You are on page 1of 15

ELECTROCHEMICAL AND COPPER DISINFECTION WITH MINIMAL

CHLORINATION: A TWO-YEAR CASE STUDY IN AN INTENSE BATHER


LOADED PUBLIC TRAINING POOL

P. Lievens*, R. Calders**, R. Vanlangendonck*, J. Van den Bulcke*, A. Jobse*

*E-clear.eu, Halensbroek 1114, Mosstraat 38a, 3545 Halen, Belgium


(E-mail: lievens@e-clear.eu)
**P.I.H., Kronenbrugstraat 45, 2000 Antwerpen, Belgium
(E-mail: rudy.calders@provincieantwerpen.be)

Abstract
A two-year case study in a training pool treated with three disinfecting mechanisms;
chlorination, electrochemical and copper disinfection showed that it was possible to
work with chlorine levels as low as 0.1 ppm FAC. No fecal indicator was found
during the whole test. The chlorine demand was proportional to the bather load and
decreased with lower FAC to 4.3 g Cl2/bather. The net chlorine demand decreased
spectacular. The free chlorine disappeared almost completely in the dual-media filter,
which represented 4.8 L/day chlorine use. The process maximizes the effectiveness
of chlorine so that low levels achieve disinfection equivalent to that normally offered
by high levels, while minimizing the drawbacks; lower combined chlorine, AOX,
THM and chlorate.

Keywords
public pool, copper disinfection, electrochemical disinfection, chlorination

INTRODUCTION.
Traditionally, chlorine and chlorine compounds have been used to disinfect swimming pool
waters. Free available chlorine (FAC) levels of 1.0 parts per million (ppm) or greater are
usually maintained to ensure effective control of microorganisms and to make the general
sanitary quality of swimming pool waters acceptable. Although a very effective disinfectant,
chlorine has several disadvantages. The lifetime of the FAC residual varies with climatic
conditions and bather load (chlorine demand). When chlorine reacts with organic compounds
and nitrogen compounds present in the water, it forms objectionable by-products such as
trihalomethanes (THMs) and chloramines. FAC levels of 0.6 ppm or greater have been
associated with irritation of eyes, nasal passages, and skin, as well as with objectionable odors
in the pool environment. Also, chlorine has a corrosive effect on pool structures, necessitating
expensive maintenance work (Borkow, 2005). The discharge of DBPs (Disinfection By-
Products) such as AOX is restricted in several countries. A greater number (100) of DBPs
were identified in the indoor pools, including most classes of DBPs found in chlorinated
drinking water, but also nitrogen-containing DBPs (“N-DBPs”) such as haloamides,
halonitriles, haloanilines, haloamines, haloanisoles, and halonitro- compounds. N-DBPs are
likely present in pool waters due to nitrogen-containing precursors from swimmers (urine,
sweat, etc.) (LaKind, 2010)

Electrochemical disinfection
Microorganisms can be electrochemically inactivated either directly or via the generation of
very active chemical species, such as free radicals and other ions (OH·, [O], HO 2·, Cl2, OCl-,
etc). The disinfecting properties are due to very unstable and short-lived active oxygen
compounds produced at the electrodes (Pulido, 2005) (Drees, 2003) (Patermarakis, 1990).
The hydroxyl radicals are highly reactive, non-selective electrophilic reagents that serve as a
precursor to the oxidative breakdown of organic molecules, such as halogenated hydrocarbons
into smaller, less toxic organic compounds and mineralized acids. Indeed, advanced oxidation
processes (AOP) are a class of emerging technologies in drinking water treatment that utilizes
hydroxyl radicals to oxidize DBP precursors. (Pulido, 2005)
Electric fields are also themselves harmful to cells. This is due to the electrochemical
oxidation of intracellular Coenzyme A (CoA), which leads to decreased respiration and
consequent cell death (Matsunaga, 1992). The carbonate and sulphate ions present in the
water could be oxidized at the anode to form percarbonate or persulphate, which are excellent
oxidizing agents for bacteria (Porta, 1986). Electrochemical disinfection is most effective in
waters containing chloride, but E. coli and bacteriophage MS2 are inactivated by 4 logs in
electrochemically treated sulphate and phosphate electrolytes in the absence of chloride. The
absence of a residual disinfectant in electrochemically treated waters without chloride infers
the mechanism of disinfection is by the generation of short-lived oxidants (Kerwick, 2005).
But Delaedt et al. reports that in all treatments where E. coli bacteria (104 CFU/ ml) were
added to water tapped downstream the electrolysis unit (residual effect), no living bacteria
could be detected by the end of the reaction time. They determined (DPD method for
combined chlorine) the concentration of free oxidants before the addition ranged from 0.27 up
to 0.49 mg/l and the concentration at the end of the reaction time varied between 0.08 and
0.33 mg/l (Delaedt, 2008).
Numerous aplications of electrochemical disinfection have been described; in agriculture &
food industries (Al-Haq, 2005), as alternative for chlorination of wastewater effluents (Pulido,
2005), in aquaculture, ballast water e.g. (Tsolaki, 2010), cooling water treatment, drinking
water treatment (Ghernaout, 2010), clinical, dental and veterinary applications. Numerous
patents and laboratory studies have been conducted. However, only a few electrochemical
water disinfection products are currently available on the market. This is due to the relative
unfamiliarity of the technology, and to fierce market competition with other technologies.
Eventually, the cost and performance advantages of electrochemical technology should lead
to its wider use (Kraft, 2008).
Only a few applications for disinfection of swimming pool water are described see (Hafer,
1995)(small field study testing a commercial apparatus) (Lee, 2010).

Copper disinfection
Results reported by several authors (Gerba, 1989) (Abad, 1994) (Beer, 1997) indicate that the
use of 300 to 400 ppb of electrolyticaly generated copper and 40 ppb of silver combined with
0.1 to 0.4 ppm of chlorine is more effective than higher levels of chlorine in limiting a number
of microorganisms, including coliforms. These studies suggest a synergistic effect upon
microorganisms subjected to copper or silver ions in the presence of low levels of chlorine.
The technology provides a high level of bacteria control, while lower chlorine levels result in
substantial reduction of the production of trihalomethanes and increase substantially the
enjoyment and satisfaction of the swimming experience for the pool user. (Beer, 1997)
In contrast to the low sensitivity of human tissue to copper, microorganisms are extremely
susceptible to copper. Copper toxicity in microorganisms may occur through the displacement
of essential metals from their native binding sites, from interference with oxidative
phosphorylation and osmotic balance, and from alterations in the conformational structure of
nucleic acids, membranes and proteins. In most microorganisms, but not in viruses, there is an
integrated set of proteins that delivers copper to specific subcellular compartments and
copper-containing proteins without releasing free copper ions. Although some organisms have
mechanisms of resistance to excess copper, generally exposure of most microorganisms to
high concentrations of this trace element results in damage to cellular components. Viruses
lack DNA repair mechanisms, permeability barriers, intra- and extra-cellular sequestration of
metals by cell envelopes, active metal transport membrane efflux pumps, and enzymatic metal
detoxification mechanisms, such as those found in bacteria and cells. The reduced capabilities
of viruses to resist copper may thus explain their high vulnerability to the metal. (Borkow,
2005) (Meyer, 2001)

Effect of pH
The solubility of copper decreases with increasing pH (Yu-sen, 2002). The rate of disinfection
potential by chlorination varies with pH because of the change in the ratio HOCl/ClO -.
(Randtke, 1992) At pH 7 there is about 75 % present as hypochlorous acid and at pH 8 only
25 %. It has been estimated that hypochlorous acid is almost 100 times more effective for
disinfection than hypochlorite ion. However, it should be remembered that ClO - is a reservoir
of HOCl. As HOCl is consumed it is replenished by the equilibrium reaction: ClO - + H+ =
HOCl. (Wojtowicz, 2001). THM formation was reduced by decreasing pH but HAN, and
NCl3 formation increased at decreasing pH whereas the formation of HAAs remained constant
(Hansen, 2012). The electrochemical disinfection is sensitive for fouling at higher pH because
of possible carbonate deposits. Lower pH favours the disinfection potential of the three
methods, but is limited by the aggresiveness of the water. For all this reasons a pH of 7.0 was
maintained during the tests.
Combined disinfection
Regulatory and end users pressure to achieve high inactivation for various pathogens in water
has pushed the industry towards more effective disinfectants.
The process where two (or more) disinfectants produce a synergistic effect by either
simultaneous or sequential application to achieve more effective pathogen inactivation is
referred to as interactive disinfection and it is still considered an emerging technology.
(Pulido, 2005; Yahya, 1990)
By combining three disinfecting mechanisms; chlorination, electrochemical and copper
disinfection, the authors believed that:
Drawbacks of one method are not prevalent
Advantages of each method are exploited
Resistance is suppressed
Synergies are possible: research has shown that the application of sequential
disinfectants is more effective than the added effect of the individual disinfectants.
The combination should allow to work at minimal chlorination levels and therefore with
minimal DBP formation. AOX and THM where monitored to prove this thesis.
Legislation and variance
VLAREM II (Flemish legislation) requires swimming pool waters to be disinfected with
chlorine at a FAC of 0.5 to 1.5 ppm. Any alternative disinfection method that uses lower
FAC levels requires a special protocol. A test protocol was written by the Flemish agency for
care and health (VAZG) for a two-year use of alternative disinfection at the training pool in
Leuven (Flanders, Belgium). The protocol required two-weekly sampling and analysis of
physical, chemical and bacteriological parameters as defined in the VLAREM II and the level
of copper, the number of E. coli and intestinal enterococs (see table 2). No criterion for
acceptance of the alternative disinfection method was proposed. But the agency promised that
it would give positive advice in the environmental permit procedure that grants a variance of
the existing legislation.
MATERIAL EN METHODS.
Training pool
With dimensions of 10 m * 20 m * 2 m the pool had a volume of 400 m 3 and is completely
separated from the other pools in the facility “Sportoase”. Only schools and clubs are allowed.
The pool is closed during winter-, spring- and summerholidays.

Fig. 1. General scheme


Figure 1 shows the general scheme of the installation. The water was pumped to the filter
surface of 7.32 m2 at 124 m3/h after adding PAC and correcting the pH. About 1.6 g/h
aluminum hydroxichloride 13 w/w % was dosed. The dual-media filter is composed of 80 cm
sand and 40 cm anthracite. Backwashing during 4.5 minutes en rinsing during 3 minutes took
place twice a week. The water was heated by a 120 kW heat exchanger. The pool was
chlorinated by dosing a sodium hypochlorite solution of 12.5 w/w% (12 w/w% active
chlorine, specific gravity 1.2 kg/L, Roam chemie). The acid dosage was averaged 1.86 L
sulfuric acid (40 w/w %, specific gravity 1.31 kg/L, Roam chemie) per day.

The pool water quality (8/2/2012-8/20/2012, lab Derva) and operational parameter values at
the start of the project are summarized in table 1.

Table 1. Pool water quality and operational parameter values


Water quality
Total alkalinity (mg HCO3-/L) 34
Free chlorine residual (mg/L “as Cl2”) 1.03
Combined chlorine residual (mg/L “as Cl2”) 0.31
Conductivity (mS/cm) 2.58
pH 7.12
Temperature (°C ) 30
Totale hardness (mg CaCO3/L) 310
Turbidity (NTU) 0.09
Langelier index (LSI)
Nitrate ( mg NO-3/L) 69
Sulfate (mg SO42-/L) 277
Chloride (mg Cl-/L) 660
Table 2. Analytical methods for (physico)chemical parameters
Chlorate (mg ClO3-/L) 36
KMnO4 (mg O2/L) 0.83
Ureum (mg/L) < 0.5

Pool operation
Water flow rate (m3/h) 124
Water turnover (h) 4
Filter surface (m2) 7.32
Filtration velocity (m3/u/ m2) 17
Bather load (/day) 300
Maximum (/hour) 65
Mean (/hour) 30
Chlorine consumption (kg per month 47/50 12,5% sodiumhypochlorite) < 600
Acid consumption (kg sulfuric acid 40% per month) < 40
Make-up water flow rate (m3/day/person) 32

E-clear.eu
On 14/8/2012 E-Clear.eu installed 1588 cm2 iridium coated titanium electrodes inline, directly
after filtration, operating at 1.26 mA/cm 2, after the filter and before the injection of sodium
hypochlorite (see fig. 1). Electrodes were regenerated after fouling by reversing the electrode
polarity every half hour.
Furthermore a constant amount of copper was added to the water by electrolysis of 710 cm 2
copper electrodes after filtration, before chlorination. The current was regulated to obtain 0.3
ppm Cu in the pool water, as weekly measured by staff people.
Analytical methods
Table 2 shows the analytical methods used by two different certified labs for several
(physico)chemical parameters measured during the course of the case study.

Lab Parameter Method


DERVA Chloride, nitrate, sulfate Ionchromatography, ISO 10304-1
DERVA, PIH Turbidity Nefelometric method ISO 7027
DERVA, PIH Urea Enzymatic, spectrophotometry,
NEN6494
DERVA, PIH Oxidisability Titrimetry
DERVA, PIH Conductivity at 20°C ISO 7888
DERVA, PIH Bicarbonate, alkalinity Titrimetry, ISO 9963-1
DERVA, PIH Combined chlorine Colorimetry, ISO 7393/2
DERVA, PIH Free available chlorine Colorimetry, ISO 7393/2, SM§4500/Cl
G
DERVA Chlorate Ionchromatography, ISO 10304-4
PIH Copper ICP-AES
PIH NPOC Catalytic oxidation and IR
PIH AOX Coulometry
PIH Trihalomethanes VOC determination with HS/GCMS

The two certified labs also controlled following microbiological parameters, as shown in
table 3.
Table 3. Analytical methods for bacteriological parameters

Lab Parameter Method


DERVA, PIH Colony count plate count agar ISO 6222
DERVA, PIH Coliform bacteria tergitol-7 agar ISO 9308-1 + ISO
9308-2
DERVA, PIH Fecal Coliform tergitol-7 agar ISO 9308-1 + ISO
9308-2
DERVA, PIH Intestinal Enterococci slanetz and bartley agar ISO 7899-2
DERVA, PIH Pathogenic Stafylococci Baird-Parker SM 9213, XPT90-412
DERVA, PIH Pseudomonas aeruginosa ISO 16266 , EN12780
DERVA, PIH Fungi and yeasts sabouraud dextrose agar

The chlorine level in the pool was regulated by an amperometric sensor for total available
chlorine, type CGE 2 (0,01-10 mg/L) (calibrated against Lovibond DPD method) from
Prominent in a Dulcomarin unit, which also controlled pH and measured ORP.

RESULTS AND DISCUSSION.


Startup
The set point pH was lowered gradually from 7.25 to 7.0 during the first weeks of the project.
Simultaneously the copper concentration rose to 0.4 ppm. The set point value for free chlorine
was decreased slowly from 1.2 to 0.5. Figure 2 shows the evolution of those parameters. The
photometrically (DPD metod) determined combined chlorine increases a little. The average
value of ORP was 844 mV during startup.

Fig. 2. Gradual change of the chemical parameters during startup


End phase results
(Physico)chemical parameters
pH, chlorine, AOX, THM and copper.
The free chlorine was further slowly decreased in the end phase to a value of 0.2 ppm
chlorine, as can be seen from table 4. The copper concentration fluctuated around 250 37
ppb.
Table 4. Chemical parameters in the end phase as determined by two monitoring labs

Date pH ppm ppm ppb ppb ppb


Free Combined AOX THM Copper
chlorine chlorine
25/09/13 7.18 0.54 0.26
26/09/13 7.10 0.60 0.40 280 22 174
1/10/13 7.36 0.96 0.10 194
15/10/13 7.10 0.31 0.41 190 21 190
14/11/13 7.22 0.53 0.17 220
20/11/13 7.00 0.34 0.43 180 21 253
4/12/13 7.10 0.41 0.35 190 19 247
23/12/13 7.11 0.44 0.25 270
8/01/14 7.10 0.44 0.32 250 12 276
16/01/14 7.03 0.35 0.28 310
5/02/14 7.10 0.44 0.40 270 22 306
19/02/14 7.23 0.52 0.34 300
5/03/14 7.00 0.32 0.35 180 16 282
18/03/14 7.01 0.30 0.06 230
2/04/14 6.90 0.26 0.37 180 16 246
17/04/14 6.89 0.37 0.24 250
22/05/14 6.95 0.28 0.14 230
23/05/14 7.1 0.23 0.41 254
11/06/14 7.1 0.26 0.29 232
19/06/14 7.02 0.1 0.22 260
2/07/14 7 0.25 0.24 269
24/07/14 7,15 0,17 0,13 188
7/08/14 7,1 0,34 0,16 199
21/08/14 7,06 0,26 0,14 210

The results indicated that the process was able to minimize production of DBP’s. Normal
averages are situated around 300-600 ppb AOX and 70-170 ppb THM in the province of
Antwerp, Belgium (R. Calders, personal communication).
The concentration of AOX was measured in a number of public pools from Sportfondsen
Nederland NV. The figures ranged from 400 to 2000 µg/L. Filtration over activated carbon
had no or little effect (Vlaardingerbroek, 2007). The concentrations of chloroform collected in
eight different indoor swimming pools in London showed a geometric mean concentration of
total THMs (TTHMs) 132.4 μg/L and for chloroform 113.3 μg/L (Chu, 2002). An average
concentration of 121.2 μg/L was found in samples collected at the swimming pool water in
the Town of Brookline, Massachusetts' Municipal Swimming Pool Recreational Facility,
working at 1.0 ppm FAC. After lowering the FAC to 0.4-0.5 ppm and electrolitically dosing
0.3 ppm copper/silver an average concentration of 48.5 μg /L was found. This represented a
150% reduction in THM concentrations. But other studies found much lower concentrations
in their swimming pool studies, e.g. mean CHCl3 water from 10 to 80 μg/L (LaKind, 2010).

Conductivity, turbidity, redoxpotential, chloride, chlorate.


Conductivity values during end phase fluctuated around 1500 µS/cm and turbidity was always
about 0.13 NTU. The average value of ORP was 803 mV. Chloride and chlorate were halved,
compared to startup, with end values of resp. 339 ± 25 and 20 ± 6 mg/L. This means that also
the objectionable chlorate concentration is decreasing with lower FAC. Average concentration
of chlorate in poolwater treated with sodium hypochlorite was 17 mg/L (Erdinger, 1999)
(Ribeiro, 2011) - 30 mg/L (Michalski, 2007). Germany allows <30 mg/L in swimming pools
and the provisional guideline value for drinking-water of the World Health Organization is
0,7 mg /L (Guidelines for drinking-water quality).

Alkalinity
The alkalinity of the pool was always to low compared to legal standards and good practice.
At the startup the figures averaged 34.2 mg bicarbonate/L and lowered to 21.7 mg/L in the
end phase, while legally > 60 mg/L is needed. Addition of 25 kg NaHCO3 every two days did
not cure the situation. Former tests showed that CO2 neutralisation demanded a very high
amount of CO2. Although Gomà et al. (Gomà, 2010) showed that with 4 kg/m3,year it is
possible to maintain 137 mg bicarbonate/L. This could potentially adjust the alkalinity,
thereby converting the swimming pool in a depository of greenhouse gas, reducing the carbon
footprint and lowering the chlorine use and the formation of THM.

Nitrate, sulfate, urea, DOC, KMnO4


The nitrate concentration lowered to 42 ± 3 ppm, and the sulfate to 205 ± 17 ppm. Urea
values rose a little to 0.7 ppm. KMnO4 consumption stayed < 1 mg O2/L while DOC was
around 2.8 mg/L.

Bacteriological parameters
At the same time, the two certified labs took samples of the poolwater and determined the
bacteriological quality. The results are summarized in table 5. Total plate count was lower
than 100 cfu/ml. Other research for fecal indicators (Pseudomonas, Staphylococs, E. coli,…)
showed that no fecal indicator (cfu/100 ml) was found during the whole test. The
bacteriological tests indicated that the disinfection procedure worked very well, even at 0.1
ppm free chlorine (19/06/14).

Table 5. Bacteriological parameters in the end phase as determined by two monitoring labs
Lab Date
Total plate count (cfu/ml)

Fungi and yeasts

Pseudomonas

Stafylococci
Enterococci
Coliforms

fecal Coli

DERVA 25/09/13 0 0 0 0 0 0 0
PIH 26/09/13 0 0 0 0 0
DERVA 1/10/13 0 0 0 0 0 0 0
PIH 15/10/13 0 0 0 0 0
DERVA 14/11/13 61 0 0 0 0 0 0
PIH 20/11/13 1 0 0 0 0
PIH 4/12/13 0 0 0 0 0
DERVA 23/12/13 1 0 0 0 0 0 0
DERVA 16/01/14 0 0 0 0 0 0 0
VAZG 30/01/14 8 0 2
PIH 5/02/14 0 0 0 0 0
DERVA 19/02/14 2 0 0 0 0 0 0
VAZG 24/02/14 5 0 0
PIH 5/03/14 7 0 0 0 0
VAZG 10/03/14 0 0 0
DERVA 18/03/14 0 0 0 0 0 0 0
DERVA 2/04/14 0 0 0 0 0
PIH 17/04/14 0 0 0 0 0
DERVA 22/05/14 0 0 0 0 0 0 0
PIH 23/05/14 0 0 0
PIH 11/06/14 6 0 0 0 0
DERVA 19/06/14 0 0 0 0 0 0 0
PIH 2/07/14 0 0 0 0
DERVA 24/07/14 0 0 0 0 0 0 0
PIH 7/08/14 0 0 0 0
DERVA 21/08/14 0 0 0 0 0 0 0

Bather load evolution


The bather load, expressed as the number of visitors per day, increased gradually during the
project as can be seen from figure 3. The lower attendance during holidays is obvious.

Fig. 3. Evolution of bather load (in average number of visitors/day) during the project
Chlorine demand
Background consumption of chlorine.
The use of sodium hypochlorite 12.5 w/w % in liter per day was proportional to the number of
visitors, as can be seen from figure 4.

Fig. 4. Use of sodium hypochlorite (liter 12 %/day) in relation to the number of visitors

Extrapolation of this data showed a background consumption of about 8 liter sodium


hypochlorite 12.5 w/w% per day in an unloaded pool. After subtraction of the background
consumption, a “net chlorine consumption per 1000 bathers” was obtained. The background
consumption is due to (photo)catalytic decay to chloride in the pool (e.g. catalysed by copper
ions), slow but continuous release of oxygen gas through a decomposition process, reaction
with oxydisable matter and chemical reactions in the dual-media filter. Closed bottle tests
showed a complete degradation of residual active chlorine in the pool water in 24 hours. To
test the chlorine decay in the filter, DPD measurements of free available chlorine residual
before and after filtration were determined as shown in table 6.
Table 6. Change in FAC after filtration
Date Free available chlorine ∆ FAC
Before After
5/2/14 0.45 0.18 0.27
13/2/14 0.54 0.33 0.21
17/4/14 0.26 0 0.26
2/5/14 0.34 0.05 0.29
23/5/14 0.23 0 0.23
11/6/14 0.26 0.08 0.18
2/7/14 0.25 0.09 0.16

From those figures and the water flow rate a chlorine use of 686 g chlorine per day was
calculated. This equals a consumption of averaged 4,8 liter sodium hypochlorite 12.5 w/w%
per day. This could have been the result of a reaction with organic carbon (e.g. in the
anthracite layer) (Potwora, 2009). The combined chlorine did not change significantly in the
filter. Free chlorine reacts rapidly with both powdered activated carbon (PAC) and granular
activated carbon (GAC). The reaction is heterogeneous, involving both solid and liquid
phases, and can be expressed as follows:
C(s) + HOCl CO(s) + H+ + Cl−
If significant amounts of HOCl are allowed to react with the carbon, some of the oxygen
attached to the surface may be emitted as CO or CO2 gas. The stoichiometry of the reaction
will be as follows:
C + 2Cl2 + 2H2O 4HCl + CO2
In this reaction, 1.0 part of chlorine will destroy 0.00845 part of carbon.
The surface of the carbon is oxidized, chlorine is reduced to chloride, and acidity is produced.
If the chlorine is present in the form of HOCl, removal of 1 mg/l of residual chlorine as Cl 2
consumes 0.7 mg/l of alkalinity as CaCO3.
Removal of combined chlorine occurs much more slowly than removal of free chlorine and
the reaction can be expressed as follows:
C(s) + NH2Cl + H2O CO(s) + NH+4 + Cl−
CO(s) + 2NH2Cl C(s) + N2 + 2 H++ 2Cl− + H2O
C(s) + 2NHCl2 + H2O CO(s) + N2 + 4 H++ 4Cl− (Randtke, 1992)
Vogel et al. (Vogel, 2009) reported about a dual-media filter media with anthracite-sand that
removed chlorine quickly from chlorinated water. They concluded that virgin anthracite
became coated with carbon-oxide that prevented futher interaction. The filter media in our
case were however over eight years old and still removed free chlorine.
Chlorine demand and free chlorine.
There was a correlation between the average FAC and the calculated “net chlorine
consumption per 1000 bathers” as shown in figure 5. This means that a lower threshold
decreased the use of chlorinating agent. This is also visible in figure 6.

Fig. 5. Correlation between the average FAC and the calculated “net chlorine consumption per 1000 bathers”

Figure 6 represents the decreasing total chlorine use per 1000 bathers vs. the average FAC
data chronologically. Higher residual chlorine levels generate significant higher chlorine
usage. White (White, 2000) expects chlorine use to rise as a square of the free-chlorine level.
Fig. 6. Chronologic representation of the average FAC and the total chlorine consumption per 1000 bathers
The extreme values at the end of 2013 were due to shock chlorination during a holiday period.
Acid demand
The 12.5 w/w% sodium hypochlorite solution (2.0 N) could be theoretically neutralised by 1.0
mole of sulfuric acid per liter sodium hypochlorite solution. Thus 0.19 L of the 40 % sulfuric
acid solution (10.7 N) is needed to neutralise one liter of sodium hypochlorite solution.
The average consumption of sodium hypochlorite solution per day was 14 L whereof the
dual-media filter destroyed 4.8 L, the net amount to neutralise was therefore 9.2 L chlorine.
This could be theoretically neutralised with 1.7 L sulfuric acid. But the sodium hypochlorite
solution contains a certain amount of sodiumhydroxide and sodiumcarbonate. This explains
the real consumption of about 1.9 liter sulfuric acid per day.
Airquality
Up till now, only subjective experiences of the air quality by the lifeguards were done. They
confirmed that air quality was much better in the training pool area than in the recreational
pool space.

SUMMARY.
A two-year case study in a training pool treated with three disinfecting mechanisms;
chlorination, electrochemical and copper disinfection showed that it was possible to work
with chlorine levels as low as 0.1 ppm FAC. The bather load was continuously rising from
250 – 350 visitors/day, or 0.6 – 0.9 bathers.m-3.h-1, while lowering the FAC, maintaining 0.3
ppm electrolyticaly generated copper and < 0,2 W/m 3 flow rate gentle electrochemical
desinfection by water electrolysis at pH 7.
The experiment was followed by the health inspection agency every two weeks by a complete
report with analysis and operational data. Two certified labs performed complete chemical
and microbiological follow up. The lowest measured value for FAC was 0.1 ppm with
combined chlorine of 0.22 ppm. Chlorine by-products AOX decreased to 180 µg/L and total
THM to 16 µg/L. The chlorate content was substantially lower at lower chlorine levels.
Colony counts were absent and fecal indicators research (Pseudomonas, Staphylococs, E. coli,
…) showed that no fecal indicator was found during the whole test. The bacteriological results
were outstanding, pointing towards a very effective disinfection. The chlorine demand was
proportional to the bather load and decreased with lower FAC to 30 L per 1000 bathers. This
corresponds to 4.3 g Cl2/bather. An average chlorine demand of 14 L/day was partly due to
background use of 8 L/day (the chlorine demand of the unloaded pool). The free chlorine
disappeared almost completely in the dual-media filter, which represented 4.8 L/day chlorine
use. The net chlorine demand decreased spectacular. The drawbacks of chlorination could be
minimalised. Lower residual chlorine levels generated reduced chlorine usage and DBP’s
(chlorine by-products), indicated by lower combined chlorine, AOX, THM and chlorate. The
process maximizes the effectiveness of chlorine so that low levels achieve disinfection
equivalent to that normally offered by high levels, while minimizing the drawbacks. Air
quality was experienced by the lifeguards as substantially improved. Determination of
nitrogen trichloride (NCl3) levels in the air with the impinger method (Predieri, 2012) are
planned for the near future.

REFERENCES
Abad, F. P. (1994, July). Disinfection of human enteric viruses in water by copper and silver
in combination with low levels of chlorine. Appl. Environ. Microbiol. , 2377-2383.
Al-Haq, M. S. (2005). Applications of electrolyzed water in agriculture & food industries.
Applications of electrolyzed water in agriculture & food industries , 11 (2), 135-150.
Beer, C. G. (1997). Swimming pool disinfection efficacy of Copper/Silver ions with reduced
chlorine levels. Indian Journal of Environmental Health. , 39 (3), 9-13.
Borkow, G. a. (2005). Copper as biocidal tool. Current Medicinal Chemistry , 12, 2163-2175.
Chu, H. (2002). Distribution and determinants of trihalomethane concentrations in indoor
swimming pools. Occup Environ Med , 59, 243-247.
Delaedt, Y. D. (2008). The impact of electrochemical disinfection on Escherichia coli and
Legionella pneumophila in tap water. Microbiological Research , 163, 192-199.
Drees, K. A. (2003). Comparative electrochemical inactivation of bacteria and bacteriophage.
Water Res. , 37, 2291-2300.
Erdinger, L. K.-G. (1999). Chlorate as an inorganic disinfection by product in swimming
pools. Zbl Hyg Umweltmed , 202, 61-75.
Gerba, C. K. (1989). Swimming Pool Disinfection, An Evaluation of the Efficacy of Copper:
Silver Ions. Journal of Environmental Health , 51 (5), 282-285.
Ghernaout, D. a. (2010). From chemical disinfection to electrodisinfection: The obligatory
itinerary? Desalination and Water Treatment , 16 (1-3), 156-175.
Gomà, A. G. (2010). Benefits of carbon dioxide as pH reducer in chlorinated indoor
swimming pools. Chemosphere , 80, 428-432.
Guidelines for drinking-water quality. WHO.
Hafer, D. (1995). A Field Evaluation of the Bi–Polar Oxygen Sanitation System/Mineral
Purification System. Journal of the Swimming Pool and Spa Industry , 1 (3), 39-51.
Hansen, K. (2012). Effect of pH on the formation of disinfection byproducts in swimming
pool water - Is less THM better? Water Research , 46, 6399-6409.
Kerwick, M. R. (2005). Electrochemical disinfection, an environmentally acceptable method
of drinking water disinfection? Electrochimica Acta , 50(25-26), 5270-5277.
Kraft, A. (2008). Electrochemical water disinfection: a short review. Platinum Metals , 52 (3),
177-185.
LaKind, J. R. (2010). The good, the bad, and the volatile: can we have both healthy pools and
healthy people? Environmental science & technology , 44 (9), 3205-3210.
Lee, J. (2010). Production of various disinfection byproducts in indoor swimming pool
waters treated with different disinfection methods. International Journal of Hygiene and
Environmental Health , 213 (6), 465-474.
Matsunaga, T. N. (1992). Disinfection of drinking water by using a novel electrochemical
reactor employing carbon-cloth electrodes. Appl. Environ. Microbiol. , 58, 6, 686-689.
Meyer, W. C. (2001). Resistance to Copper-Silver Disinfection. Water Engineering &
Management , 25-27.
Michalski, R. (2007). Occurrence of chlorite, chlorate and bromate in disinfected swimming
pool water. Polish Journal of Environmental Studies , 237-241.
Patermarakis, G. (1990). Disinfection of water by electrochemical treatment. Water research ,
24 (12), 1491-1496.
Porta, A. (1986). Patent No. 4,619,745. U.S.
Potwora, R. (2009). Chlorine and Chloramine Removal with Activated Carbon. Water
Conditioning & Purification , 14-16.
Predieri, G. (2012). Determination of nitrogen trichloride (NCl3) levels in the air of indoor
chlorinated swimming pools: an impinger method proposal. International Journal of
Environmental Analytical Chemistry , 92, 6, 645-654.
Pulido, M. (2005). Evaluation of an Electro-Disinfection Technology as an Alternative to
Chlorination of Municipal Wastewater Effluents.
Randtke, S. (1992). Chemistry of Aqueous Chlorine. In G. White, White's handbook of
chlorination and alternative disinfectants. (5th edition ed., Vol. 2, pp. 145-146). Van
Nostrand Reinhold, New York.
Ribeiro, I. H. (2011). Occurrence of chlorite and chlorate in swimming pool water. Fourth
International Conference Swimming Pool and Spa , 15-18 march 2011. Porto, Portugal.
Tsolaki, E. P. (2010). Electrochemical disinfection of simulated ballast water using Artemia
salina as indicator. Chemical engineering journal , 156 (2), 305-312.
Vlaardingerbroek, A. v. (2007). Oriënterend onderzoek naar desinfectietechnieken voor
zwembadwater. Kiwa Water Research. KIWA.
Vogel, J. K.-K. (2009). Meeting Chlorine Demand for Virgin Anthracite. Opflow , 35, 12, 16-
19.
White, T. (2000). Several questions and comments in response to article in Journal of
Environmental Health, May 1999. Journal of Environmental Health , 1-10.
Wojtowicz, J. (2001). Survey of Swimming Pool/Spa Sanitizers and Sanitation Systems.
Journal of the Swimming Pool and Spa Industry , 4:11, 9-29.
Yahya, M. T. (1990). Disinfection of bacteria in water systems by using electrolytically
generated copper:silver and reduced levels of free chlorine. Canadian Journal of
Microbiology , 36 (2), 109-116.
Yu-sen, E. V. (2002). Negative effect of high pH on biocidal efficacy of copper and silver
ions in controlling Legionella pneumophila. Applied and environmental microbiology , 68 (6),
2711-2715.

LIST OF TABLES
Table 1. Pool water quality and operational parameter values
Table 2. Analytical methods for (physico)chemical parameters
Table 3. Analytical methods for bacteriological parameters
Table 4. Chemical parameters in the end phase as determined by two monitoring labs
Table 5. Bacteriological parameters in the end phase as determined by two monitoring labs
Table 6. Change in FAC after filtration
LIST OF FIGURES
Figure 1. General scheme
Figure 2. Gradual changing of the chemical parameters during startup
Figure 3. Evolution of bather load (in average number of visitors/day) during the project
Figure 4. Use of sodium hypochlorite (liter 12 %/day) in relation to the number of visitors
Figure 5. Correlation between the average FAC and the calculated “net chlorine consumption
per 1000 bathers”
Figure 6. Chronologic representation of the average FAC and the total chlorine consumption
per 1000 bathers
AKNOWLEDGEMENTS
The authors wish to express their gratitude for significant assistance and cooperation by the
people of Cofely Services, Sportoase and VAZG.

You might also like