Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Chlorine Dose

Point C in the chlorine dose-residual curve represents the breakpoint chlorination


dose at which all chloramines get decomposed to nitrogen trichloride, N2, or N2O,
and free chlorine residuals (Cl−, OCl−, HOCl) start increasing sharply.

From: Industrial Water Treatment Process Technology, 2017

Related terms:

Disinfection, Ammonia, Chlorination, Chloramine, Coliforms, Free Chlorine, Or-


ganic Carbon, Residual Chlorine

View all Topics

Disinfection of Water
Malcolm J. Brandt BSc, FICE, FCIWEM, MIWater, ... Don D. Ratnayaka BSc, DIC,
MSc, FIChemE, FCIWEM, in Twort's Water Supply (Seventh Edition), 2017

11.11 Typical Chlorine Dose


Typically chlorine doses to final treated waters are in the range 0.2–2.0 mg/l of
free chlorine to give a residual of about 0.02–0.3 mg/l at the consumer’s tap. The
lower doses (0.2–0.5 mg/l) tend to be used on clear groundwaters not subject to
pollution; the higher doses relate to treated surface waters or to well or borehole
supplies which are liable to sudden pollution, where superchlorination followed by
partial dechlorination after adequate contact time may be advised. Some raw waters,
particularly surface waters, can have a high chlorine demand of 6–8 mg/l (Smith,
1990).

Free chlorine has a taste threshold concentration of 0.6–1.0 mg/l. Chlorine dosing


therefore often causes taste and odours, principally by the reaction of chlorine with
some of the many trace compounds in the water (Section 10.32).

Equipment for storing and dosing chlorine is described in Section 12.13.

> Read full chapter


Physico-chemical techniques for the re-
moval of disinfection by-products pre-
cursors from water
Tanwi Priya, ... Majeti Narasimha Vara Prasad, in Disinfection By-products in Drink-
ing Water, 2020

2.1.1.5 Impact of chlorine dose on trihalomethans formation


The amount of chlorine used for disinfection is referred to as the chlorine dose.
Studies have also shown that higher disinfectant doses increase the DBP formation
potential in water (Watson, 1993). Longer reaction time generally leads to higher
consumption of residual disinfectant and results in formation of more DBPs (Chen
and Weisel, 1998). However, THMs formation was not found to increase signif-
icantly when the chlorine doses were increased further (greater than breakpoint
doses). This may be due to the fact that the chlorines beyond breakpoints had an
insignificant amount of organics to react. This lends partial support to the fact
that excess chlorine beyond the breakpoint does not necessarily contribute to a
significant increase in THMs formation. However, the slow reactions of NOM and
chlorine in the distribution system may exert partial chlorine demand. The chlorine
demand is generally determined to satisfy the actual needs of the water disinfection
in the plants and distribution systems. The overall amount of chlorine required for
oxidation is typically insignificant in comparison to the chlorine required by the
NOM (Rodriguez et al., 2003). As the residue-free chlorine content decreases, the
concentration of THMs and HAAs production increases. However, at last, if the
residue chlorine was insufficient to satisfy the further reaction of THMs formation,
the quantity of THMs produced would be changed little. Furthermore, because of
THMs decomposition, with the residue-free chlorine decrease, THMs concentration
decreases.

> Read full chapter

Chemical Treatment Technology


Parimal Pal, in Industrial Water Treatment Process Technology, 2017

2.8.7 Chlorine Residuals


When chlorine is added to water for disinfection, it reacts with ammonia if present
forming chloramine-based residuals such as monochloramine (NH2Cl), dichlo-
ramine (NHCl2), and nitrogen trichloride (NCl3). These residuals are collectively
called combined available chlorine. Chlorine also forms residuals like Cl−, OCl−, and
HOCl where the total amount of these residuals (Cl−, OCl−, and HOCl) is called free
available chlorine. The sum of the combined available chlorine and free available
chlorine is called the total residual chlorine (TRC).

Residual chlorine plays a significant role in ensuring protection of water from any
microbial contamination during pipeline transport of potable water from treatment
plants to user points. However, concentration and effectiveness of residual chlorine
depends on the chlorine dose applied, as described in the next section.

The chlorine dose-residual curve as shown in Fig. 2.7 indicates the consumption of
chlorine and formation of chlorine residuals through a series of reactions represent-
ed by Eqs. (2.44)–(2.54) in Sections 2.7.4 and 2.7.5.

Figure 2.7. Chlorine dose-residual curve.

In Fig. 2.7, zone OA represents the chlorine-destruction zone when almost all
applied chlorine is used for useful reactions (2.48–2.51) through which iron (Fe+2),
manganese (Mn+2), and hydrogen sulphide (H2S) are separated from water thereby
improving the quality. Thus in zone OA, no residual chlorine is observed. Zone
AB represents the zone of formation of combined chlorine residuals compris-
ing organo-chloro compounds like chloramines such as monochloramine, dichlo-
ramine, nitrogen trichloride, and disinfection byproducts like trihalomethanes, bro-
mochloromethane, etc., through reactions (2.50)–(2.52). Zone BC represents the
formation of free chlorine residuals through reactions (2.42)–(2.45). Zone CD repre-
sents formation and stockpiling of free and combined chlorine residuals as chlorine
is continually added to water. Point C in the chlorine dose-residual curve repre-
sents the breakpoint chlorination dose at which all chloramines get decomposed
to nitrogen trichloride, N2, or N2O, and free chlorine residuals (Cl−, OCl−, HOCl)
start increasing sharply. Beyond this breakpoint chlorination, chlorine residuals in
drinking water remain as free and combined residuals. The breakpoint chlorination
dose is the dose of chlorine that satisfies all chloride demand of water so that further
addition of chlorine will result only in an increase in free residual chlorine. The safe
dosage of chlorine is 1–2 mg/L but during an outbreak of any epidemic it has to be
increased by several fold (e.g., 10–15 mg/L) and is called superchlorination.

> Read full chapter

Occurrence of trihalomethanes in
drinking water of Indian states: a criti-
cal review
Binota Thokchom, ... Snigdha Dutta, in Disinfection By-products in Drinking Water,
2020

4.3.2.1.4 Kanpur city


Mishra et al. (2012) studied the TFP of the groundwater of Jajmau in Kanpur which
is the hub of tanneries. Sample collection was done in January and May 2009.
Groundwater samples were chlorinated with free chlorine residuals and combined
chlorine residual doses. For the ground water samples collected in January 2009, the
total THMs (TTHMs) levels (i.e., sum of the ratio of the concentrations of each of the
THM to its corresponding GV) were below the WHO GV ≤1 except the sample with
17 mg/L chlorine dose which was 320.4 μg/L. Likewise in the samples collected in
May 2009, the TTHMs levels were found to be below WHO GV ≤1 except in the value
at 17 mg/L of chlorine dose. In both January 2009 (winter) and May 2009 (summer),
CHCl3 was found to be the only THM species identified in all the samples and the
samples collected in summer (May) had higher TFP values. This work also studied
the relationship between reaction time and TFP. It showed that the THM formation
rate in the first 24 hours is the fastest and higher reaction time led to higher CHCl3
concentration in higher dose of chlorine. Jajmau being the hub of tanneries had
groundwater with high TOC both in summer and winter due to percolation of
liquid-phase toxic pollutants drained out from industries.

Another interesting study (Mishra and Dixit, 2013) was carried to examine the effects
of anthropogenic sources and other precursors on THM formation. Sampling of
treated water just before disinfection was done from Ganga Barrage WTP, Kanpur,
during January to June 2009. The TTHMs levels of April, May, and June were above
the WHO GV ≤1. Observing the seasonal variation of TFP, it was seen that the TFP
values in May and June were three times higher than those in January and there is
abrupt rise in the month of February. This is attributed to rapid decay of vegetation
during spring. CHCl3 levels were higher than all the other three forms of THMs.
Studying the relationship between reaction time and varying chlorine dosages with
THM formation in water, it was seen that the formation rate was very fast in the first
24 hours. CHCl3 formation was influenced by the higher reaction time causing rise
in formation of CHCl3. However, CHBrCl2, CHBr3, and CHClBr2 were independent
of the reaction time exceeding 24 hours. With the exception in February the value
of TFP increased as the TOC level increased in samples with higher chlorine doses.
They reported that reduction in TOC causes higher Br−/TOC ratio and consequently
there is higher Br−-THM concentration, and this explains abrupt elevation of TFP in
February.

> Read full chapter

Chemical Storage, Dosing and Control


Malcolm J. Brandt BSc, FICE, FCIWEM, MIWater, ... Don D. Ratnayaka BSc, DIC,
MSc, FIChemE, FCIWEM, in Twort's Water Supply (Seventh Edition), 2017

12.29 Cascade Control


Cascade control is a method of control combining two feedback loops, with the
output of one controller (the primary controller) adjusting the set-point of a second
controller (the secondary controller). It is particularly useful on processes with a
long process time, such as chlorine dose control across a chlorine contact tank
(Section 11.5), where loop times in excess of 30 minutes can be expected. The
primary controller receives a signal from the water quality monitor at the outlet
from the contact tank and compares this with the set-point entered by the operator.
The output from this controller varies the set-point of the secondary controller. The
secondary controller then adjusts the chemical dose rate to maintain this internal
set-point, as measured by a water quality monitor at the inlet to the chlorine contact
tank. The secondary controller thus acts on the dosing pump or regulating valve to
bring both the residual at the inlet to the chlorine contact tank and the residual at
the outlet from the chlorine contact tank back to their desired values. In this case,
it is important that the set-point for the primary controller is changed at an interval
no less than the overall process loop time.

> Read full chapter


Drinking Water Quality and Treatment
Dan Askenaizer, in Encyclopedia of Physical Science and Technology (Third Edition),
2003

VII.A Chloramines
Chloramines (referred to as combined chlorine) are formed when water containing
ammonia is chlorinated. There are three inorganic chloramine species: monochlo-
ramine (NH2Cl), dichloramine (NHCl2), and trichloramine (NCl3). The species of
chloramines that are formed depends on factors such as the ratio of chlorine to
ammonia–nitrogen, chlorine dose, temperature, pH, and alkalinity.

The principal reactions for chloramine formation are presented below:

At a typical water treatment plant, the dominant chloramine species will be mono-
chloramines. Chloramine generating reactions are 99% complete within a few
minutes. Chloramines are a weak disinfectant that are less effective against viruses
or protozoa than free chlorine but produce fewer disinfection by-products. The use
of chloramines as a DBP control strategy is well established in the United States.
Chloramines are generated onsite at the treatment plant. Anhydrous ammonia and
ammonia sulfate are examples of ammonia containing chemicals used by water
systems to form chloramines. In most situations in the United States, chloramines
are used as a secondary disinfectant to maintain a residual in the distribution system.

> Read full chapter

Transport and Fate of Chlorinated


By-Products Associated with Cooling
Water Discharges
Adeyinka E. Adenekan, ... Joseph P. Smith, in Proceedings of the 1st Annual Gas
Processing Symposium, 2009

2 Chemistry of Seawater Chlorination


When added to seawater, chlorine oxidizes bromide ions present (approximately
80 mg/L in Arabian Gulf water compared to an average of 67 mg/L in the world's
oceans) yielding hypobromous acid (HOBr) and hypobromite ion (OBr ). In addition,
depending on the chlorine dose relative to the concentration of bromide ions,
some portion of the initial products of chlorine hydrolysis - hypochlorous acid
(HOCl) and hypochlorite ion (OCl ) - could remain in solution. These four substances
togethermake up what is referred to as "Free Residual Oxidants (FRO)". Figure 2
illustrates the situation.

Figure 2. The chemistry consequent upon addition of sodium hypochlorite to natural


seawater(After Khalanski, 2003).

The free residual oxidant components react with dissolved natural organic matter
present in the seawater and it is these reactions that form the chlorination by-prod-
ucts that are the subject of this paper. These by-products include trihalomethanes
(e.g., bromoform), haloacetic acids (e.g., dibromoacetic acid), halophenols (e.g.,
2,4,6-tribromophenol), and haloacetonitriles (e.g., dibromoacetonitrile).

When ammonia is present, HOBr and HOCl participate in reactions that lead to
the formation of bromamines and chloramines. The bromamines and chloramines
together make up what is referred to as "Combined Oxidants". The term "Total
Residual Oxidants (TRO)" is used to represent the sum of Free Residual Oxidants
and Combined Oxidants.

> Read full chapter

Disinfection Systems
Chittaranjan Ray Ph.D., P.E., Ravi Jain Ph.D., P.E., in Low Cost Emergency Water
Purification Technologies, 2014

4.4.1 Liquid Chlorine as a Disinfectant


Household bleach can be used to disinfect drinking water. A specific number of
drops should be added to the water, depending on the concentration of the chlorine
in the bleach solution, the quantity of the water being treated, and the quality of
the water being treated. Water quality factors that increase the necessary chlorine
dose include increased turbidity, increased pH, and decreased temperature (Burch
and Thomas, 1998). As turbidity impedes disinfection, water should be filtered using
a cloth filter or a coffee filter before chlorination. Inactivation of microorganisms
such as Vibrio, Salmonella, and Shigella generally requires a contact time of 30 min,
depending on the dose used. Chlorine has been found to be fairly ineffective against
protozoans because they can form spores, so a much longer contact time of 4 h
is required. Even with the longer contact time, chlorine might still be ineffective
against some microbes, such as Cryptosporidium (Table 4.2).

Table 4.2. EPA (2006) Recommendations for Disinfecting Contaminated Water Using
Household Bleach

Available Chlorine Drops Per Quart-Gallon of Drops Per Liter of Clear Water
Clear Water
1% 10 per quart-40 per gallon 10 per L
4-6% 2 per quart-8 per gallon 2 per L
7-10% 1 per quart-4 per gallon 1 per L

Double the amount of drops if water is murky or turbid.

An advantage to using chlorine as a disinfectant is that it leaves chlorine residual


in the water. This residual helps prevent recontamination of the water if it’s stored
correctly. Chlorination using sodium hypochlorite (NaOCl) was found to improve
stored water quality after the Indonesian tsunami (Gupta et al., 2007; Loo et al.,
2012). NaOCl can also be used for cleaning storage vessels, but, in some cases, it
did not prevent recontamination of storage vessels (Steele et al., 2008).

When water is treated as a batch solution with chlorine drops, little maintenance of
the treatment device is required. Filtering the water and adding the chlorine drops
does not require skilled operation. Chlorine degrades quickly, so the treated water
needs to be stored in a closed container in a cool, dark place so the degradation
process is slowed. Even then, the half-life of chlorine is 2 months. Because of
this quick degradation, chlorine supplies need to be replenished often to ensure
sufficient disinfection. The costs associated with chlorine use are mainly derived
from the cost of replenishing the chlorine. Labor takes around 20 min to produce a
batch of disinfected water and thus is not highly time sensitive (Burch and Thomas,
1998). However, if turbid water must be filtered, costs and labor will increase.

> Read full chapter

Analysis and Formation of Disinfection


Byproducts in Drinking Water
Jean-Luc Boudenne, ... Bruno Coulomb, in Comprehensive Analytical Chemistry,
2021

1 Introduction
Inorganic chloramines comprise three chemical compounds formed by reactions
between chlorine and nitrogen-containing substances in aqueous solution: mono-
chloramine (NH2Cl, MCA), dichloramine (NHCl2, DCA) and trichloramine (NCl3,
TCA). The formation of inorganic chloramines is illustrated in Fig. 1 exemplified
with ammonium (NH4+) as a nitrogen-containing substance and hypochlorous acid
(HClO) as a chlorinated disinfectant [1]. MCA is first formed at a chlorine dose ratio
below 1:1 (Cl2:NH4+, mol:mol), DCA is then formed before being eliminated by reac-
tion with MCA at a chlorine dose ratio between 1 and 1.5 (Cl2:NH4+, mol:mol). TCA
is finally formed at higher dose and is relatively stable at neutral pH. Chloramines
generation is thus highly dependent on the ratio chlorine:nitrogen, but also on pH,
of and, in a lesser extent, on temperature and contact time [2]. Fig. 2 illustrates the
pH- and Cl:N ratio-dependence of the speciation of chloramines [3].
Fig. 1. Reaction mechanism between chlorine and ammonium leading to the for-
mation of chloramines.From C.T. Jafvert, R.L. Valentine, Reaction scheme for the
chlorination of ammoniacal water, Environ Sci. Technol. 26 (1992) 577.
Fig. 2. Speciation of chloramines (A vs pH; B. vs molar ratio N:Cl (pH = 7.5)).From
C.T. Jafvert, R.L. Valentine, Reaction scheme for the chlorination of ammoniacal
water, Environ. Sci. Technol. 26 (1992) 577.

Once formed in waters, an equilibrium exists between MCA, DCA, and TCA which is
dependent on the pH solution and the ammonium ion concentration. The principal
specie present in solutions of pH less than 3 is NCl3; of pH in the range 3–5 is NHCl2
and of pH above 8 is NH2Cl. These equilibria are represented by Eqs. (1) and (2):

(1)

H+ + 2 NH2Cl   NHCl2 + NH4+

(2)

H+ + 3 NHCl2   2 NCl3 + NH4+

Approximate equilibrium constants for Eqs. (1) and (2) were estimated as 106 and 104
respectively [4]. In strongly alkaline solutions, chloramine is converted to hypochlo-
rite (Eq. 3):

(3)

NH2Cl + OH−   NH3 + ClO−

Other nitrogenous compounds have been proved to be precursors of chloramines


in presence of free chlorine:urea [2], creatinine [5,6], formamide, lysine, asparagine,
arginine, acetamide, glutamic acid, serine and -alanine [6,7].

Inorganic chloramines are not commercially-available but could be produced in


situ when needed. Therefore, monochloramine-used as a secondary disinfectant
treatment of drinking waters in several countries (mainly in Japan, USA, Australia,
and in some Northern European countries) [8,9] is generated by the reaction of
chlorine (hypochlorite, hypochlorous acid, or gaseous chlorine) with anhydrous
ammonia (NH3, gaseous form), ammonia (NH4OH, liquid form) and ammonium
sulphate ((NH4)2SO4, solid form) [10]. Despite its known ability to form iodo and
bromo-disinfection by-products (DBPs) and nitrogen-containing DBPs (such as
nitrosamines or haloacetamides) [11–14], MCA is also used to disinfect wastewaters
before discharge in natural environments or for reuse purposes, ballast waters,
industrial waters in need of clean waters, or of cooling waters to prevent biofouling
[15–18]. NHCl2 is considered to be a stronger disinfectant than NH2Cl [19] but its
lifetime in aqueous solutions prevents its use in distribution systems (disappearance
either by autodecomposition or by autocatalytic reduction) [20], whereas NCl3 has no
disinfection properties and is also volatile: their presences in waters result therefore
from unwanted reactions, and they are thus considered as DBPs. NCl3 is mainly of
health concern in the air of indoor swimming pools where it has been associated
with an increased risk of respiratory effects in highly exposed populations [5,21–23].

Inorganic chloramines can act as tracers in waters, by indicating insufficient


water treatment (as elimination of nitrogenous compounds) and by preventing
formation of more potent DPBs, such as nitrosamines and dichloromethylamine
[22,24]. Indeed, despite chloramines are known to survive in water for several days
after treatment (in absence of other reactive species), their hydrolysis may produce
hydroxylamines and, by subsequent oxidation, nitrous and nitric acid. Chlorine
added or chlorine transfer from inorganic chloramines to organic amines may then
lead to the production of organic chloramines [25]. Inorganic chloramines, that
are powerful oxidants, are also able to react with iodide (I−) and bromide (Br−).
MCA can react with Br− to form bromochloramine (NHBrCl), as well as mono- and
di-bromamine and hypobromous acid [26,27], while MCA and TCA may oxidize I−
into hypoiodous acid (HOI) in typical drinking water treatment conditions [28,29],
DCA can also oxidize I− but at a much lesser rate [30]. Once formed, these halogenous
ions with natural or organic matter to form brominated an iodinated by-products
whose toxicities are known to be higher than their chlorinated analogues [31].

Analysis of inorganic chloramines in waters is therefore of particular interest in the


field of DBPs as they are at the origin of many others DBPS and as they induce
higher global toxicity of waters. Their analysis is also challenging because of their
high reactivities, unstabilities in waters, and several methods are available but often
presenting low limits of quantification and detection, or suffering from lack of
sensitivity (interferences from organic chloramines or matrix effects) [32]. However,
relative recent use of mass-spectrometry could allow to overcome these analytical
bias, according to kind of waters analysed.

> Read full chapter

Chlorination disinfection by-products


in municipal drinking water – A review
Mohd Aamir Mazhar, ... Viola Vambol, in Journal of Cleaner Production, 2020

3.4.3 Optimizing chlorine dosing through disinfection bench-


marking
The amount of disinfection given by the water treatment plant can be benchmarked
in terms of CT values to reduce the microbial growth and to optimize reduction of
DBPs formation. By using the history of operating conditions like flow rate, chlorine
doses and residuals, water levels in tanks etc. and water quality data like temperature,
pH, turbidity, etc., the reduction in chlorine doses can be determined. The improved
values of CT by treatment units can also be provided by tracer tests or computational
fluid dynamics (CFD) modelling especially by which we can identify the non-ideal
mixing zones which will influence the contact time as well as formation of DBPs.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like