Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

This article was downloaded by: [University of California, San Diego]

On: 13 June 2015, At: 05:46


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Journal of Environmental Science and Health, Part


A: Toxic/Hazardous Substances and Environmental
Engineering
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/lesa20

Corrosion control in water supply systems: Effect of


pH, alkalinity, and orthophosphate on lead and copper
leaching from brass plumbing
a b
Y. S. Tam & P. Elefsiniotis
a
Watercare Services Ltd. , Auckland, New Zealand
b
Department of Civil and Environmental Engineering , University of Auckland , Auckland,
New Zealand
Published online: 04 Sep 2009.

To cite this article: Y. S. Tam & P. Elefsiniotis (2009) Corrosion control in water supply systems: Effect of pH, alkalinity, and
orthophosphate on lead and copper leaching from brass plumbing, Journal of Environmental Science and Health, Part A:
Toxic/Hazardous Substances and Environmental Engineering, 44:12, 1251-1260, DOI: 10.1080/10934520903140009

To link to this article: http://dx.doi.org/10.1080/10934520903140009

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Journal of Environmental Science and Health Part A (2009) 44, 1251–1260
Copyright C Taylor & Francis Group, LLC

ISSN: 1093-4529 (Print); 1532-4117 (Online)


DOI: 10.1080/10934520903140009

Corrosion control in water supply systems: Effect of pH,


alkalinity, and orthophosphate on lead and copper leaching
from brass plumbing

Y.S. TAM1 and P. ELEFSINIOTIS2


1
Watercare Services Ltd., Auckland, New Zealand
2
Department of Civil and Environmental Engineering, University of Auckland, Auckland, New Zealand
Downloaded by [University of California, San Diego] at 05:46 13 June 2015

This study explored the potential of lead and copper leaching from brass plumbing in the Auckland region of New Zealand. A five-
month field investigation, at six representative locations, indicated that Auckland’s water can be characterized as soft and potentially
corrosive, having low alkalinity and hardness levels and a moderately alkaline pH. More than 90% of the unflushed samples contained
lead above the maximum acceptable value (MAV) of 10 µg/L (New Zealand Standards). In contrast, the copper level of unflushed
samples remained consistently below the corresponding MAV of 2 mg/L. Flushing however reduced sharply metal concentrations,
with lead values well below the MAV limit. Generally, metal leaching patterns showed a limited degree of correlation with the
variations in temperature, dissolved oxygen and free chlorine residual at all sampling locations. Furthermore, a series of bench-scale
experiments was conducted to evaluate the effectiveness of pH and alkalinity adjustment, as well as orthophosphate addition as
corrosion control tools regarding lead and copper dissolution. Results demonstrated that lead and copper leaching was predominant
during the first 24 hr of stagnation, but reached an equilibrium state afterwards. Since the soluble fraction of both metals was
small (12% for lead, 29% for copper), it is apparent that the non-soluble compounds play a predominant role in the dissolution
process. The degree of leaching however was largely affected by the variations in pH and alkalinity. At pH around neutrality, an
increase in alkalinity promoted metal dissolution, while at pH 9.0 the effect of alkalinity on leaching was marginal. Lastly, addition
of orthophosphate as a corrosion inhibitor was more effective at pH 7.5 or higher, resulting in approximately 70% reduction in both
lead and copper concentrations.
Keywords: Water supply, corrosion control, lead, copper, metal leaching, orthophosphate, alkalinity.

Introduction by leaching from fittings.[3,4] In particular, lead has been


recognized as a cumulative poison to the human body,
Corrosion (or plumbosolvency) is one of the major chal- which can cause irreversible damage to the nervous system
lenges water utilities face worldwide. From an asset man- and create neurological problems especially to infants and
agement perspective, the principal concerns for a water children.[5,6]
utility are twofold: (i) corrosion of the distribution sys- To minimize such potential hazards to public health
tem pipes which may eventually lead to water leakage and infrastructure, the metal content in water supplies has
and loss of hydraulic capacity due to the accumulation been regulated by various authorities worldwide. For in-
of corrosion products,[1] and (ii) premature failure of water stance, strict limits on lead concentrations in drinking water
pipes which impacts the structural integrity of the distri- have been promulgated by the United States Environmen-
bution systems.[2] From a health viewpoint, drinking wa- tal Protection Agency (U.S. EPA) due to lead’s toxicity.[7]
ter suppliers are concerned about corrosion of plumbing In addition, the Guidelines for Canadian Drinking Water
fittings, which may result in the staining of fixtures and Quality have a maximum acceptable concentration (MAC)
expose consumers to heavy metal contamination caused of 10 µg/L for lead.[8] The European Commission (EC)
also plans to tighten the drinking water lead standard
Address correspondence to P. Elefsiniotis, Department of to 10 µg/L by 2013, in line with the World Health Or-
Civil and Environmental Engineering, University of Auckland, ganization (WHO) guideline value.[9] Presently, the max-
Private Bag 92019, Auckland 1030, New Zealand. E-mail: imum acceptable value (MAV) of lead in New Zealand
t.elefsiniotis@auckland.ac.nz is 10 µg/L,[10] which is consistent with several interna-
Received December 9, 2009. tional drinking water standards.[8,9] Regarding cooper, the
1252 Tam and Elefsiniotis

corresponding MAV is 2 mg/L for health reasons with a (asbestos-cement and concrete-lined steel) in Auckland’s
guideline value (GV) of 1 mg/L for aesthetic purposes.[10] drinking water distribution system. The experimental pro-
Corrosion is a complex phenomenon as its behavior is cedure followed was established by the New Zealand Min-
dependent on water chemistry as well as a number of envi- istry of Health.[10] Specifically, each brand new yellow brass
ronmental factors including pH, alkalinity, chlorine resid- sample tap (containing 58.0% copper, 37.9% zinc and 3.78%
ual, temperature, dissolved oxygen, chloride, phosphate, lead) was completely filled with water, kept stagnated for
and organic matter.[11,12] Although certain factors, such as 4 hr and then flushed to waste prior to installation to the
pH and alkalinity, may be controlled during water treat- water mains. This procedure was repeated for a week to
ment to reduce the impact of corrosion,[13] other factors minimize variability.
such as temperature are difficult to control. To estimate the To evaluate the effect of overnight storage on metal re-
corrosion potential of a certain water type, several indices lease in brass fittings, all sample taps were flushed for 2
have been introduced, including the Langelier Saturation min, 24 hr prior to sampling to achieve a uniform resi-
Index (LSI) and the Aggressiveness Index (AI). The LSI is dence time. On each sampling occasion, both unflushed
commonly used to predict whether calcium carbonate will samples (i.e., first-draw water) and flushed samples (i.e.,
precipitate or dissolve in a particular case, thereby reducing running water collected after 2 min) were taken. In total,
Downloaded by [University of California, San Diego] at 05:46 13 June 2015

the water’s tendency for corrosion.[14] 84 samples (14 from each of the 6 sampling locations) were
At an LSI of zero, the water is in equilibrium with cal- collected on a weekly basis during a 5-month period. The
cium carbonate. A negative LSI suggests that the water sampling locations were selected based on the ease of acces-
dissolves calcium carbonate, and a positive LSI value indi- sibility, provision of a representative coverage of different
cates that the water will tend to deposit calcium carbonate. water types, and pipe materials used for conveyance. For
Usually a LSI between – 0.5 and 0.0 is desirable for infras- brevity purposes, the six locations have been designated as
tructure as well as public health protection.[15] On the other S1 through S6 (Table 1).
hand, the AI can predict the degree of deterioration in wa-
ter quality that arises from pipe dissolution by measuring
pH, alkalinity and calcium hardness.[16] If the AI ≤ 10.0, Bench-scale experiments
the water is categorized as highly aggressive; if the AI lies The bench-scale experiments examined the performance of
between 10.0 and 11.9, the water is considered moderately corrosion control techniques under stagnant conditions to
aggressive; and if the AI ≥ 12.0, the water is classified as simulate the dead-end sections (i.e. worst-case scenario) of
completely non-aggressive.[17] a distribution system. Pipe sections made up of concrete-
Currently, the most common methods for corrosion con- lined steel were connected with identical brass fittings uti-
trol involve either adjustment of pH and alkalinity or ad- lized in the field study as illustrated in Fig. 1. The pipes had
dition of corrosion inhibitors such as orthophosphate and been in service for approximately 3 years and no apparent
silicate. The pH and alkalinity adjustment affects the ther- tubercles were observed. Each pipe section had an inner di-
modynamic properties of the carbonate system to achieve ameter of 0.20 m, a length of 0.50 m and a carrying capacity
reduction in the corrosive nature of the water.[18,19] Alterna- of approximately 16 L. All pipe sections were covered with
tively, addition of corrosion inhibitors aids the formation of a polyurethane transparent lid to ensure minimal airspace
protective coating on a pipe, thereby preventing corrosive between the surface of the water sample and the covering
water from attacking the metal surface.[20−22] to prevent contact with the atmosphere.
Both field and bench-scale investigations can be used to Initially all pipes were rinsed three times with 0.1N
determine water quality and quantify the corrosion poten- sodium hydroxide (NaOH) to remove organic deposits and
tial, thus enabling water utilities to gain an insight into
corrosion problems and implement the necessary correc-
tive strategies. The objectives of this research are therefore Table 1. Sampling locations, pipe material and corresponding
as follows: (i) conduct a field study to assess the degree of water source.
corrosion in the drinking water of the Auckland region in
Sampling location Pipe material Water source
New Zealand and (ii) evaluate the effect of pH, alkalin-
ity and inhibitor (orthophosphate) addition on corrosion Ardmore & Concrete-lined steel (CLS) Dam & river
control in bench-scale experiments. Waikato (S1)
Waikato (S2) Concrete-lined steel (CLS) River
Ardmore (S3) Concrete-lined steel (CLS) Dam
Ardmore (S4) Asbestos-cement (AC) Dam
Materials and methods Waitakere (S5) Concrete-lined steel (CLS) Dam
Ardmore, Concrete-lined steel (CLS) Dam
Field investigation Waikato, Huia
& Waitakere
The primary focus of the field work was to investigate the
mix (S6)
leaching potential of lead and copper in water conduits
Corrosion control in water supply systems 1253

During the corrosion inhibition tests, disodium hydro-


gen orthophosphate (Na2 HPO4 ), containing 21.8% P, was
used as a corrosion inhibitor. To evaluate the effect of or-
thophosphate addition, a passivation dose of 1.6 mg/L P
was applied for 10 consecutive days to ensure that a pro-
tective layer was formed on the appropriate pipes prior to
orthophosphate treatment.[22,23] Following this, an amount
of 0.8 ± 0.05 mg/L P was added to all samples. The exper-
imental design was identical to the one followed previously
(Table 2), except that there was no need to conduct experi-
ments at pH 9.0 with orthophosphate addition (explained
later).

Analytical methods
Downloaded by [University of California, San Diego] at 05:46 13 June 2015

During both the field and the bench-scale parts of the study,
Fig. 1. Experimental set-up for bench-scale tests.
pH, temperature, dissolved oxygen (DO), and free chlorine
residual were measured on-site during sampling. All other
then rinsed five times with distilled water. Since it was ev-
parameters such as lead and copper (both total and solu-
ident that the naturally low alkalinity levels (i.e. 20 to 50
ble), total alkalinity, total hardness, total dissolved solids
mg/L as CaCO3 ) in Auckland’s water had induced concrete
(TDS), chloride, and sulfate were determined in the labo-
leaching in the distribution system,[23] a higher range of al-
ratory. Each water sample was stirred using a cleaned glass
kalinity was applied in this study. To that effect, sodium
rod prior to obtaining the aliquot for analysis to ensure
bicarbonate (NaHCO3 ) was added to achieve concentra-
uniform distribution.
tions of 100, 125, and 150 mg/L as CaCO3 , respectively,
With respect to the on-site measurements, pH and
while the control samples had an alkalinity of 20 mg/L as
temperature were recorded after 20 sec while stirring
CaCO3 . At each alkalinity level tested, the pH was subse-
the solution using a portable kit (pHTestr30, Eutech
quently adjusted to 7.0, 7.5, 8.0, and 9.0 by bubbling carbon
dioxide gas into the water and was measured at 1-hr, 4-hr, InstrumentsR ). The DO was measured with a portable dis-
24-hr, 48-hr and 72-hr intervals. To minimize any potential solved oxygen meter (Microprocessor Oximeter Oxi 325R )
pH changes, the pH was continuously monitored and ad- using an oxygen sensor (CellOx325). Free chlorine residual
justed during the initial 4 hr, and then corrected to within was measured at a wavelength of 530 nm using a HACH
±0.1 unit of the initial values at the time of sampling there- colorimeter by the colorimetric version of the N,N-diethyl-
after. The range of pH and alkalinity tested is shown in p-phenylenediamine (DPD) method.[24]
Table 2. All tests were conducted in duplicate to enhance For total metal (lead and copper) determination, sam-
the statistical reliability of the measurements. ples were preserved with concentrated nitric acid at pH <
2 and then digested in accordance with the nitric acid di-
Table 2. Range of pH and alkalinity tested during bench-scale gestion #3030E method prior to analysis.[24] Samples for
experiments. soluble metal analysis were collected into cleaned 100-ml
polyethylene bottles and then filtered through a 0.45 µm fil-
pH Alkalinity (mg/L as CaCO3 )
ter. Lead and copper were analyzed by inductively coupled
7.0 100 plasma mass spectrometry (ICP-MS). Alkalinity and hard-
125 ness were determined by digital titration methods (Hach
150 #8203 and #8213, respectively) and TDS by a gravimetric
Control method (Hach #8277).[24] Chloride and sulfate determina-
7.5 100 tion was carried out by a computer-controlled Dionex DX-
125 120 ion chromatograph equipped with an AS9-HC column.
150 Further details on all sampling and analytical procedures
Control
are available elsewhere.[23]
8.0 100
125
150
Control Statistical methods
9.0 100 Regression analysis was performed in Microsoft Excel
125 Analysis ToolPakR . The statistical significance of a wa-
150
ter quality parameter was evaluated based on the p-value
Control
obtained from regression analysis. In general, a p-value
1254 Tam and Elefsiniotis
Table 3. Average water quality parameters at each sampling quality indices (LSI and AI). In general, Auckland’s wa-
location. ter is soft in nature, with low alkalinity (19 to 55 mg/L as
CaCO3 ), low hardness (25 to 60 mg/L as CaCO3 ), and a
Parameters S1 S2 S3 S4 S5 S6
moderately alkaline pH (7.80 to 8.33). The majority of the
Temperature, ◦ C 21.4 22.5 21.4 21.3 20.7 20.5 water samples from the six sampling locations were com-
pH 8.04 7.99 7.97 7.95 8.33 7.80 parable, with the exception of location S2, where higher
DO, mg/L 7.71 6.89 7.51 7.64 7.90 7.80 alkalinity, hardness, sulfate and TDS concentrations were
Residual 0.64 0.56 0.79 0.81 0.63 0.54 consistently recorded. The higher alkalinity, hardness and
chlorine, mg/L TDS values observed in location S2 can be attributed, at
Total alkalinity, 25.4 54.8 19.2 19.7 25.4 21.6 least partially, to lime addition at the Waikato water treat-
mg/L as
ment plant. Furthermore, the LSI values indicated that
CaCO3
Total hardness, 31.2 60.2 24.8 24.6 42.4 34.7
all samples were corrosive in nature. This observation is in
mg/L as agreement with the interpretation of the AI, which revealed
CaCO3 that the water was moderately aggressive.
Figure 2 displays the distribution of total (i.e. soluble
Downloaded by [University of California, San Diego] at 05:46 13 June 2015

Total dissolved 91.2 154 81.4 84.1 107 103


solids, mg/L plus non-soluble) lead release following a 24-hr standing
Chloride, mg/L 16.2 19.3 15.3 15.3 23.8 23.0 time in unflushed samples. It was found that more than
Cl 90% of the values exceeded the MAV of 10 µg/L and were
Sulfate, mg/L 11.5 20.1 8.27 8.22 12.4 13.1 classified as “aggressive” according to the New Zealand
SO4 standards.[10] Samples from locations S1 and S2 exhibited
Langelier −1.10 −0.52 −1.34 −1.35 −0.67 −1.41 similar water characteristics and therefore had higher me-
Saturation
dian lead concentrations. The other samples (locations S3
Index (LSI)
Aggressiveness 10.8 11.4 10.5 10.5 11.2 10.5
to S6) had lower lead values and the relative concentrations
Index (AI) were reasonably consistent.
It is worth noting that the S2 samples, despite having a
relatively high alkalinity at 55 mg/L as CaCO3 , also ex-
between 0.1 and 0.03 is considered to indicate a moder- hibited a median lead concentration of 28.0 µg/L, which is
ate correlation, while a p value below 0.03 reveals a strong the second highest mean value (Fig. 2). It has been reported
correlation.[25] that alkalinity up to 75 mg/L as CaCO3 may reduce cor-
rosion in surface waters.[18,26,27] Furthermore, the LSI of
the S2 water was −0.52 which suggested the least corrosion
Results and discussion
potential; nevertheless, the median lead concentration was
relatively high. It should be noted that there has been some
Field investigation
concern raised about the reliability of LSI to predict the
Results summarized in Table 3 include the mean values of aggressiveness of water.[17,28] Results from this study tend
the various water quality parameters that were monitored to support the notion that LSI may not be a dependable
during the field investigation, plus the corresponding water indicator of water quality.

80

70

60
Lead Concentration (µg/L)

50

40

30

20

10

0
S1 S2 S3 S4 S5 S6
Location

Fig. 2. Median total lead concentrations (including the 25th and 75th percentiles) in unflushed samples from the field locations.
Corrosion control in water supply systems 1255
40

35

Mean Lead Concentration (µg/L)


30

25

20

15

10

0
S1 S2 S3 S4 S5 S6
Location
Downloaded by [University of California, San Diego] at 05:46 13 June 2015

Unflushed sample Flushed sample

Fig. 3. Mean total lead concentrations in unflushed and flushed water samples from the field locations.

It has been reported that lead leaching can be also af- no clear correlation between pH and lead dissolution was
fected by the presence of sulfate and chloride ions, as well as observed in the field study. The mean pH values were rea-
TDS, since water with a high TDS content becomes a better sonably consistent in the range of 7.80 ± 0.48 to 8.33 ±
conductor.[29,30] Results revealed much higher average sul- 0.38. Regression analysis indicated that pH did not affect
fate and TDS concentrations in location S2, compared to all lead leaching from brass taps (P-value between 0.105 and
other samples (Table 3). The chloride concentration, how- 0.924). With respect to temperature, only location S6 re-
ever, appeared to be within the same range in all locations. vealed a strong correlation with the amount of lead leached
To examine the effect of flushing on metal leaching, from brass fittings (P-value = 0.019). In location S3, both
flushed samples also were taken (followed the procedure temperature and DO appeared to have an influence on lead
described previously) and compared to the unflushed ones. release to a varying degree (P-value = 0.081 and 0.040, re-
The mean lead concentrations detected in both types of spectively). Furthermore, the S1 and S2 samples suggested
samples are shown in Fig. 3. It is apparent that flushing a moderate correlation between lead dissolution and the
greatly reduced lead dissolution by 90 to 99%, resulting amount of free chlorine residual (P-value = 0.074 and
in concentrations below the MAV. The significant drop in 0.078, respectively).
lead levels confirms that lead dissolution occurred in the In addition to lead, copper was analyzed and the re-
brass fittings. sults have been summarized in Fig. 4. Location S1 had
An attempt was made to correlate the effect of the pa- the highest median total copper concentration (22.5 µg/L)
rameters measured to lead dissolution data. Regarding pH, and S4 had the lowest one (11.5 µg/L). In all locations,

50

45

40
Copper Concentration (µg/L)

35

30

25

20

15

10

0
S1 S2 S3 S4 S5 S6

Location

Fig. 4. Median total copper concentrations (including the 25th and 75th percentiles) in unflushed samples from the field locations.
1256 Tam and Elefsiniotis

the maximum copper levels fluctuated between 31 µg/L CaCO3 , provide a representative example of the pattern
and 46 µg/L. It is therefore apparent that the unflushed of total lead release as a function of the initial pH value.
water samples collected had copper levels well below the Overall, lead levels increased rapidly during the first 4 hr
corresponding MAV limit of 2 mg/L, according to the of stagnation; however, they appear to fluctuate around
New Zealand Standards.[10] As observed in the case of an “equilibrium” state after 24 hr, at all pH levels tested,
lead, flushing reduced the copper concentration in water except at the highest pH of 9.0. This denotes that a pH
by more than 90%. This provides additional evidence that higher than 8.0 may limit the leaching potential of lead.
copper leaching originated from the brass plumbing. Similar trends were observed in the tests with higher ini-
Although investigating the role of natural organic mat- tial alkalinity content as well (i.e., of 125 and 150 mg/L as
ter (NOM) on metal leaching was not part of this research CaCO3 ), suggesting that a moderate increase in alkalinity
project, as background information, the mean total organic has practically no effect on lead leaching. The pattern of
carbon (TOC) content of the samples tested within the lead release observed in this study is typical of that reported
Auckland region was 1.26 ± 0.56 mg/L. It should be noted in the literature.[13,33,34]
that the presence of NOM in surface waters, even at rela- The effect of stagnation time on copper leaching was
tively low concentrations (i.e., less than 2 mg/L as TOC) has found to be similar to that described for lead (Fig. 5). That
Downloaded by [University of California, San Diego] at 05:46 13 June 2015

been indicated to enhance lead and copper solubility.[31,32] is, a substantial increased in copper release was observed
In contrast, organic matter has been also found to pro- within the first 24 hr, which was followed by an “equilib-
mote coating in water pipes, thus reducing the corrosion rium” state. The leaching of copper was more pronounced
potential.[8] It is therefore apparent that the significance of at the lower pH values investigated (i.e. 7.0 and 7.5), be-
NOM on corrosion has not been conclusively determined. coming practically negligible at pH 9.
During the bench-scale experiments, DO, free chlorine
residual and temperature were also monitored. The initial
DO values varied between 6.7 to 7.2 mg/L, while the fi-
Bench-scale experiments
nal DO at the end of the 72 hr period was in the range
of 5.9 to 6.2 mg/L, while the temperature fluctuation was
Effect of stagnation time on lead and copper release
minimal (17 ± 1◦ C) throughout the tests. The decrease
The effect of stagnation time on lead and copper leached in DO concentration over the stagnation period followed
from the brass taps at varying pH and alkalinity values was first-order kinetics (Fig. 6), with an average rate constant
investigated by comparing metal concentrations at differ- k DO of 0.0036 ± 0.0004 hr −1 . Moreover, chlorine decay
ent stagnation times (i.e., 1-hr, 4-hr, 24-hr, 48-hr and 72- was also found to follow a first-order reaction (kCl = 0.16
hr). Since the field study revealed that flushing reduced the ± 0.01 hr−1 ), with no free chlorine residual detected af-
metal concentrations drastically, only unflushed samples ter 24 hr of stagnation. Both the DO and chlorine deple-
were analyzed during the bench-scale investigation. Results tion patterns observed are consistent with several other
illustrated in Fig. 5, at an alkalinity level of 100 mg/L as studies.[35,36]

240
220
200
Lead concentrations ( g/L)

180
160
140
120
100
80
60
40
20
0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75
Stagnation Time (hr)

pH 7.0 pH 7.5 pH 8.0 pH 9.0

Fig. 5. Effect of pH and stagnation time on total lead dissolution at an alkalinity of 100 mg/L as CaCO3.
Corrosion control in water supply systems 1257
2 concentrations were significantly lower, accounting for an
average of 11.6 ± 7.1% of the total lead content, almost
1.95
all samples (expect the control ones) were above the MAV
1.9
limit of 10 µg/L.[10] It is therefore apparent that only a small
fraction of the total lead was in the dissolved state which
1.85 indicates that the presence of insoluble lead compounds
ln(D O )

play an important role in water chemistry, as explained in


1.8 the next paragraph.
The amount of lead in drinking water depends upon
1.75
the solubility of several lead corrosion by-products such as
cerussite [PbCO3 ], hydrocerussite [Pb3 (OH)2 (CO3 )2 ] and
1.7
2
R = 0.9728 lead hydroxide [Pb(OH)2 ].[26,37,38] Although any of these
1.65
compounds can play an important role under a given pH
0 10 20 30 40 50 60 70 80 and alkalinity range, it has been reported that hydrocerus-
site is the dominant corrosion by-product formed in drink-
Downloaded by [University of California, San Diego] at 05:46 13 June 2015

Time (hr)

ing water supplies.[29,39] According to the Pourbaix dia-


Fig. 6. Typical dissolved oxygen (DO) depletion profile . gram, hydrocerussite is formed under typical conditions
existing in drinking water systems (i.e., pH 7.0 to 9.0 and
Impact of pH and alkalinity on lead release alkalinity 20 to 200 mg/L as CaCO3 ).[40] Moreover, con-
sidering the consistency of the experimental trends with
The total lead concentrations as a function of pH and al- the theoretical solubility model, which states that increased
kalinity, at the end of the 72-hr stagnation period, have lead solubility occurs at higher alkalinity values, it is rea-
been summarized in Fig. 7. Results demonstrated that at sonable to postulate that Pb3 (OH)2 (CO3 )2 may have played
the lower pH values tested (i.e., 7.0 and 7.5), an increase in a predominant part in controlling lead dissolution during
alkalinity resulted in a substantial increase in lead release. this study.
For instance, at pH 7.0, samples with alkalinity in the range
of 100 to 150 mg/L as CaCO3 , exhibited a lead concentra-
tion 3 to 5 times higher than that of the control, which had Impact of pH and alkalinity on copper release
an alkalinity of 20 mg/L as CaCO3 . The relative increase in Figure 8 illustrates the total copper concentrations as a
lead leaching became less pronounced as the pH increased, function of pH and alkalinity, at the end of the 72-hr stag-
and it was practically eliminated at pH 9.0. This suggests nation period. As was observed in the case of lead, at an
that at pH around neutrality, an increase in alkalinity pro- “elevated” alkalinity ranging between 100 and 150 mg/L as
motes lead dissolution and corrosion, while at pH > 8.0 CaCO3 , copper leaching displayed its highest potential at
the effect of alkalinity on lead leaching is marginal. With pH 7. At a given alkalinity level, the copper concentrations
respect to soluble lead behavior, a trend similar to the one measured declined progressively as the pH increased and
just described was observed. Although the corresponding reached their lowest values at pH 9.0. It is apparent that a

Fig. 7. Total lead concentration as a function of pH and alkalinity Fig. 8. Total copper concentration as a function of pH and alka-
at a stagnation time of 72 hr. linity at a stagnation time of 72 hr.
1258 Tam and Elefsiniotis

combination of pH around neutrality and alkalinity in the Table 4. Percent decrease in total lead concentration due to or-
100 to 150 mg/L as CaCO3 range can induce an adverse im- thophosphate addition as a function of pH and alkalinity.
pact on drinking water quality due to copper dissolution.
Decrease in total Pb concentration (%)
This phenomenon can be attributed to the formation of
soluble cupric bicarbonate [Cu(HCO3 )] and other carbon- Alkalinity (mg/L as CaCO3 )
ate complexes.[41−43] Regarding the soluble copper content,
it averaged 28.9 ± 12.1% of the total copper. Although pH 20 (Control) 100 125 150 Mean value
soluble copper appeared to be a comparatively “minor”
7.0 57.5 42.7 17.1 26.5 36.0 ± 15.4
component in the water; however, expressed as a percent- 7.5 88.2 73.9 67.9 64.5 73.6 ± 9.1
age, it was more than twice the corresponding soluble lead 8.0 81.4 67.6 60.5 68.8 70.0 ± 7.5
value of 11.6%. The concentration of copper detected in
oxidizing systems depends on the water chemistry and the
relative stabilities of copper (I) and copper (II) aqueous
species. According to the Pourbaix diagram, within the that the presence of orthophosphate reduced remarkably
typical pH range between 6.5 to 8.0 for water supplies, the lead dissolution potential.
Downloaded by [University of California, San Diego] at 05:46 13 June 2015

the principal stable copper species are cuprite [Cu2 O] and To further elucidate and quantify the performance of the
malachite [Cu2 (OH)2 CO3 ].[44] inhibitor, the percent decrease in total lead concentration
was computed at each condition studied, and the results
have been summarized in Table 4. Orthophosphate was
Effect of orthophosphate addition on lead and copper found to be much more effective at pH 7.5 or higher, result-
release ing in approximately 70% corrosion reduction. In contrast,
at a neutral pH, its effectiveness was reduced to almost half
The effect of orthophosphate addition (0.8 mg/L as P) of the above value (i.e., a 36% decrease). The variation in
as a corrosion inhibitor was investigated performing a se- alkalinity had a rather moderate effect on corrosion, with
ries of tests, at the same conditions as described previously the control samples demonstrating the greatest improve-
(Table 2). Since it was found that there was a limited metal ment at all pH values investigated. In addition, the soluble
leaching potential at pH 9.0 (Figs. 7 and 8), no further ex- lead concentrations were below the MAV limit of 10 µg/L
periments at this pH value were conducted. The total lead in all samples analyzed,[10] and accounted for only 2.4 ±
concentrations as a function of pH and alkalinity at the 1.9% of the total lead content.
end of the 72-hr stagnation period are depicted in Fig. 9. Figure 10 displays the total copper concentrations of the
Overall, the pattern was similar to that observed without or- tests with orthophosphate addition. As expected, the be-
thophosphate addition (Fig. 7), with lead reaching a peak havior was comparable to the one followed by lead (Fig. 9),
at the lowest pH of 7.0 and the highest alkalinity of 150 showing a sharp decrease in copper release as the pH in-
mg/L as CaCO3 tested. Comparing the relative concentra- creased in the high alkalinity samples, and very low levels
tion values in Figs. 7 and 9, however, it becomes apparent of copper in the control samples. The percent decrease in

Fig. 10. Total copper concentration as a function of pH and


Fig. 9. Total lead concentration as a function of pH and alkalinity alkalinity at a stagnation time of 72 hr (with orthophosphate
at a stagnation time of 72 hr (with orthophosphate addition). addition).
Corrosion control in water supply systems 1259
Table 5. Percent decrease in total copper concentration due to ues was apparent, with the lowest concentration obtained
orthophosphate addition as a function of pH and alkalinity. at pH 9.0. It was also found that dissolution of lead and
copper was intensified with an increase in alkalinity, par-
Decrease in total Cu concentration (%)
ticularly at pH ≤ 7.5. The addition of orthophosphate as
Alkalinity (mg/L as CaCO3 ) corrosion inhibitor was more effective at pH ≥ 7.5 resulting
in approximately 70% reduction in both lead and copper.
pH 20 (Control) 100 125 150 Mean value Overall, regarding the effect of pH and alkalinity, the pat-
terns were similar to those observed without orthophos-
7.0 11.5 48.1 9.7 40.9 27.6 ± 17.2
7.5 66.3 72.7 56.4 72.7 67.0 ± 6.7
phate addition, albeit at lower metal concentrations.
8.0 77.3 68.8 52.7 70.9 67.4 ± 9.1

Acknowledgments
total copper as a result of orthophosphate addition has
been included in Table 5. It is clear that application of the The authors wish to thank the staff at the Environmen-
Downloaded by [University of California, San Diego] at 05:46 13 June 2015

inhibitor reduced significantly the corrosion potential, by tal Engineering, and Environmental Science Laboratories,
an average of 67% at pH 7.5 and 8.0, and by 28% at pH 7.0. University of Auckland as well as at Watercare Service
Furthermore, soluble copper averaged at 21.9 ± 6.6% of Ltd. for their assistance throughout this study. Funding
the total copper detected, a slightly lower value than that was provided by a Technology in Industry Fellowship and
observed without orthophosphate addition. Watercare Services Ltd.
According to solubility models for lead and copper, the
optimal pH for orthophosphate film formation ranges be-
tween 7.0 and 8.0 on lead surfaces, and between 6.5 and
7.5 on copper surfaces.[8,41,45] Although the findings from References
this investigation are in agreement with respect to the opti-
[1] Brongers, M.P.H. Appendix K. Drinking water and sewer systems.
mal pH range indicated for lead, in the case of copper, or- Corrosion control costs and preventive strategies in the United States,
thophosphate addition exhibited the highest performance FHWA-RD-01-156, U.S. Department of Transportation Federal
at pH 8.0, which falls outside the predicted range. This Highway Administration, Houston, TX, 2002; 1–33.
study explored the corrosion pattern at a 72 hr stagnation [2] Cantor, A.F.; Bushman, J.B.; Glodoski, M.S.; Kiefer, E.; Bersch,
period; however, it is possible that metal dissolution behav- R.; Wallenkamp, H. Copper pipe failure by microbiologically influ-
enced corrosion. Mater. Perform. 2006, 45(6), 38–41.
ior could differ in the long term. As a final note, it has been [3] Churchill D.M.; Mavinic, D.S.; Neden, D.G.; MacQuarrie, D.M.
reported that although the use of orthophosphate may re- The effect of zinc orthophosphate and pH-alkalinity adjustment on
duce copper levels in the short term, in the long term the metal levels leached into drinking water. Can. J. Civil Eng. 2000,
formation of more stable protective films such as those con- 27(1), 33–43.
taining malachite [Cu2 (OH)2 CO3 ] and tenorite (CuO) may [4] Ahmedna, M.; Marshall, W.E.; Husseiny, A.A.; Rao, R.M.;
Goktepe, I. The use of nutshell carbons in drinking water filters
be prevented.[46−48] for removal of trace metals. Water Res. 2004, 38(4), 1062–1068.
[5] Clement, M.; Seux, R.; Rabarot, S. A practical model for estimat-
ing total lead intake from drinking water. Water Res. 2000, 34(5),
1533–1542.
Conclusions [6] Elci, L.; Kartal, A.A.; Soylak, M. Solid phase extraction method for
the determination of iron, lead and chromium by atomic absorption
The field investigation indicated that Auckland’s water is spectrometry using Amberite XAD-2000 column in various water
aggressive in nature, with a low alkalinity and hardness samples. J. Hazard. Mater. 2008, 153(1–2), 454–461.
content. Over 90% of the unflushed samples contained lead [7] United States Environmental Protection Agency. Revised Guidance
Manual for Selecting Lead and Copper Control Strategies. Report
above the corresponding MAV of 10 µg/L; however, copper No. EPA-816-R-06-001; U.S. Environmental Protection Agency,
was consistently well below the regulatory limits. Flushing Washington, DC, 2003.
reduced metal concentrations by more than 90%, with lead [8] Health Canada. Corrosion control in drinking water distribution
values below the MAV level. The bench-scale experiments systems. Canada, 2007. http://www.hc-sc.gc.ca/ewh-semt/alt
revealed that the lead and copper leaching behavior fol- formats/hecs-sesc/pdf/pubs/water-eau/consultation/corrosion/
corrosion-eng.pdf (accessed April 2009).
lowed similar trends during the stagnation period, showing [9] World Health Organization. Guidelines for Drinking-water Quality
a sharp increase within the first 24 hr and then approaching 3rd Ed. Volume 1: Recommendations. WHO, Geneva, Switzerland,
an equilibrium state. The relatively small soluble fraction 2004; 145–196.
for each metal (12% for lead, 29% for copper) denotes that [10] New Zealand Ministry of Health. Drinking-water Standards for
the role of non-soluble compounds is prevalent during dis- New Zealand 2005, The Ministry of Health: Wellington, New
Zealand, 2005.
solution. Furthermore, the highest degree of leaching was [11] Cardew, P.T. A method for assessing the effect of water quality
consistently observed at the lowest pH 7.0 tested. As pH changes on plumbosolvency using random daytime sampling. Water
increased, a considerable decrease in lead and copper val- Res. 2003, 37(12), 2821–2832.
1260 Tam and Elefsiniotis
[12] Xiao, W.; Hong, S.; Tang, Z.; Taylor, J.S. Effects of blending on total [31] Korshin, G.V.; Ferguson, J.F.; Lancaster, A.N. Influence of natural
copper release in distribution systems. J. Am. Water Works Assoc. organic matter on the corrosion of lead brass in potable water.
2007, 99(1), 78–88. Corros. Sci. 2000, 42(1), 53–56.
[13] Boyd, G.R.; Shetty, P.; Sandvig, A.M.; Pierson, G.L. Lead in tap [32] Korshin, G.V.; Ferguson, J.F.; Lancaster, A.N. Influence of natu-
water following simulated partial lead pipe replacements. J. Environ. ral organic matter on the morphology of corroding lead surfaces
Eng. 2004, 130(10), 1188–1197. and behavior of lead-containing particles. Water Res. 2005, 39(5),
[14] Melidis, P.; Sanozidou, M.; Mandusa, A.; Ouzounis, K. Corrosion 811–818.
control by using indirect methods. Desalination, 2007, 213(1–3), [33] Britten, A.; Richards, W.N. Factors influencing plumbosolvency in
152–158. Scotland. J. Instit. Water Eng. Sci. 1981, 35(4), 349–364.
[15] Prisyazhniuk, V. Prognosticating scale-forming properties of water. [34] Hulsmann, A.D. Particulate lead in water supplies. Journal of the
Appl. Thermal Eng. 2007, 27(8–9), 1637–1641. Institution of Water and Environ. Manage. 1990, 4, 19–25.
[16] Al-Rawajfeh, A.E.; Al-Shamaileh, E.M. Inhibition of corrosion [35] Boulay, N.; Edwards, M. Role of temperature, chlorine and organic
in steel water pipes by ammonium pyrrolidine dithiocarbamate matter in copper corrosion by-product release in soft water. Water
(APDTC). Desalination, 2007, 206(1–3), 169–178. Res. 2001, 35(3), 685–690.
[17] Chan, M.W.M. A critical study on corrosion and corrosion control [36] Huang, Y.H.; Zhang, T.C. Effects of dissolved oxygen on formation
methods for piping systems in Hong Kong. Ph.D. Thesis. The Hong of corrosion products and concomitant oxygen and nitrate reduc-
Kong Polytechnic University, Department of Building Services En- tion in zero-valent iron systems with or without aqueous Fe2+ .
Downloaded by [University of California, San Diego] at 05:46 13 June 2015

gineering, Hong Kong, 2003. Water Res. 2005, 39(9), 1751–1760.


[18] Edwards, M.; McNeill, L.S. Effect of phosphate inhibitors on [37] Leroy, P. Lead in drinking water—Origins; solubility; treatment. J.
lead release from pipes. J. Am. Water Works Assoc. 2002, 94(1), Water Supply: Res. Technol.—Aqua, 1993, 42(4), 223–238.
79–90. [38] Miranda M.L.; Kim, D.; Hull, A.P.; Paul, C.J.; Overstreet Galeano,
[19] Edwards, M. Controlling corrosion in drinking water distribution M.A. Changes in blood lead levels associated with use of chlo-
systems: a grand challenge for the 21st century. Water Sci. Technol. ramines in water treatment systems. Environ. Health Perspect. 2007,
2004, 49(2), 1–8. 115(2), 221–225.
[20] McNeil, L.S.; Edwards, M. Phosphate inhibitors and red water stag- [39] Schock, M.R.; Wagner, I.; Oliphant, R.J. The corrosion and sol-
nant pipes. J. Environ. Eng. 2000, 126(12), 1096–1102. 1 ubility of lead in drinking water.” In: Internal Corrosion of Water
[21] Becker, A. The effect of corrosion inhibitors in drinking water in- Distribution Systems. 2nd Ed.; AWWA Research Foundation and
stallations of copper. Mater. Corros. 2002, 53(8), 560–567. DVGW Technologiezentrum Wasser, Denver, CO, 1996, 213–308.
[22] Ebrahimi Mehr, M.; Shahrabi, T.; Hosseini, M.G. Determination of [40] Conio, O.; Ottaviani, M.; Formentera, V.; Lasagna, C.; Palumbo,
suitable corrosion inhibitor formulation for a potable water supply. F. Evaluation of the lead content in water for human consumption.
Anti-Corros. Methods Mater. 2004, 51(6), 399–405. Microchem. J. 1996, 54(4), 355–359.
[23] Tam, Y.S. Investigation into the options to control plumbosolvency [41] United States Environmental Protection Agency. Effect of pH, DIC,
in drinking water. Master of Engineering Thesis. University of orthophosphate and sulfate on drinking water cuprosolvency, Report
Auckland, Department of Civil and Environmental Engineering, No. EPA/600/R-95/085; U.S. Environmental Protection Agency:
Auckland, New Zealand, 2007. Cincinnati, OH, 1995.
[24] HACH Water Analysis Handbook-ISO9001 Certified. HACH [42] Edwards, M.; Jacobs, S.; Dodrill, D.M. Desktop guidance for miti-
Company, Loveland, CO, USA, 1997. gating Lead and Copper corrosion by-products. J. Am. Water Works
[25] Berthouex, P.M.; Brown, L.C. Statistics for Environmental Engi- Assoc. 1999, 91(5), 66–77.
neers. 2nd Edition, Lewis Publishers, Boca raton, FL, USA, 2002. [43] Letelier, M.V.; Lagos, G.E.; Reyes, A. Chemical characterization
[26] Schock, M.R. Response of lead solubility to dissolved carbon- of blue stains in domestic fixtures in contact with drinking water.
ate in drinking water. J. Am. Water Works Assoc. 1980, 72(12), Environ. Monitor. Assess. 2008, 139(1–3), 307–315.
695–704. [44] Merkel, T.H.; Grosz H.-J.; Werner W.; Dahlke T.; Reicherter
[27] Dodrill, D.M.; Edwards, M. Corrosion control on the basis S.; Beuchle G.; Eberle S.H. Copper corrosion by-product re-
of utility experience. J. Am. Water Works Assoc. 1995, 87(7), lease in long-term stagnation experiments. Water Res. 2002, 36(6),
74–85. 1547–1555.
[28] Benjamin, M.M.; Sontheimer, H.; Leroy, P. Corrosion of iron and [45] Schock, M.R. Understanding corrosion control strategies for lead.
steel. In: Internal Corrosion of Water Distribution Systems, 2nd J. Am. Water Works Assoc. 1989, 81(7), 88–100.
Ed.; AWWA Research Foundation and DVGW Technologiezen- [46] Schock, M.R.; Clement, J.A. Lead and copper control with non-
trum Wasser: Denver, CO, 1996; 29–68. zinc orthophosphate. J. New England Water Works Assoc. 1998,
[29] Tang, Z. Effect of source water blending on iron and lead release: 112(1), 20–33.
thermodynamic and statistical modeling. Ph.D. Thesis. University [47] Edwards, M.; Sprague, N. Organic matter and copper corrosion
of Central Florida, Civil and Environmental Department, Orlando, by-product release: A mechanistic study. Corros. Sci. 2001, 43(1),
FL, 2003. 1–18.
[30] Volk, C.J. Corrosion control in drinking water systems. In: Water [48] Cantor, A.F.; Park, J.K.; Vaiyavatjamai, P. Effect of chlorine on
Encyclopedia, John Wiley & Sons, Inc.: New York, 2005; 5–11. corrosion. J. Am. Water Works Assoc. 2003, 95(5), 112–123.

You might also like