K13bacteriophages, E. coli, intestinal enterococci, and AdV-genomes after infiltration of 1% wastewater at 900 m/d PWV and
90 cm depth.

Somatic phages (26 samples of both depths)


K13-phages (26 samples of both depths)
E. coli (26 samples of both depths)
Intestinal enterococci (26 samples of both depths)
AdV-Genomes (4 samples of both depths)

Depth

Arithmetic Means at 900 cm/day

log(C/Co)

0 cm
90 cm
0 cm
90 cm
0 cm
90 cm
0 cm
90 cm
0 cm
90 cm

19,735
36
7675
2
195,273
15
35,920
8
7600  2180
<10

0
2.74
0
3.58
0
4.11
0
3.65
0
>2.88

446

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

Table 5 e Arithmetic means (organisms or genome copies), removal values (Llog(C/Co)) and removala rate (RR) of somatic
and K13-phages (pfu/100 ml), and of AdV-genomes (genomes/100 ml) at different depths. COD and SS concentrations of the
infiltrate and exfiltrates.
Depth
Somatic phages

K13-phages

AdV-Genomes

COD (mg/l) 0.12 samples each

SS (mg/l) 12 samples each

Arithmetic Mean

0 cm
30 cm
120 cm
0 cm
30 cm
120 cm
0 cm
30 cm
120 cm
0 cm
30 cm
120 cm
0 cm
30 cm
120 cm

2445
82
2.1
1708
6.1
<0.1
766
26
<10
2.8 
2.4 
3.1 
1.5 
<0.2
<0.2

0.7
0.8
1.0
1.0

log(C/Co)

RR

0
1.47
3.07
0
2.45
>4.23
0
1.47
>1.88
e
e
e
e
e
e

e
4.91
1.77
e
8.16
>1.98
e
4.90
>0.46
e
e
e
e
e
e

a Removal Rate (RR) removal between two adjacent depths, d1 and d2, expressed as log(Cd2/Cd1)*m1.Cd concentration at the depth d.

could not be operated at velocities of higher than 400 cm/


d without considerably risk of distorting the PWVs between
the different depths, the filtration facility of Fig. 1C was used.
This facility could only be sampled at 90 cm depth. As in the
previous experiment, wastewater-spiked surface water,
which had been homogenized in a reservoir using a pump
stirring the water/wastewater mixture, was infiltrated at
a PWV of 900 cm/d during 14 days. The experiment was conducted from 16th to 30th October 2005, the temperature of the
infiltrate ranged from 12 to 14.5  C. Beside the somatic and the
K13-phages, in this experiment the fecal indicators EC and IE
were additionally examined. In addition, four samples of the
infiltrate and exfiltrate were assayed for AdV-genomes in nonconsecutive days. The residence time above the filter surface
(height of the water: 20 cm) was less than 2 h, the microorganisms decay therefore negligible (Dizer et al., 2005). The
arithmetic means of the organisms and viral genomes and
their log10 removals are given in Table 4. The removal of
somatic phages was 2.74 log10, the ones of E. coli and intestinal
enterococci 4.11 log10 and 3.65 log10 respectively. The log10
removal of AdV-genomes was more than 2.88.

3.3.
Removal of somatic and K13-phages, and of
adenovirus (AdV) genomes by slow sand filtration at 30 and
120 cm migration distance in the facility of Fig. 1A
In order to estimate the removal of enteric AdV-genomes (as
a potential surrogate for the AdV) during SSF, we carried out
a similar experiment as the one shown in Fig. 2 in the same
facility. The experiment was carried out from July 30th till
August 15th 2005; the infiltrate temperature ranged between
18 and 25  C. The PWV was kept constant during 48 days at
50 cm/d (calculated for the depth at 30 cm). Due to the flow
characteristics of the facility (see its description in the Material section) this PWV corresponded to approximately 29 cm/d
at a depth of 120 cm. The arithmetic means, the removal, and
the removal rates of both phages and of AdV-genomes are
shown in Table 5. No AdV-genomes were detected at 120 cm

(values for 90 cm depth are not available), which corresponds


to a log10 removal of at least 1.88. As in the first experiment
shown in Fig. 2, the oxygen concentration in the effluents
decreased below 0.1 mg/l after two days operation.

4.

Discussion

The concentration of phages remaining after the infiltration of


wastewater-spiked surface water through different filtration
distances is shown in Fig. 2 and in Tables 1 and 1a. From the
tenth day on, removal between 0 and 30 cm was at least 1 log10
for both phages. Between 30 and 60 cm the removal was less
pronounced. The higher removal of the upper layer presumably was due to the presence of a schmutzdecke above the top
layer, known to significantly contribute to the removal efficiency (Ellis, 1985; Unger and Collins, 2006). Removal at 90 cm
was at least 2.25 log10. The removal of the phages in the
experiment of Fig. 3 and Table 4 were similar to the ones in
Fig. 2 and Tables 1 and 1a.
As Figs. 2 and 3 show, the estimated influent phage
concentration varied from day to day, likely due to the changing
quality of the wastewater used for spiking. The concentrations
at the different migration distances varied as well. This variability made an accurate calculation of the quantitative
removal very difficult since such a calculation would require
that the phage concentration of the influent be constant at least
within certain limits. Nonetheless, this has been done to obtain
an approximate estimate of the extent of phage removal.
Influent and effluent phage concentrations at a given PWV were
averaged and log removals were subsequently calculated using
those mean values as [log10(C/Co)] (Tables 1, 1a, 4, and 5).
These mean concentrations are uncertain because of likely
temporal variability (i.e., individual concentrations were estimated on different days, and concentration peaks or valleys
might have been gone unnoticed if they happened to occur
between two adjacent measurements). Moreover, the plating
methods used to obtain the phage concentration estimates

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

1,E+04

pfu/100 ml

1,E+03
1,E+02
Somatic phages
1,E+01
1,E+00
1,E-01
0

10

20

30

40

50

1,E+04
K13-phages

pfu/100 ml

1,E+03

1,E+02

1,E+01

1,E+00

1,E-01
0

10

20

30

40

50

Genome copies/100 ml

1,E+04
1,E+03
1,E+02
AdV-Genomes
1,E+01
1,E+00
1,E-01
0

Days

10

20

30

40

50

Fig. 3 e Removal of somatic and K13-coliphages, as well as


enteric adenoviruses (as genome copies) during sand
filtration at 50 cm/d at depths of 30 and 120 cm. A mixture
of surface water and wastewater was infiltrated in the
facility of Fig. 1A; water samples were analyzed from
inflow (A), and after migration distances of 30 cm (-) and
120 cm (:). The concentrations of somatic phages (top),
K13-phages (middle) and genome copies of enteric
adenoviruses (AdV-genomes) (bottom) are shown at
various migration times. Table 3 is companion to this
figure. The detection limit for the phages was 1 pfu/100 ml,
for the adenovirus DNA 10 genome copies/100 ml (lowest
gridline in the figure); symbols below this line represent
values below detection limit.

presented in Fig. 2 are prone to unavoidable measurement


errors (Emelko et al., 2010; Schmidt et al., 2010). The uncertainty
in these estimates increases when counts are low; particularly
when counts are less than w10 pfu (Emelko et al., 2008), as
shown in Fig. 2. As a result, the slight differences in the phage
removals observed at 50 cm/d compared to those at higher

447

PWVs cannot necessarily be attributed to the differences in


PWV because the relative contribution of measurement error is
unknown.
A number of removal studies have been conducted with
groundwater, or with surface water low or free of fecal
pollution, spiked with laboratory-grown viruses (Table 6, for
a more extensive treatment of removal studies in saturated
columns, see Jin and Flury, 2002). As a rule, the removal rates
observed in the studies mentioned in Table 6 were less
pronounced than in studies carried out using wastewater as
the carrier water (Table 7), what suggests that contents in
fecally polluted water enhances viral removal. Some workers
also found strong evidence for a better removal at low PWVs
(Lance et al., 1982; Quanrud et al., 2003). Using tertiary treated
wastewater, Vaughn et al. (1981) found very little virus
removal at a PWV of 21 m/d, but decreasing the PWV to 1.4 m/
d, or even lower, greatly improved the removal. However,
tertiary treated wastewater, having been treated with
precipitating salts and/or coagulating agents, might be not the
proper representative matrix for the highly fecally polluted
waters addressed in the present paper, since tertiary treated
sewage has been found to foster adsorption of viruses to sand
as compared to secondary effluent (Dizer et al., 1984). On the
other side, polioviruses, used in the study Vaughn et al. do not
seem to be conservative indicators for the removals of viruses,
as they have been found to be much more efficiently removed
by soil filtration than the MS2, PRD1 or FX174-phages
(DeBorde et al., 1999).
We have carried out the present work in order to obtain
additional information to evaluate the viral removal of sand
filtration plants that infiltrate fecally polluted or very polluted
surface water, as often encountered in emergent countries.
We have used phages in this study, but as laboratory-grown
phages might not be in the same state of aggregation as
indigenous phages (and pathogenic viruses), we have used
indigenous phages as index organisms, in spite of the lower
numbers sometimes found in the wastewater and of their
somewhat erratic concentrations. Besides, indigenous
phages, being not the clonal progeny of a single E. coli strain,
but a mixture of clones generated in many different indigenous E. coli strains, avoid biases when a monoclonal population is used: Landry et al. (1979) have reported that
a guanidine-resistant clone of poliovirus 1 displayed different
soil adsorption properties than the parent strain. As tests
organisms we have aimed at using a relatively broad spectrum
of phages to make the study more representative for the
human-pathogenic viruses present in fecally polluted water;
in addition of using the E. coli-indicator strain WG5, which
mainly sustains multiplication of somatic phages, we have
used the strain K13 of E. coli. This strain multiplies both, Fphages and T-phages, albeit at poorer efficiencies than other
indicator strains that multiply either phage type only
(unpublished results of HD and JMLP). Tables 1 and 1a show no
clear dependency between the removals at different PWVs.
Only when the PWV was raised from 50 to 100 cm/d the
breakthrough of phages increased as well, but it receded back
shortly thereafter to levels existing before raising the PWV.
The increase of breakthrough might be explained by the
enhanced fluid shear stress dislodging some of the phages
from their attachment sites, the further building up of biomass

448

Table 6 e Transport of viruses in waters free of, or low in, fecal pollution.
Virus infiltrated

Nature of the Infiltrate

PWV
(m/d)

MS2; PRD1; VX174;


Poliovirus 1
MS2; PRD1; VX174;
Poliovirus 1
MS2; PRD1

Virus suspension
in groundwater
Same as above

Experimental field
(Aquifer: clasts and gravel)
Same as above

22e45 Bromide:
22.5e30
As above

7.5

Phage suspension
in pretreated
river water
Same as above

Infiltration plant for


groundwater recharge.
(Aquifer: Fine sand)
Experimental field.
(Aquifer: sandy layers
of fluvial sediments)

ca. 1.6

up to 30

3.33e1.01

up to 40

ca. 6 after 8 m for


both phages, then 2.3
for MS2 after
additional 32 m.

PRD1; MS2; VX174;


Poliovirus 1
MS2

Same as above

Same as above

ca. 132

21.5

Natural aquifer

MS2

Phage suspension in
pretreated (coagulation,
rapid sand filtration,
softening, etc.)
surface water
Phage suspension in
pretreated (rapid sand
filtration) surface water

Pilot plant, 2.56 m2,


sand of 0.3 mm diameter

7.2

1.5

0.26; 0.77; 1.6; 2.9


respectively
Ca. 0.6/day initial
removal; 0.2/day
afterwards
1.7e1.8

Pilot plant, 2.56 m2,


sand of 0.3 mm diameter

7.2

1.5

Phage suspension in
pretreated (coagulation,
rapid sand filtration,
softening, etc.)
surface water
Phage suspension in
anoxic groundwater

Columns (0.4 m;
0.09 m diameter)

7.2

0.4

Experimental field.
(Aquifer: coarse sand)

0.33e0.56

up to 37.7

MS2; PRD1

MS2

MS2

MS2, VX174

Filtration
distance (m)

19.4

Total removal
log10 (C/Co)
0.3; 0.54; 0.46; 2.0
respectively
0.82; 0.92; 1.22; 2.7
respectively
ca 8.3 in 30 m

Removal rate
(log10 (C/Co)/m)
0.034; 0.072; 0.061; 0.27
respectively
0.042; 0.047; 0.063; 0.14
respectively
Ca. 0.7 in the initial
part; ca. 0.15
in the linear part.
Ca. 0.7 in the
initial part for both
phages. Ca 0.08
between 8 and
38 m for MS2
0.012; 0.036; 0.073;
0.136 respectively
e

1.13e1.2

Reference
DeBorde et al. (1999)
DeBorde et al. (1999)
Schijven et al. (1999)

Schijven et al. (2000)

Woessner et al. (2001)


Medema and Stuyfzand
(2002) cited by van der
Wielen et al. (2008)
Hijnen et al. (2004)

1.7e1.9 (4 days old


schmutzdecke)
1.8e2.2
(81 days old
schmutzdecke)
0.1
0.2
0.4

1.13e1.27 (4 days
old schmutzdecke)
1.2e1.47 (81 days
old schmutzdecke)

Hijnen et al. (2004)

0.25
0.5
1.0

Hijnen et al. (2004)

4.4 (after 37.7 m)


and 2.8 (after 7.8 m)
respectively

0.12 and 0.36


respectively

van der Wielen


et al. (2008)

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

Infiltration field

Table 7 e Transport of viruses suspended in wastewater.


Infiltrated Virus(es)

Nature of
the Infiltrate

Infiltration field

PWV
(m/d)

Filtration
distance (m)

Total removale(log10)

Removal rate (log10/m)

Reference

Soil basins (6.1  213.4 m)


Fine loamy sand;
gravel, cobbles.

0.25

>4

>0.45

Gilbert et al. (1976)

Viral suspension in
secondary effluent

Saturated soil columns,


2.75 m and 10 cm diameter.
Loamy sand 89%;
silt 8%; clay 3%

15; 55

Up to 1.60

5.4 and 5.0 after 0.4 depth;


2.1 and 2.0 between 0.4 and
1.6 m respectively for
15 and 55 PWV

Lance et al. (1976)

Poliovirus 1

Viral suspension in
tertiary effluent

0.24

Up to 0.75

10 during the first 10 cm; 6.7


during 10 and 25 cm; None
between 25 and 75 cm

Landry et al. (1980)

Poliovirus 1

Viral suspension in
tertiary effluent

Soil cores from a


experimental
recharge basin for tertiary
effluent. Coarse sand/fine
gravel with 1.25% silt and clay
Experimental recharge basin
for tertiary effluent (25 m2) t.
Coarse sand/fine gravel with
1.25% silt and clay

2.1 and 2.0 for


15 and 55 after 0.4 m
depth; 2.52 and 2.42
between 0.4 and 1.60 m
respectively for 15 and 55 PWV
1 after 10 cm; 1 between
10 and 25. Between 25 and
75 cm no apparent removal

21; 1.4;
0.24; 0.12

Up to 7.62

0.42; 4.59; 4.29; 6.80


respectively (after
an infiltration
distance of 0.75 m)

Vaughn et al. (1981)

Poliovirus 1 Echo
1 Echo 29

Viral suspension in
secondary effluent

0.5e0.6

Up to 2.75

5 between 0 and
0.4 m; >0.83
between 0.4 and 1.6 m

Lance, et al. (1982)

Echo 1

Viral suspension in
secondary effluent

Saturated soil columns,


2.75 m  10 cm (i.d.).
Loamy sand 89%;
silt 8%; clay 3%
Saturated soil columns,
2.75 m  10 cm (i.d.).
coarse sand 93%;
silt 6%; clay 1%

0.6; 1.2; 2.4

250

0.31; 3.44; 3.22; 5,10


respectively for the different
PWV (Pulse experiments.
Removal is expressed as the
log10 of the ratio between the
maximal concentration in the
exfiltrates after 0.75 m
infiltration to the
concentration in the infiltrate)
2 after 0.4 m infiltration: >1
between 0.4 and 1.6 m
infiltration. Undetectable
afterwards
Qualitatively: better removal
at the lower PWV

Lance, et al. (1982)

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

Secondary
effluent

Enteroviruses,
reoviruses
indigenous in
secondary effluent
Poliovirus 1

449

450

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

might have been responsible for the resumption of the


removal. According to the results of Tables 1, 1a, and 2 the
phage removal was similarly high at all pore water velocities
between 1 and 9 m/d. This affirm regulations of the German
Water Association (Deutscher Verein des Gas- und Wasserfaches, DVGW) which specify that the velocity of the water
penetrating the filter should be between 0.05 and 0.30 m/h,
corresponding to an approximate PWV between 3.6 and
21.6 m/d (Anonymous, 2005).
The stabilization (maturation) of the slow sand filter in the
experiment of Fig. 2 was apparently well advanced after 10
days. During this period the removals of the phages were, at
all depths, lower than the ones corresponding to the following
14 days (Table 1). Maturation might take several more weeks,
though, as suggested by Fig. 3 for the values at 30 cm depth.
During this time, the removal efficiency for the organisms
studied increased steadily. Maturation time probably greatly
depends on the formation of a biological matrix not only
above (schmutzdecke), but within the filter as well. During the
experiments in the plant of Fig. 1A a vigorous development of
biofilms was observed, and the density of biofilm bacteria
decreased with filter depth (personal communication of
Szewzyk and Szewzyk (2009)).
As shown in Table 4 somatic phages (log10 removal: 2.74) as
process indicators are more conservative than either K13phages (log10 removal: 3.58) or the bacterial indicators (log10
removal for IE: 3.58; for E. coli: 4.11). Somatic phages seem also
to be more conservative than adenovirus DNA, since from the
experiment in the facility of the Fig. 1C the log10 removal of
viral genomes was higher than that of somatic phages.
Although indigenous phages, taking advantage of the fecal
bacteria present in the filter, might multiply to some extent
during filtration, this possibility does not hamper their value
as conservative indicators of viral attenuation. The concentration of both, somatic and K13-phages decreased with
increasing filter depth. In the case of phage multiplication
during filtration the filter performance for viral removal could
be rather under than overestimated.
Systems relying on sand or soil filtration might be
a considerable help to reduce the incidence of water borne
gastroenteritis in countries with highly polluted rivers. If, for
instance, a stream contains 100 infectious particles of a pathogenic virus per liter which, after ingestion of one viral
particle, causes illness in 10% of the cases, then, in order to
reduce the risk of illness down to 104/d and person, the viral
concentration will have to be reduced by five logs to a final
concentration of 103 viruses/l (assuming that each person
consumes 1 l water per day, which is an underestimate).
Following the suggestions of WHO (2006b), this removal
should be accomplished by a multiple-barrier system, each
component of the system resilient to potential malfunctions
of the other(s). Assuming that the last part of the system
should be chemical disinfection with chlorine, hypochlorite or
chlorine dioxide, the other part could be chosen to be SSF (or
RBF) which, carried out with a filtration path of e.g. 10e20 m at
PWV not above 10 m/d, would reduce the viral load of the
water three to four logs and probably much more. The
resulting filtrate, having been practically freed of the suspended solids and partially of the organic substances originally present in the surface water, would allow a much more

efficient chlorination than if the disinfection had been carried


out directly with the surface water. In fact, chlorination has
been reported to reduce indigenous somatic and male phages
less than one log10 only, if it was carried out in secondary
effluent (Tyrrell et al., 1995). Other alternatives for drinking
water treatment which encompass filtration (as opposite to
chemical treatment like e.g., lime softening) are fast sand
filtration following precipitation/coagulation, and membrane
filtration. Both filtration techniques require higher economic
investments than SSF, both in equipment and maintenance.
Membrane filtration plants, especially in the versions which
efficiently remove viruses, seem prohibitive for low-income
communities. Fast sand filtration, more affordable, does not
outperform SSF in the removal of viruses (WHO, 2006c).

5.

Conclusions

Slow sand filtration removes indigenous somatic phages from


fecally polluted water very efficiently; after 90 cm filter path
the minimum removal was 2.25 log10 (Table 1), the maximum
3.92 log10 (Table 1a). The removal of indigenous K13-phages
was even higher, Removal of indigenous AdV (as AdVgenomes) was >2.88 log10. The time for the filters to stabilize
and reach a constant performance (maturation) was found to
be approximately 10 days, but progression to maturity up to 40
days was also observed. At least in mature sand filters, PWVs
between 1 and 9 m/d were equally efficacious in removing
indigenous phages, an important fact if the filters are to be
efficiently operated under cost/benefits aspects.
SSF (and by inference RBF) is able to greatly reduce the viral
concentration of highly fecally polluted surface waters, such
as are found in many emergent countries.

Acknowledgments
This investigation was carried out partly as a contribution to
the Natural and Artificial Systems for Recharge and Infiltration project (NASRI project) with the financial support of the
Kompetenzzentrum Wasser Berlin (KWB). The support of the
German Bundesministerium fur Gesundheit for the project
Kleinbadeteiche is greatly acknowledged. The authors thank
Mrs. Ch. Mekonnen and Mrs. Ch. Arndt for excellent logistical
and technical assistance.

references

Anonymous, 2005. Filtrationsverfahren zur Partikelentfernung. Teil


4: Langsamsandfiltration (Filtration Procedures for the Removal
of Particles. Part 4: Slow Sand Filtration). Deutscher Verein des
Gas- und Wasserfaches, Bonn. DVGW Arbeitsblatt W 213-4-LSF.
Anonymous, 2010. Virobathe, An EU Specific Targeted Research
Project. http://www.virobathe.org/.
Ashbolt, N., 2004. Microbial contamination of drinking water and
disease outcomes in developing regions. Toxicology 198 (1e2),
229e238.
Calgua, B., Mengewein, A., Grunert, A., Bofill-Mas, S., ClementeCasares, P., Hundesa, A., Wyn-Jones, A.P., Lopez-Pila, J.M.,

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

Girones, R., 2008. Development and application of a one-step


low cost procedure to concentrate viruses from seawater
samples. Journal of Virological Methods 153, 79e83.
DeBorde, D.C., Woessner, W.W., Kiley, Q.T., Ball, P., 1999. Rapid
transport of viruses in a floodplain aquifer. Water Research
33 (10), 2229e2238.
DIN 38409-1 Norm, 1987. Deutsche Einheitsverfahren zur
Wasser-, Abwasser- und Schlammuntersuchung;
Summarische Wirkungs- und Stoffkenngroen (Gruppe H);
Bestimmung des Gesamttrockenruckstandes, des
Filtrattrockenruckstandes und des Gluhruckstandes (H 1).
Beuth, Berlin, vol. 6. Deutsches Institut fur Normung e.V.
Burggrafenstrae, Berlin. 10787.
DIN 38414-9, 1986. Deutsche Einheitsverfahren zur Wasser-,
Abwasser- und Schlammuntersuchung; Schlamm und
Sedimente (Gruppe S); Bestimmung des Chemischen
Sauerstoffbedarfs (CSB) Beuth, Berlin, vol. 6. Deutsches
Institut fur Normung e.V. Burggrafenstrae, Berlin. 10787.
DIN EN ISO 10705-2, 2002. Nachweis und Zahlung von
BakteriophageneTeil 2: Zahlung von Somatischen Coliphagen
(ISO 10705-2:2002); Deutsche Fassung EN ISO 10705-2:2002.
Beuth, Berlin, vol. 6. Deutsches Institut fur Normung e.V.
Burggrafenstrae, Berlin. 10787.
DIN EN ISO 19458, 2006. Ausgabe: 2006-12,
WasserbeschaffenheiteProbenahme fur mikrobiologische
Untersuchungen. Beuth, Berlin, vol. 6. Deutsches Institut fur
Normung e.V. Burggrafenstrae, Berlin. 10787.
DIN EN ISO 7899-1, 1999. Wasserbeschaffenheit - Nachweis und
Zahlung von intestinalen Enterokokken in Oberflachenwasser
und Abwasser - Teil 1: Miniaturisiertes Verfahren durch Animpfen
in Flussigmedium (MPN-Verfahren) (ISO 7899-1, 1999); Deutsche
Fassung EN ISO 7899-1:1999. Beuth, Berlin, vol. 6. Deutsches
Institut fur Normung e.V. Burggrafenstrae, Berlin. 10787.
DIN EN ISO 7899-2, 2000. Wasserbeschaffenheit - Nachweis und
Zahlung von intestinalen Enterokokken - Teil 2: Verfahren
durch Membranfiltration (ISO 7899-2:200); Deutsche Fassung
EN ISO 7899-2:2000. Beuth, Berlin, vol. 6. Deutsches Institut fur
Normung e.V. Burggrafenstrae, Berlin. 10787.
DIN EN ISO 9308-3, 1998. WasserbeschaffenheiteNachweis und
Zahlung von Escherichia coli und coliformen BakterieneTeil 3:
Miniaturisiertes Verfahren durch Animpfen in Flussigmedium
(MPNeVerfahren) (ISO 9308-3:1998); Deutsche Fassung EN ISO
9308-3:1998. Beuth, Berlin, vol. 6. Deutsches Institut fur
Normung e.V. Burggrafenstrae, Berlin. 10787.
Dizer, H., Nasser, A., Lopez, J.M., 1984. Penetration of different
human pathogenic viruses into sand columns percolated with
distilled water, groundwater, or wastewater. Applied and
Environmental Microbiology 47 (2), 409e415.
Dizer, H., Wolf, H.S., Fischer, S.M., Lopez-Pila, J.M., Roske, I.,
Schmidt, R., Szewzyk, R., Wiedenmann, A., 2005. Die Novelle der
EU-Badegewasserrichtlinie Aspekte der Risikobewertung bei der
Grenzwertsetzung (The EU Bathing Water Directive. Risk
assessment and standards). Bundesgesundheitsblatt 48, 607e614.
Ellis, K.V., 1985. Slow sand filtration. CRC Critical Reviews in
Environmental Control 15, 315e354.
Emelko, M., Schmidt, B.P., Roberson, J.A., 2008. Regulatory
updateequantification of uncertainty in microbial
dataereporting and regulatory implications. Journal of the
American Water Works Association 100 (3), 94e104.
Emelko, M., Schmidt, B.P., Roberson, J.A., Reilly, P.M., 2010.
Particle and microorganism enumeration data: enabling
quantitative rigor and judicious interpretation. Environmental
Science & Technology 44 (5), 1720e1727.
Gadgil, A., 1998. Drinking water in developing countries. Annual
Review of Energy and the Environment 23, 253e286.
Gilbert, R.G., Gerba, C.P., Rice, R.C., Bouwer, H., Wallis, C.,
Melnick, J., 1976. Virus and bacteria removal from wastewater

451

by land treatment. Applied and Environmental Microbiology


32 (3), 333e338.
Glass, R.I., Bresee, J., Jiang, B., Gentsch, J., Ando, T., Fankhauser, R.,
Noel, J., Parashar, U., Rosen, B., Monroe, S.S., 2001.
Gastroenteritis Viruses: An Overview, vol. 238. Novartis
Foundation Symposium. 5e25.
Goldschneider, A., Haralampides, K., Macquarrie, K., 2007. River
sediment and flow characteristics near a bank filtration water
supply: implications for riverbed clogging. Journal of
Hydrology 344 (1e2), 55e69.
Grabow, W., 2001. Bacteriophages: update on application as
models for viruses in water. Water SA 27 (2), 251e268.
Grunert, A., Lopez-Pila, J., Girones, R., Selinka, H., Ronghang, M.,
Mittal, A., Sprenger, C., Lorenzen, G., 2008. Removal of Viruses
from Surface Water during Groundwater Recharge via River
Bank Filtration in Delhi (India) COST-Symposium: Current
Developments in Food and Environmental Virology Pisa
9the11th October 2008.
Havelaar, A.H., Hogeboom, W.M., 1983. Factors affecting the
enumeration of coliphages in sewage and sewage-polluted
waters. Antonie van Leeuwenhoek 49, 387e397.
Hernroth, B., Conden-Hansson, A., Rehnstam-Holm, A.,
Girones, R., Allard, A., 2002. Environmental factors influencing
human viral pathogens and their potential indicator
organisms in the blue mussel Mytilus edulis: the first
scandinavian report. Applied and Environmental Microbiology
68 (9), 4523e4533.
Hijnen, W., Schijven, J., Bonne, P., Visser, A., Medema, G., 2004.
Elimination of viruses, bacteria and protozoan oocysts by slow
sand filtration. Water Science and Technology 50 (1), 147e154.
Jin, Y., Flury, M., 2002. Fate and transport of viruses in porous
media. Advances in Agronomy 77, 39e102.
Lance, J.C., Gerba, C.P., Melnick, J.L., 1976. Virus movement in soil
columns flooded with secondary sewage effluent. Applied and
Environmental Microbiology 32 (4), 520e526.
Lance, J., Gerba, C., Wang, D., 1982. Comparative movement of
different enteroviruses in soil columns. Journal of
Environmental Quality 11 (3), 347e351.
Landry, E.F., Vaughn, J.M., Thomas, M.Z., Beckwith, C.A., 1979.
Adsorption of enteroviruses to soil cores and their subsequent
elution by artificial rainwater. Applied and Environmental
Microbiology 38 (4), 680e687.
Landry, E.F., Vaughn, J.M., Penello, W.F., 1980. Poliovirus retention
in 75-cm soil cores after sewage and rainwater application.
Applied and Environmental Microbiology 40 (6), 1032e1038.
Medema, G.J., Stuyfzand, P.J., 2002. Removal of microorganisms
upon basin recharge, deep well injection and river bank
filtration in the Netherlands. In: Proceedings of the 4th
International Symposium on Artificial Recharge Adelaide,
Australia, Sep 22e26.
Moe, C., Rheingans, R., 2006. Global challenges in water, sanitation
and health. Journal of Water and Health (Suppl. 04), 41e57.
Quanrud, D.M., Carroll, S.M., Gerba, C.P., Arnold, R.G., 2003. Virus
removal during simulated soil-aquifer treatment. Water
Research 37 (4), 753e762.
Ray, C., Soong, T., Lian, Y., Roadcap, G., 2002. Effect of floodinduced chemical load on filtrate quality at bank filtration
sites. Journal of Hydrology 266 (3e4), 235e258.
Ray, C.J., Schubert, P.E., Linsky, R.B., Melin, G., 2003. What is
riverbank filtration? Historical significance. Unrecognized RBF
plants. Similarities between RBF and slow sand, filtration.
Surface-water contaminants of concern. case studies of log
removal credit in the United States. The value of applying RBF
as a pretreatment technology. In: Ray, C., Melin, G., Linsky, R.B.
(Eds.), River Bank Filtration. Improving Source-water Quality.
Series: Water Science and Technology Library, vol. 43. IWA
Publishing, London, UK.

452

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 4 3 9 e4 5 2

Schijven, J., Hassanizadeh, S., 2000. Removal of viruses by soil


passage: overview of modeling, processes, and parameters.
Critical Reviews in Environmental Science and Technology
30 (1), 49e127.
Schijven, J.F., Hoogenboezem, W., Hassanizadeh, S.M., Peters, J.H.,
1999. Modelling removal of bacteriophages MS2 and PRD1 by
dune infiltration at Castricum, The Netherlands. Water
Resources Research 35 (4), 1101e1111.
Schijven, J.F., Medema, G., Vogelaar, A.J., Hassanizadeh, S.M.,
2000. Removal of microorganisms by deep well injection.
Journal of Contaminant Hydrology 44, 301e327.
Schijven, J., 2001. Virus Removal from Groundwater by Soil
Passage: Modeling, Field and Laboratory Experiments. Ponsen
& Looijen B. V, Wageningen.
Schmidt, P.J., Emelko, M.B., Reilly, P.M., 2010. Quantification of
analytical recovery in particle and microorganism
enumeration methods. Environmental Science & Technology
44 (5), 1705e1712.
Shamrukh, M., Abdel-Wahab, A., 2008. Riverbank filtration for
sustainable water supply: application to a large-scale facility
on the Nile River. Clean Technologies and Environmental
Policy 10 (4), 351e358.
Szewzyk, R., Szewzyk, U., 2009. Personal Communication.
Regine Szewzyk, Umwelbundesamt, Berlin. regine.szewzyk@
uba.de. Ulrich Szewzyk. Technical University, Berlin. www.
tu-berlin.de.
Tyrrell, S.A., Rippey, S.R., Watkins, W.D., 1995. Inactivation of
bacterial and viral indicators in secondary sewage effluents,
using chlorine and ozone. Water Research 29 (11), 2483e2490.

Unger, M., Collins, M.R., 2006. Assessing the role of the


schmutzdecke in pathogen removal in riverbank and slow
sand filtration. In: Gimbel, R., Graham, N.J.D., Collins, M.R.
(Eds.), Recent Progress in Slow Sand and Alternative
Biofiltration Processes. IWA publishing, London, UK.
USEPA-1602, 2001. Method 1602: Male-Specific (F) and Somatic
Coliphages in Water by Single Agar Layer (SAL) Procedure EPA821-R-01-029, Washington, USA.
van der Wielen, P.W.J.J., Senden, W.J.M.K., Medema, G., 2008.
Removal of bacteriophages MS2 and FX174 during transport in
a sandy anoxic aquifer. Environmental Science & Technology
42 (12), 4589e4594.
Van-Heerden, J., Ehlers, M., Zyl, W., Grabow, W., 2004. Prevalence
of human adenoviruses in raw and treated water. Water
Science and Technology 50 (1), 39e43.
Vaughn, J., Landry, E., Beckwith, C., McHarrel, Z., 1981. Virus
removal during groundwater recharge: effects of infiltration
rate on adsorption of poliovirus to soil. Applied and
Environmental Microbiology 41 (1), 139e147.
WHO, 2006a. Viral pathogens. In: Guidelines for Drinking Water
Quality, 2006. WHO, Geneva, pp. 247e259.
WHO, 2006b. System assessment and design. In: Guidelines for
Drinking Water Quality, 2006. WHO, Geneva, pp. 51e68.
WHO, 2006c. Occurence and treatment of pathogens. In:
Guidelines for Drinking Water Quality, 2006. WHO, Geneva,
pp. 135e141.
Woessner, W.W., Ball, P.N., DeBorde, D.C., Troy, T.L., 2001. Viral
transport in a sand and gravel aquifer under field pumping
conditions. Groundwater 39 (6), 886e894.

Вам также может понравиться