logo
CONTENT TYPES

Figure 1Loading Img

Release of Pb(II) from Monochloramine-Mediated Reduction of Lead Oxide (PbO2)

View Author Information
Department of Civil and Environmental Engineering, University of Iowa, Iowa City, Iowa 52242-1527
* Corresponding author phone: (65) 6516-4729 (Y.P.L.), (319) 335-5653 (R.L.V.); e-mail: [email protected] (Y.P.L.), [email protected] (R.L.V.).
†Current Address: Division of Environmental Science and Engineering, Faculty of Engineering, National University of Singapore, Singapore 117576
Cite this: Environ. Sci. Technol. 2008, 42, 24, 9137–9143
Publication Date (Web):November 11, 2008
https://doi.org/10.1021/es801037n
Copyright © 2008 American Chemical Society
Subscribed Access
Article Views
730
Altmetric
-
Citations
LEARN ABOUT THESE METRICS

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

PDF (2 MB) OpenURL UNIV OF NORTH TEXAS
Supporting Info (1)»

Abstract

A contributing factor causing the sudden release of excessive lead into drinking water is believed to involve the change in redox conditions occurring when monochloramine (NH2Cl) replaces free chlorine as a disinfectant. Studies suggest that NH2Cl cannot effectively oxidize Pb(II) to form PbO2, a Pb(IV) mineral scale formed from the oxidation of metallic lead and Pb(II) species by free chlorine. Unexpectedly, we observed that NH2Cl is actually capable of reducing PbO2 to form Pb(II). We systematically investigated this reaction by varying important water chemistry factors such as solution pH, total carbonate concentration, and the Cl/N molar ratio to control chloramine speciation and its rate of decomposition via a complex set of autodecomposition reactions. The amount of Pb(II) formed was found to be proportional to the amount of NH2Cl that autodecomposed regardless of the rate of this reaction. This implies that the rate of Pb(II) release is proportional to the absolute rate of NH2Cl decomposition. It is proposed that the species responsible for the reduction of PbO2 is likely a reactive intermediate produced during the decay of NH2Cl. This finding is the first to report that NH2Cl can act as a reductant.

Synopsis

Monochloramine, a disinfectant generally presumed an oxidant, can act as a reductant via a mechanism hypothesized to involve an intermediate formed from its decomposition.

Introduction

ARTICLE SECTIONS
Jump To

The change of disinfectant from free chlorine to monochloramine has been a strategy adopted by water utilities with difficulties to meet the standards of disinfection byproduct (DBPs) and maintain a sufficient disinfectant residual in the distribution system (1).
Recent public concern on this practice aroused from the abrupt rise of lead levels in Washington, DC, in 2003 (2, 3). Lead concentration as high as 4800 μg/L, about 300 times higher than the 15 μg/L action level of the 1991 Lead and Copper Rule set by the U.S. EPA (4), was found in the customer residence after the water utility adopted NH2Cl for secondary disinfection (3). Free chlorine is a strong oxidant and has been demonstrated to be able to oxidize metallic Pb((0)) and Pb(II) ion or solid phases to form PbO2, a Pb(IV) solid phase (2, 3, 5-9). PbO2 is a strong oxidant and almost insoluble (5). Its formation on inside pipes may help to reduce soluble Pb concentration to an acceptable level in systems that historically use free chlorine to maintain a disinfectant residual in the distribution system. It is believed that the switch to NH2Cl, a weaker oxidant than free chlorine, creates an environment with a lower oxidation potential that may limit the formation of PbO2 or alter the stability of PbO2(2, 10). The mechanisms and processes involving this release of lead are not yet well understood. However, recent work has shown that PbO2 is not stable in water and is slowly reduced, releasing Pb(II) into water (11). It is hypothesized that this is caused by the oxidation of water itself. In addition, natural organic matter is also capable of reducing PbO2(7, 11).
The initial motivation of this work was to investigate the influence of NH2Cl on the stability of PbO2 in water. The initial hypothesis was that the presence of NH2Cl would not significantly influence the rate of PbO2 reduction and subsequent Pb(II) formation. If anything, a minor reduction in the rate of Pb(II) release was expected because of the minor oxidation of Pb(II) resulting from the water-induced PbO2 reduction by the extremely low free chlorine concentration that exists in the NH2Cl equilibrium (Table S1, Supporting Information). This hypothesis was based on a thermodynamic argument that NH2Cl is slightly lower in redox potential than PbO2 around neutral pH so that NH2Cl can not directly oxidize Pb(II) to PbO2(10). Quite unexpectedly, we observed that the presence of NH2Cl greatly increased the rate of PbO2 reduction.
NH2Cl is not stable in water and autodecomposes primarily to nitrogen by a complex set of reactions (12, 13). While a comprehensive reaction model incorporating these reactions has been developed to successfully describe the autodecomposition of NH2Cl, some of these reactions have only been hypothesized to involve unidentified intermediates that must be redox active (12, 14, 15). Identification of these intermediates has been shown challenging (16-18), and only a few, such as amidogen radical (NH2) and hydrazine (N2H4), have been proposed (18). Both oxidation and reduction of substances are therefore theoretically possible in a system involving multiple oxidants and reductants. For example, both oxidations and reductions of organic substances have been observed in solutions of Fenton’s reagent, typically used as a source of the powerful oxidant hydroxyl radical (19-24).
It is hypothesized that the autodecomposition of NH2Cl, which results in its slow loss in solutions, can also cause the reduction of PbO2, presumably via the reduction by reactive intermediates. Furthermore, consideration of the NH2Cl autodecomposition mechanism also supports the idea that these intermediates should exist at pseudo-steady-state at exceedingly low concentrations in proportion to the rate of NH2Cl decay. If the rate of PbO2 reduction by an intermediate is first order in the concentration of intermediate, then the overall rate of PbO2 reduction should occur in proportion to the rate of NH2Cl decay resulting in a simple apparent stoichiometry between the amount of NH2Cl decomposed and the amount of Pb(II) produced for a given concentration of PbO2. The variables that increase the decay of NH2Cl should therefore also increase the rate of PbO2 reduction and Pb(II) formation.
The detailed reactions of NH2Cl autodecomposition is provided in the Supporting Information (Table S1). The overall rate of NH2Cl decay has been approximated by a second-order relationship (25)(1) where kvcsc is a water quality-dependent coefficient which can be described as (2) where CT and [NH3]T are total carbonate and free ammonia concentrations, respectively; α0 and α1 are the ionization constants for carbonate system, and α0,N is the ionization constant for the neutral ammonia species, which are all pH dependent; kH+, kH2CO3, and kHCO3 are rate constants that account for the general acid catalysis of NH2Cl loss; k3 is a rate constant characterizing the reaction between NH2Cl and hypochlorous acid to form dichloramine; Ke is an equilibrium constant describing dichloramine and hypochlorous acid equilibrium.
This approximation can be derived from the more comprehensive mechanism by assuming that the rate limiting reactions involve the formation and subsequent loss of dichloramine, which is maintained at pseudo-steady state (25-27). This assumption is reasonable only at pH 8 or above at NH2Cl concentrations generally less than about 0.1−0.2 mM. It is presented only as a way of easily seeing the influence of several water quality parameters on NH2Cl decay, not as an exact predictor. Integration of this relationship yields only an approximate relationship between NH2Cl concentration and time under all reaction conditions because free ammonia concentration increases as NH2Cl decomposes (25-27). However, this relationship is still useful in understanding monochloramine decay kinetics. This relationship indicates that the absolute rate of NH2Cl decay increases with increasing initial NH2Cl concentration, carbonate concentration, which acts as a general acid catalyst (28), and Cl/N molar ratio which influences the rate of NH2Cl hydrolysis. The decay of NH2Cl also increases with decreasing pH because of the acid-catalyzed dichloramine formation.
The objective of this study was to investigate how the reduction of PbO2 as measured by the subsequent formation of Pb(II) depends on the rate and extent of NH2Cl decay under conditions where its overall loss was governed by autodecomposition. Systematic studies were conducted over a range of water quality parameters including initial NH2Cl concentration, pH, total carbonate concentration, and Cl/N molar ratio that can greatly influence the rate of NH2Cl decay and encompass the conditions typically encountered in drinking water supplies subjected to chloramination.

Material and Methods

ARTICLE SECTIONS
Jump To

Chemicals

Reagent grade ammonia chloride and sodium bicarbonate (Fisher Scientific) were used as the source ammonia and total carbonate. Deionized water obtained from a Barnstead ULTRO pure water system (Barnstead-Thermolyne Corp.) was used to prepare working solutions. Stock NH2Cl (∼3.95 mM or 280 mg/L as Cl2) was freshly prepared by the addition of concentrated NaOCl solution (∼1 M, Fisher Scientific) to a carbonate-buffered ammonium chloride solution according to the published procedures (14). This NH2Cl stock solution was calibrated by the DPD-FAS method before use. Ammonia concentration was varied to prepare NH2Cl solutions with different Cl/N molar ratios.

Lead Oxide

PbO2 particles were prepared in our laboratory as follows: To 250 mL of 0.1 M Pb(NO3)2 solution, the desired amount concentrated NaOCl solution (∼1M, Fisher Scientific) was added to make a solution with a NaOCl/Pb2+ molar ratio of 1.1. A white-yellow solid phase quickly precipitated and turned into brown-reddish color as the reaction proceeded. After 24 h, the brown-reddish slurry was transferred to several 50 mL polypropylene tubes and centrifuged at 6000 rpm for 6 min. The supernatant was decanted and replaced with deionized water. The new slurry was centrifuged again. The above procedure was repeated several times to remove Pb2+ as much as possible. The solid phase was then transferred to a dialysis membrane tube (MWCO = 6−8000, Spectra/Por) submerged in a DI water bath with regular change of the DI water for three days to remove residual Pb2+. The solid phase was collected in polypropylene tubes followed by centrifuge and decanting the supernatant. Dry particles were obtained by freeze-drying the solid phase. Final solid phase weighted 3.34 g, approximately 55.8% recovery of the added Pb2+. Most of the unrecovered lead was due to the attachment of particles on the container walls and loss during the solid transfer. A dialysis membrane was used to remove residual Pb(II) because PbO2 is reduced to Pb(II) at low pH (11).
Scanning electron microscope (SEM) images of the PbO2 particles is shown in Figure S1 in the Supporting Information. The size of individual particles was less than 100 nm. However, the primary particles appeared to be composed of small particles that had aggregated. The specific surface area determined by 7-point BET method was 24.29 m2/g. X-ray diffraction (XRD) analysis of the particles is shown in Figure S2 in the Supporting Information. The diffraction pattern indicated that the laboratory-prepared PbO2 is plattnerite, a tetragonal polymorph of PbO2 that has been found in distribution systems (6, 29)

PbO2 Reduction Studies

Experiments were conducted headspace-free using 125 mL polypropylene bottles covered by aluminum foil at 25 °C. All experiments utilized 10 mg/L of PbO2, an amount which did not significantly influence the rate of NH2Cl autodecomposition. Solution pH values were adjusted by 1 N HNO3 and NaOH. Experimental variables investigated including initial NH2Cl concentration (14.1, 42.3, and 70.4 μM or 1, 3, and 5 mg/L as Cl2), solution pH (6.0, 7.0, and 8.0), total carbonate concentration from added bicarbonate (CT = 2, 5 and 10 mM) and Cl/N molar ratio (0.2, 0.5 and 0.7). After they were filled with the working solution, the polypropylene bottles were placed on a shaking table rotating at 200 rpm. The Pb(II) concentration, NH2Cl residual, and solution pH value were periodically measured over a course of 7 days. The variations of pH values during the course of the experiments were within target pH ±0.2.

Analytical Methods

NH2Cl concentration was measured by a modified iodometric method. Because iodide can react with PbO2 to produce a positive interference in this method (30), 5 mL of solution was filtered by a nylon syringe filter with a 0.2 μm pore size (Fisher Scientific) to remove PbO2 particle aggregates before NH2Cl measurement; 50 μL of acetate buffer was added to the filtrate to adjust the pH to about 4. Sufficient potassium iodide was then added to the filtrate to allow the oxidation of iodide by NH2Cl. The triiodate (I3) liberated from the oxidation was determined by its UV absorbance at 351 nm using a predetermined molar absorbance of 23 325 cm−1 M−1. This method showed comparable results with the widely used DPD-FAS method (less than 10% difference) with the advantage that a much smaller sample volume was required.
A Nano-Band Explorer II (TraceDetect, Seattle) anodic stripping voltammetry (ASV) was used for total Pb(II) measurements (11). The pH value of unfiltered sample taken from the well mixed solution was adjusted to 4 by addition 0.1 M acetate buffer. At this pH value, colloidal and Pb(II) carbonate solid phases quickly dissolve to soluble Pb2+ ion (31, 32). Thus, total Pb(II) including soluble, colloidal, and solid Pb(II) is measured by the ASV. A Hitachi S-4000 scanning electron microscope (SEM) was used for acquiring the images of PbO2 particles. The Xay-ray diffraction (XRD) pattern was determined by a MiniFlex II desktop X-ray diffractometer (Rigaku Americas). The solution pH value was measured by an Accument pH meter (A15, Fisher Scientific) coupled with a calomel combination pH electrode pre-equilibrated by standard buffers.

Results and Discussion

ARTICLE SECTIONS
Jump To

PbO2 Reduction Dependence on Initial NH2Cl Concentration and NH2Cl Decay

Figure 1 graphically shows the loss of NH2Cl concentration, formation of total Pb(II) with time, and the amount of total Pb(II) produced as a function of the change in NH2Cl concentration for experiments conducted using initial NH2Cl concentration of 14.1, 42.3, and 70.4 μM (1.0, 3.0, and 5.0 mg/L as Cl2). The decay of NH2Cl approximately followed the second-order kinetics described in eq 1 and the decay rate was faster at the higher initial NH2Cl concentration (see Figure S2 in Supporting Information for the second-order kinetics plot). It should be pointed out that the presence of 10 mg/L PbO2 only slightly influenced the rate of NH2Cl decay (Figure 1a). The rate of total Pb(II) formation increased with increasing initial NH2Cl concentration (Figure 1b), which was concurrent with the trend of NH2Cl decay rate. A linear relationship between total Pb(II) formation and loss in NH2Cl was observed (Figure 1c), which is consistent with the hypothesis that PbO2 reduction is caused by a reaction with an intermediate that is linearly related to the decay of NH2Cl. In a control experiment in the absence of NH2Cl, Pb(II) was formed but at a greatly reduced rate, with a concentration of 0.31 μM (65 μg/L) formed after 7 days (Figure 1b). This has been attributed to a water-induced PbO2 reduction (11). We hypothesize that the water reaction may occur in parallel to the NH2Cl-induced PbO2 reduction, although it may not be significant when a high concentration of NH2Cl is present. The SEM image of the solid phase after the experiment with 70.4 μM NH2Cl is shown in Figure 2. A secondary solid phase was formed discretely. The identity of this solid was not certain because of the limited amount of sample that could not be analyzed by XRD. However, based on the solution composition and solubility consideration, this solid phase should be either cerrusite (PbCO3) or hydrocerrusite (Pb3(CO3)2(OH)2) (33). The discrete solid phase did not form a passivating layer to physically protect PbO2 from being reduced to form Pb(II) in NH2Cl solutions. However, it can provide a sink to regulate soluble Pb(II) concentration (i.e., Pb2+ and Pb(II)−carbonate complex).

Figure 1

Figure 1. Effect of NH2Cl concentration on the formation of total Pb(II) and NH2Cl decay: (a) NH2Cl decay as a function of time, (b) total Pb(II) formation as a functiom of time, and (c) total Pb(II) formation vs NH2Cl decay (data of duplicate experiments were shown by separate points). Fixed experimental parameters: PbO2 = 10 mg/L, pH 7.0, CT = 5 mM, Cl/N = 0.7. In control experiment for NH2Cl decay, no PbO2 was added.

Figure 2

Figure 2. SEM image of the solid phase after 7 day experiment. NH2Cl = 70.4 μM, PbO2 = 10 mg/L, pH 7.0, CT = 5 mM, Cl/N = 0.7.

Influence of Solution pH

Figure 3 summarizes the results for experiments conducted at pH 6.0, 7.0, and 8.0 at a constant initial NH2Cl concentration, carbonate concentration, and Cl/N ratio. The Pb(II) formation from control experiments in the absence of NH2Cl is shown in the Supporting Information. The decay of NH2Cl is controlled by the rate-determining step that is believed to be the formation of dichloramine (NHCl2), a product of NH2Cl hydrolysis and disproportionation that quickly decomposes upon its formation (12, 34). Dichloramine formation is enhanced at lower pH and, in turn, leads to the faster decay of monochloramine at lower pH as observed in our experiments (Figure 3a). For the control experiment without PbO2 conducted at pH 7.0, the rate of NH2Cl decay was only slightly slower than that in the presence of PbO2. The rates of total Pb(II) formation increased with decreasing solution pH value. At pH 6.0, the rate of total Pb(II) formation sharply decreased after 2 days (Figure 3b), which was accompanied by a decrease in NH2Cl decomposition rate. The exhaustion of the most reactive sites on the PbO2 surface may also contribute to the decrease in total Pb(II) formation. Figure 3c again shows that total Pb(II) formation is linearly related to the amount of NH2Cl decomposed.

Figure 3

Figure 3. Effect of solution pH on the formation of total Pb(II) and NH2Cl decay: (a) NH2Cl decay as a function of time, (b) total Pb(II) formation as a function of time, and (c) total Pb(II) formation vs NH2Cl decay vs total Pb(II) formation (data of duplicate experiments were shown by separate points). PbO2 = 10 mg/L, NH2Cl = 42.3 μM, CT = 5 mM, Cl/N = 0.7. In control experiment for NH2Cl decay, no PbO2 was added.

Influence of Total Carbonate Concentration, CT

The rates of NH2Cl decay and total Pb(II) formation increased with increasing total carbonate concentration at a constant pH and Cl/N molar ratio (Figure 4a and b). The increasing rate of NH2Cl decay with increasing CT can be attributed to the general-acid catalyzed nature of the NH2Cl disproportionation (14, 28). Valentine and Jafvert (28) experimentally determined the general-acid catalyzed NH2Cl disproportionation by sulfate and phosphate, and proposed that carbonate may act at the same fashion based on a linear free energy relationship (LFER), which was later experimentally demonstrated by Vikesland et. al (14). Our results are consistent with this nature of NH2Cl decay. In the control experiment without PbO2 conducted at 10 mM, CT, no statistically significant difference was observed for the NH2Cl decay when comparing to that in the presence of PbO2. Figure 4c shows that the amount of lead formed is linearly related to NH2Cl decay.

Figure 4

Figure 4. Effect of total carbonate concentration on the formation of total Pb(II) and NH2Cl decay: (a) total Pb(II) formation as a functiom of time, (b) NH2Cl decay as a function of time, and (c) NH2Cl decay vs total Pb(II) formation (data of duplicate experiments were shown by separate points). PbO2 = 10 mg/L, NH2Cl = 42.3 μM, pH 7.0, Cl/N = 0.7. In control experiment for NH2Cl decay, no PbO2 was added.

Influence of Cl/N Molar Ratio

The rates of NH2Cl decay and total Pb(II) formation increased with increasing Cl/N molar ratio (Figure 5a and b). The free ammonia content of the working solution increased as the Cl/N ratio decreased because a fixed free chlorine concentration was used to prepare these NH2Cl solutions. Ammonia is the product of the NH2Cl hydrolysis and disporportionation, the reactions that lead to the decay of NH2Cl (12). The stability of NH2Cl increases with increasing ammonia concentration (or decreasing Cl/N molar ratio) because of the greater backward rates of NH2Cl hydrolysis and disporportionation (14). The slower decay rate of NH2Cl at a smaller Cl/N molar ratio was consistent with this reaction scheme of NH2Cl decomposition. Figure 5(c) shows a similar linear relationship between Pb(II) formation and NH2Cl decay.

Figure 5

Figure 5. Effect of Cl/N molar ratio on the formation of total Pb(II) and NH2Cl decay: (a) total Pb(II) formation as a functiom of time, (b) NH2Cl decay as a function of time, and (c) NH2Cl decay vs total Pb(II) formation (data of duplicate experiments were shown by separate points). PbO2 = 10 mg/L, NH2Cl = 42.3 μM, pH 7.0, CT = 5 mM. In control experiment for NH2Cl decay, no PbO2 was added.

Comparison of Correlations between PbO2 Reduction and NH2Cl Autodecomposition

The formation of total Pb(II) was related to NH2Cl loss by a single linear relationship that was independent of how the rate of NH2Cl autodecomposition was controlled (Figure 6). The overall slope of 0.204(±0.011) is statistically identical to the slopes of the linear relationships obtained by varying one parameter at a time (Table 1).

Figure 6

Figure 6. Summary plot of total Pb(II) formation vs NH2Cl decay.

Table 1. Equations of Least-Square Linear Regression between Total Pb(II) Formation and ΔNH2Cl for Each Water Parameter with a 95% Confidence Level in the Slopea
water parameterleast-square linear regression equation
NH2Cly = 0.169(±0.015)x − 0.039, r2 = 0.938
pH valuey = 0.210(±0.018)x − 0.046, r2 = 0.947
total carbonatey = 0.206(±0.020)x − 0.119, r2 = 0.931
Cl/N molar ratioy = 0.173(±0.022)x − 0.038, r2 = 0.908
a

y: total Pb(II) formation (μM). x: ΔNH2Cl (μM).

The linear relationship supports the hypothesis that the autodecomposition of NH2Cl leads to the formation of total Pb(II) (or the reduction of PbO2) and that the product(s) from the NH2Cl decomposition should be responsible for this instead of a direct reaction with NH2Cl. On the basis of this linear relationship, the rate of total Pb(II) formation in NH2Cl solutions can be approximated by the following equation: (3) where α is a constant approximately equal to 0.2 in the present study. While it is tempting to interpret the constant as a measure of a true stoichiometry of approximately 1 μM Pb(II) formed to 5 μM of NH2Cl decomposed, this is not likely because the value of α is expected to depend on the type, size, and concentration of PbO2(30) and can only be determined by experiments. Increasing the amount of PbO2 should increase the amount of total Pb(II) formed in the presence of independently autodecomposing monochloramine.
The nature of the reaction causing the formation of Pb(II) in the presence of NH2Cl is hypothesized to involve a reaction of a reactive intermediate produced during the relatively slow NH2Cl autodecomposition. The reduction of PbO2 by free ammonia can be ruled out because we observed a slower rate of total Pb(II) formation at lower Cl/N ratio (or higher free ammonia concentration). It should be noted that the rate of NH2Cl decomposition under our experimental conditions was approximately second-order with respect to NH2Cl concentration (see Figures S3−S6 in Supporting Information); therefore, the formation of total Pb(II) should also be similarly approximately second-order with respect to NH2Cl concentration as shown in eq 3.

Environmental Significance

Monochloramine has been widely considered as an alternative disinfectant by water utilities to combat the formation of DBPs when an unacceptable level occurs. However, field evidence has suggested that NH2Cl may alter the stability of lead-containing scale, primarily PbO2 and result in a hazardous level of lead in drinking water (2, 3, 35). In this study, we revealed that total Pb(II) formation from the reduction of PbO2 is attributed to the decomposition of NH2Cl, presumably by an unidentified intermediate formed in the autodecomposition reactions. The water chemistry that enhances the decomposition, including a higher initial NH2Cl concentration, lower solution pH, higher total carbonate concentration, and higher Cl/N molar ratio, can in turn enhance the formation of total Pb(II). These water parameters need to be carefully evaluated for lead release when a water utility attempts to use monochloramine for secondary disinfection. Additional studies focused on the identification of the intermediate that leads to PbO2 reduction are warranted because they may help to develop a strategy to eliminate the undesired lead release while still take the advantage of monochloramine in water treatments.

Supporting Information

ARTICLE SECTIONS
Jump To

Additional table of NH2Cl autodecomposition reactions and figures of PbO2 SEM images, and second-order plots of monochloramine decay. This material is available free of charge via the Internet at http://pubs.acs.org.

Terms & Conditions

Electronic Supporting Information files are available without a subscription to ACS Web Editions. The American Chemical Society holds a copyright ownership interest in any copyrightable Supporting Information. Files available from the ACS website may be downloaded for personal use only. Users are not otherwise permitted to reproduce, republish, redistribute, or sell any Supporting Information from the ACS website, either in whole or in part, in either machine-readable form or any other form without permission from the American Chemical Society. For permission to reproduce, republish and redistribute this material, requesters must process their own requests via the RightsLink permission system. Information about how to use the RightsLink permission system can be found at http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

Acknowledgment

ARTICLE SECTIONS
Jump To

This work was supported by a grant from the American Water Works Association Research Foundation (Project No. 3172). The assistance of ASV measurements from TraceDetect was greatly appreciated.

References

ARTICLE SECTIONS
Jump To

This article references 35 other publications.

  1. 1

    American Water Works Association. Water Quality and Treatment: A Handbook of Community Water Supplies; 5th edition; McGraw-Hill, Inc.: New York,

    1999.
  2. 2
    Renner, R. Plumbing the depths of DCʼs drinking water crisis Environ. Sci. Technol. 2004, 38, 224A 227A
  3. 3
    Edwards, M.; Dudi, A. Role of chlorine and chloramine in corrosion of lead-bearing plumbing materials J. Am. Water Works Assoc. 2004, 96, 69 81
  4. 4
    . Maximum Contamination Level Goals and National Primary Drinking Water Regulations for Lead and Copper. Final Rule. Fed. Regist. 1991, 56, 2646026564.
  5. 5
    Lytle, D. A.; Schock, M. R. Formation of Pb(IV) oxides in chlorinated water J. Am. Water Works Assoc. 2005, 97, 102 114
  6. 6
    Schock, M. R.. New Insights into lead corrosion control and treatment change impacts (with some considerations towards Cu). AWWA Meeting Section Emerging Issues in Water Treatment, Michigan, 2007;
  7. 7
    Dryer, D. J.; Korshin, G. V. Investigation of the reduction of lead dioxide by natural organic matter Environ. Sci. Technol. 2007, 41, 5510 5514
  8. 8
    Schock, M. R.; Scheckel, K. G.; DeSantis, M.; Gerke, T. L. Mode of occurrence, treatment and monitoring significance of tetravalent lead. Presented at the AWWA Water Quality Technology Conference, Quebec, Canada, 2005.
  9. 9
    Schock, M. R.; Harmon, S. M.; Swertfeger, J.; Lohmann, R. Tetravalent lead: A hitherto unrecognized control of tap water lead contamination. Presented at the AWWA Water Quality Technology Conference, Nashville, TN, 2001.
  10. 10
    Rajasekharan, V. V.; Clark, B. N.; Boonsalee, S.; Switzer, J. A. Electrochemistry of free chlorine and monochloramine and its relevance to the presence of Pb in drinking water Environ. Sci. Technol. 2007, 41, 4252 4257
  11. 11
    Lin, Y. P.; Valentine, R. L. The release of lead from the reduction of lead oxide (PbO2) by natural organic matter Environ. Sci. Technol. 2008, 42, 760 765
  12. 12
    Jafvert, C. T.; Valentine, R. L. Reaction scheme for the chlorination of ammoniacal water Environ. Sci. Technol. 1992, 26, 577 586
  13. 13
    Ozekin, K.; Valentine, R. L.; Vikesland, P. J. Modeling the decomposition of disinfecting residuals of chloramine. InWater Disinfection and Natural Organic Matter; ACS Symposium Series 649; American Chemical Society: Washington, DC, 1996; pp 115125.
  14. 14
    Vikesland, P. J.; Ozekin, K.; Valentine, R. L. Monochloramine decay in model and distribution system waters Water Res. 2001, 35, 1766 1776
  15. 15
    Duirk, S. E.; Gombert, B.; Croue, J. P.; Valentine, R. L. Modeling monochloramine loss in the presence of natural organic matter Water Res. 2005, 39, 3418 3431
  16. 16
    Valentine, R. L.; Brandt, K. I.; Jafvert, C. T. A spectrophotometric study of the formation of an unidentified monochloramine decomposition product Water Res. 1986, 20, 1067 1074
  17. 17
    Diyamandoglu, V.; Selleck, R. E. Reactions and products of chloramination Environ. Sci. Technol. 1992, 26, 808 814
  18. 18
    Vikesland, P. J.; Valentine, R. L. Reaction pathways involved in the reduction of monochloramine by ferrous iron Environ. Sci. Technol. 2000, 34, 83 90
  19. 19
    Corbin, J. F.; Teel, A. L.; Allen-King, R. M.; Watts, R. J. Reactive oxygen species responsible for the enhanced desorption of dodecane in modified Fenton’s systems Water Environ. Res. 2007, 79, 37 42
  20. 20
    Smith, B. A.; Teel, A. L.; Watts, R. J. Mechanism for the destruction of carbon tetrachloride and chloroform DNAPLs by modified Fenton’s reagent J. Contam. Hydrol. 2006, 85, 229 246
  21. 21
    Watts, R. J.; Sarasa, J.; Loge, F. J.; Teel, A. L. Oxidative and reductive pathways in manganese-catalyzed Fenton’s reactions J. Environ. Eng. 2005, 131, 158 164
  22. 22
    Smith, B. A.; Teel, A. L.; Watts, R. J. Identification of the reactive oxygen species responsible for carbon tetrachloride degradation in modified Fenton’s systems Environ. Sci. Technol. 2004, 38, 5465 5469
  23. 23
    Teel, A. L.; Watts, R. J. Degradation of carbon tetrachloride by modified Fenton’s reagent J. Hazard. Mater. 2002, 94, 179 189
  24. 24
    Watts, R. J.; Bottenberg, B. C.; Hess, T. F.; Jensen, M. D.; Teel, A. L. Role of reductants in the enhanced desorption and transformation of chloroaliphatic compounds by modified Fenton’s reactions Environ. Sci. Technol. 1999, 33, 3432 3437
  25. 25
    Valentine, R. L.; Ozekin, K.; Vikesland, P. J. Chloramine Decomposition in Distribution System and Model Waters; American Water Works Association Research Foundation: Denver, CO, 1998.
  26. 26
    Valentine, R. L.; Jafvert, C. T.; Leung, S. W. Evaluation of a chloramine decomposition model incorporating general acid catalysis Water Res. 1988, 22, 1147 1153
  27. 27
    Jafvert, C. T.; Valentine, R. L. Dichloramine decomposition in the presence of excess ammonia Water Res. 1987, 21, 967 973
  28. 28
    Valentine, R. L.; Jafvert, C. T. General acid catalysis of monochloramine disproportionation Environ. Sci. Technol. 1988, 22, 691 696
  29. 29
    Schock, M. R. Hot research topics in lead control. Presented at the EPA Inorganic Contaminant Issues Workshop, Cincinnati, Ohio, 2007;
  30. 30
    Lin, Y. P.; Washburn, M.; Valentine, R. L. Reduction of lead oxide (PbO2) by iodide and formation of iodoform in the PbO2/I/NOM system Environ. Sci. Technol. 2008, 42, 2919 2924
  31. 31
    Zhang, P. C.; Ryan, J. A. Transformation of Pb(II) from cerrusite to chloropyromorphite in the presence of hydroxyapatite under varying conditions of pH Environ. Sci. Technol. 1999, 33, 625 630
  32. 32
    Zhang, P. C.; Ryan, J. A.; Yang, J. In vitro soil Pb solubility in the presence of hydroxyapatite Environ. Sci. Technol. 1998, 32, 2763 2768
  33. 33
    Schock, M. R.; Wagner, I.; Oliphant, R. J. Corrosion and solubility of lead in drinking water. In Internal Corrosion of Water Distribution Systems; AWWA Research Foundation and American Water Works Association :Denver, CO, 1996.
  34. 34
    Vikesland, P. J.; Ozekin, K.; Valentine, R. L. Effect of natural organic matter on monochloramine decomposition: Pathway elucidation through the use of mass and redox balances Environ. Sci. Technol. 1998, 32, 1409 1416
  35. 35
    Renner, R. Chloramines again linked to lead in drinking water Environ. Sci. Technol. 2005, 39, 314A 314A

Cited By


This article is cited by 38 publications.

  1. Michael K. DeSantis, Michael R. Schock, Jennifer Tully, Christina Bennett-Stamper. Orthophosphate Interactions with Destabilized PbO2 Scales. Environmental Science & Technology 2020, 54 (22) , 14302-14311. https://doi.org/10.1021/acs.est.0c03027OpenURL UNIV OF NORTH TEXAS
  2. Yeunook Bae, Jill D. Pasteris, Daniel E. Giammar. The Ability of Phosphate To Prevent Lead Release from Pipe Scale When Switching from Free Chlorine to Monochloramine. Environmental Science & Technology 2020, 54 (2) , 879-888. https://doi.org/10.1021/acs.est.9b06019OpenURL UNIV OF NORTH TEXAS
  3. Weiyi Pan, Chao Pan, Yeunook Bae, Daniel Giammar. Role of Manganese in Accelerating the Oxidation of Pb(II) Carbonate Solids to Pb(IV) Oxide at Drinking Water Conditions. Environmental Science & Technology 2019, 53 (12) , 6699-6707. https://doi.org/10.1021/acs.est.8b07356OpenURL UNIV OF NORTH TEXAS
  4. Chun-Wei Chiang, Ding-Quan Ng, Yi-Pin Lin, and Pei-Jen Chen . Dissolved Organic Matter or Salts Change the Bioavailability Processes and Toxicity of the Nanoscale Tetravalent Lead Corrosion Product PbO2 to Medaka Fish. Environmental Science & Technology 2016, 50 (20) , 11292-11301. https://doi.org/10.1021/acs.est.6b02072OpenURL UNIV OF NORTH TEXAS
  5. Sheldon Masters, Gregory J. Welter, and Marc Edwards . Seasonal Variations in Lead Release to Potable Water. Environmental Science & Technology 2016, 50 (10) , 5269-5277. https://doi.org/10.1021/acs.est.5b05060OpenURL UNIV OF NORTH TEXAS
  6. Daoping Guo, Clare Robinson, and Jose E. Herrera . Role of Pb(II) Defects in the Mechanism of Dissolution of Plattnerite (β-PbO2) in Water under Depleting Chlorine Conditions. Environmental Science & Technology 2014, 48 (21) , 12525-12532. https://doi.org/10.1021/es502133kOpenURL UNIV OF NORTH TEXAS
  7. Yuanyuan Zhang and Yi-Pin Lin . Elevated Pb(II) Release from the Reduction of Pb(IV) Corrosion Product PbO2 Induced by Bromide-Catalyzed Monochloramine Decomposition. Environmental Science & Technology 2013, 47 (19) , 10931-10938. https://doi.org/10.1021/es402733eOpenURL UNIV OF NORTH TEXAS
  8. Ding-Quan Ng, Timothy J. Strathmann, and Yi-Pin Lin . Role of Orthophosphate As a Corrosion Inhibitor in Chloraminated Solutions Containing Tetravalent Lead Corrosion Product PbO2. Environmental Science & Technology 2012, 46 (20) , 11062-11069. https://doi.org/10.1021/es302220tOpenURL UNIV OF NORTH TEXAS
  9. Yin Wang, Jiewei Wu, and Daniel E. Giammar . Kinetics of the Reductive Dissolution of Lead(IV) Oxide by Iodide. Environmental Science & Technology 2012, 46 (11) , 5859-5866. https://doi.org/10.1021/es2038905OpenURL UNIV OF NORTH TEXAS
  10. Haizhou Liu, Andrey M. Kuznetsov, Alexey N. Masliy, John F. Ferguson, and Gregory V. Korshin . Formation of Pb(III) Intermediates in the Electrochemically Controlled Pb(II)/PbO2 System. Environmental Science & Technology 2012, 46 (3) , 1430-1438. https://doi.org/10.1021/es203084nOpenURL UNIV OF NORTH TEXAS
  11. Yuanyuan Zhang and Yi-Pin Lin . Determination of PbO2 Formation Kinetics from the Chlorination of Pb(II) Carbonate Solids via Direct PbO2 Measurement. Environmental Science & Technology 2011, 45 (6) , 2338-2344. https://doi.org/10.1021/es1039826OpenURL UNIV OF NORTH TEXAS
  12. Gregory V. Korshin. Chlorine Based Oxidants for Water Purification and Disinfection. 2011,,, 223-245. https://doi.org/10.1021/bk-2011-1071.ch011OpenURL UNIV OF NORTH TEXAS
  13. Yanjiao Xie, Yin Wang, and Daniel E. Giammar. Impact of Chlorine Disinfectants on Dissolution of the Lead Corrosion Product PbO2. Environmental Science & Technology 2010, 44 (18) , 7082-7088. https://doi.org/10.1021/es1016763OpenURL UNIV OF NORTH TEXAS
  14. Yi-Pin Lin and Richard L Valentine . Reductive Dissolution of Lead Dioxide (PbO2) in Acidic Bromide Solution. Environmental Science & Technology 2010, 44 (10) , 3895-3900. https://doi.org/10.1021/es100133nOpenURL UNIV OF NORTH TEXAS
  15. Yan Zhang, Yuanyuan Zhang and Yi-Pin Lin. Fast Detection of Lead Dioxide (PbO2) in Chlorinated Drinking Water by a Two-Stage Iodometric Method. Environmental Science & Technology 2010, 44 (4) , 1347-1352. https://doi.org/10.1021/es902299bOpenURL UNIV OF NORTH TEXAS
  16. Yanjiao Xie, Yin Wang, Vidhi Singhal and Daniel E. Giammar. Effects of pH and Carbonate Concentration on Dissolution Rates of the Lead Corrosion Product PbO2. Environmental Science & Technology 2010, 44 (3) , 1093-1099. https://doi.org/10.1021/es9026198OpenURL UNIV OF NORTH TEXAS
  17. Yi-Pin Lin and Richard L. Valentine . Reduction of Lead Oxide (PbO2) and Release of Pb(II) in Mixtures of Natural Organic Matter, Free Chlorine and Monochloramine. Environmental Science & Technology 2009, 43 (10) , 3872-3877. https://doi.org/10.1021/es900375aOpenURL UNIV OF NORTH TEXAS
  18. Shengnan Zhang, Yimei Tian, Yajing Guo, Jinlin Shan, Ran Liu. Manganese release from corrosion products of cast iron pipes in drinking water distribution systems: Effect of water temperature, pH, alkalinity, SO42− concentration and disinfectants. Chemosphere 2021, 262 , 127904. https://doi.org/10.1016/j.chemosphere.2020.127904OpenURL UNIV OF NORTH TEXAS
  19. Jun Hu, Yiran Xu, Ying Chen, Jiang Chen, Huiyu Dong, Jianming Yu, Zhimin Qiang, Jiajia Qu, Jianmeng Chen. Formation of carbonaceous and nitrogenous iodinated disinfection byproducts from biofilm extracellular polymeric substances by the oxidation of iodide-containing waters with lead dioxide. Water Research 2021, 188 , 116551. https://doi.org/10.1016/j.watres.2020.116551OpenURL UNIV OF NORTH TEXAS
  20. Reyad Roy, Arumugam Sathasivan, George Kastl. Simplified chemical chloramine decay model for water distribution systems. Science of The Total Environment 2020, 741 , 140410. https://doi.org/10.1016/j.scitotenv.2020.140410OpenURL UNIV OF NORTH TEXAS
  21. Yeunook Bae, Jill D. Pasteris, Daniel E. Giammar. Impact of iron‐rich scale in service lines on lead release to water. AWWA Water Science 2020, 2 (4) https://doi.org/10.1002/aws2.1188OpenURL UNIV OF NORTH TEXAS
  22. Fu-Chun Chang, Yi-Pin Lin. Survey of lead concentration in tap water on a university campus. Environmental Science and Pollution Research 2019, 26 (24) , 25275-25285. https://doi.org/10.1007/s11356-019-05771-1OpenURL UNIV OF NORTH TEXAS
  23. Gregory Korshin, Haizhou Liu. Preventing the colloidal dispersion of Pb( iv ) corrosion scales and lead release in drinking water distribution systems. Environmental Science: Water Research & Technology 2019, 5 (7) , 1262-1269. https://doi.org/10.1039/C9EW00231FOpenURL UNIV OF NORTH TEXAS
  24. Ding-Quan Ng, Yao Chu, Shih-Wei Tan, Shan-Li Wang, Yi-Pin Lin, Chia-Hung Chu, Yun-Liang Soo, Yen-Fang Song, Pei-Jen Chen. In vivo evidence of intestinal lead dissolution from lead dioxide (PbO 2 ) nanoparticles and resulting bioaccumulation and toxicity in medaka fish. Environmental Science: Nano 2019, 6 (2) , 580-591. https://doi.org/10.1039/C8EN00893KOpenURL UNIV OF NORTH TEXAS
  25. Ding-Quan Ng, Shu-Wei Liu, Yi-Pin Lin. Lead as a legendary pollutant with emerging concern: Survey of lead in tap water in an old campus building using four sampling methods. Science of The Total Environment 2018, 636 , 1510-1516. https://doi.org/10.1016/j.scitotenv.2018.04.402OpenURL UNIV OF NORTH TEXAS
  26. Shakhawat Chowdhury, Fayzul Kabir, Mohammad Abu Jafar Mazumder, Md. Hasan Zahir. Modeling lead concentration in drinking water of residential plumbing pipes and hot water tanks. Science of The Total Environment 2018, 635 , 35-44. https://doi.org/10.1016/j.scitotenv.2018.04.065OpenURL UNIV OF NORTH TEXAS
  27. Jun Hu, Huiyu Dong, Qiang Xu, Wencui Ling, Jiuhui Qu, Zhimin Qiang. Impacts of water quality on the corrosion of cast iron pipes for water distribution and proposed source water switch strategy. Water Research 2018, 129 , 428-435. https://doi.org/10.1016/j.watres.2017.10.065OpenURL UNIV OF NORTH TEXAS
  28. Wai Lee, Jie Jia, Yani Bao. Identifying the Gaps in Practice for Combating Lead in Drinking Water in Hong Kong. International Journal of Environmental Research and Public Health 2016, 13 (10) , 970. https://doi.org/10.3390/ijerph13100970OpenURL UNIV OF NORTH TEXAS
  29. Yuanyuan Zhang, Yi-Pin Lin. Leaching of lead from new unplasticized polyvinyl chloride (uPVC) pipes into drinking water. Environmental Science and Pollution Research 2015, 22 (11) , 8405-8411. https://doi.org/10.1007/s11356-014-3999-9OpenURL UNIV OF NORTH TEXAS
  30. Konstantinos C. Makris, Syam S. Andra, George Botsaris. Pipe Scales and Biofilms in Drinking-Water Distribution Systems: Undermining Finished Water Quality. Critical Reviews in Environmental Science and Technology 2014, 44 (13) , 1477-1523. https://doi.org/10.1080/10643389.2013.790746OpenURL UNIV OF NORTH TEXAS
  31. Syam S. Andra, Konstantinos C. Makris, George Botsaris, Pantelis Charisiadis, Harris Kalyvas, Costas N. Costa. Evidence of arsenic release promoted by disinfection by-products within drinking-water distribution systems. Science of The Total Environment 2014, 472 , 1145-1151. https://doi.org/10.1016/j.scitotenv.2013.11.045OpenURL UNIV OF NORTH TEXAS
  32. Meghan Woszczynski, John Bergese, Graham A. Gagnon. Comparison of Chlorine and Chloramines on Lead Release from Copper Pipe Rigs. Journal of Environmental Engineering 2013, 139 (8) , 1099-1107. https://doi.org/10.1061/(ASCE)EE.1943-7870.0000712OpenURL UNIV OF NORTH TEXAS
  33. Yuanyuan Zhang, Ding-Quan Ng, Yi-Pin Lin. Iodide-assisted total lead measurement and determination of different lead fractions in drinking water samples. Journal of Environmental Monitoring 2012, 14 (7) , 1846. https://doi.org/10.1039/c2em10962jOpenURL UNIV OF NORTH TEXAS
  34. Yuanyuan Zhang, Yi-Pin Lin. Adsorption of Free Chlorine on Tetravalent Lead Corrosion Product (PbO 2 ). Environmental Engineering Science 2012, 29 (1) , 52-58. https://doi.org/10.1089/ees.2010.0372OpenURL UNIV OF NORTH TEXAS
  35. Yanjiao Xie, Daniel E. Giammar. Effects of flow and water chemistry on lead release rates from pipe scales. Water Research 2011, 45 (19) , 6525-6534. https://doi.org/10.1016/j.watres.2011.09.050OpenURL UNIV OF NORTH TEXAS
  36. Simoni Triantafyllidou, Marc Edwards. Galvanic corrosion after simulated small-scale partial lead service line replacements. Journal - American Water Works Association 2011, 103 (9) , 85-99. https://doi.org/10.1002/j.1551-8833.2011.tb11535.xOpenURL UNIV OF NORTH TEXAS
  37. Andrzej Wilczak, David R. Hokanson, R. Rhodes Trussell, Manouchehr Boozarpour, Andrew F. Degraca. Water conditioning for LCR compliance and control of metals release in San Francisco's water system. Journal - American Water Works Association 2010, 102 (3) , 52-64. https://doi.org/10.1002/j.1551-8833.2010.tb10072.xOpenURL UNIV OF NORTH TEXAS
  38. RuiPing Liu, JiuHui Qu. Control of health risks in drinking water through point-of-use systems. Science Bulletin 2009, 54 (12) , 1996-2001. https://doi.org/10.1007/s11434-009-0150-2OpenURL UNIV OF NORTH TEXAS
  • Figure 1

    Figure 1. Effect of NH2Cl concentration on the formation of total Pb(II) and NH2Cl decay: (a) NH2Cl decay as a function of time, (b) total Pb(II) formation as a functiom of time, and (c) total Pb(II) formation vs NH2Cl decay (data of duplicate experiments were shown by separate points). Fixed experimental parameters: PbO2 = 10 mg/L, pH 7.0, CT = 5 mM, Cl/N = 0.7. In control experiment for NH2Cl decay, no PbO2 was added.

    Figure 2

    Figure 2. SEM image of the solid phase after 7 day experiment. NH2Cl = 70.4 μM, PbO2 = 10 mg/L, pH 7.0, CT = 5 mM, Cl/N = 0.7.

    Figure 3

    Figure 3. Effect of solution pH on the formation of total Pb(II) and NH2Cl decay: (a) NH2Cl decay as a function of time, (b) total Pb(II) formation as a function of time, and (c) total Pb(II) formation vs NH2Cl decay vs total Pb(II) formation (data of duplicate experiments were shown by separate points). PbO2 = 10 mg/L, NH2Cl = 42.3 μM, CT = 5 mM, Cl/N = 0.7. In control experiment for NH2Cl decay, no PbO2 was added.

    Figure 4

    Figure 4. Effect of total carbonate concentration on the formation of total Pb(II) and NH2Cl decay: (a) total Pb(II) formation as a functiom of time, (b) NH2Cl decay as a function of time, and (c) NH2Cl decay vs total Pb(II) formation (data of duplicate experiments were shown by separate points). PbO2 = 10 mg/L, NH2Cl = 42.3 μM, pH 7.0, Cl/N = 0.7. In control experiment for NH2Cl decay, no PbO2 was added.

    Figure 5

    Figure 5. Effect of Cl/N molar ratio on the formation of total Pb(II) and NH2Cl decay: (a) total Pb(II) formation as a functiom of time, (b) NH2Cl decay as a function of time, and (c) NH2Cl decay vs total Pb(II) formation (data of duplicate experiments were shown by separate points). PbO2 = 10 mg/L, NH2Cl = 42.3 μM, pH 7.0, CT = 5 mM. In control experiment for NH2Cl decay, no PbO2 was added.

    Figure 6

    Figure 6. Summary plot of total Pb(II) formation vs NH2Cl decay.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 35 other publications.

    1. 1

      American Water Works Association. Water Quality and Treatment: A Handbook of Community Water Supplies; 5th edition; McGraw-Hill, Inc.: New York,

      1999.
    2. 2
      Renner, R. Plumbing the depths of DCʼs drinking water crisis Environ. Sci. Technol. 2004, 38, 224A 227A
    3. 3
      Edwards, M.; Dudi, A. Role of chlorine and chloramine in corrosion of lead-bearing plumbing materials J. Am. Water Works Assoc. 2004, 96, 69 81
    4. 4
      . Maximum Contamination Level Goals and National Primary Drinking Water Regulations for Lead and Copper. Final Rule. Fed. Regist. 1991, 56, 2646026564.
    5. 5
      Lytle, D. A.; Schock, M. R. Formation of Pb(IV) oxides in chlorinated water J. Am. Water Works Assoc. 2005, 97, 102 114
    6. 6
      Schock, M. R.. New Insights into lead corrosion control and treatment change impacts (with some considerations towards Cu). AWWA Meeting Section Emerging Issues in Water Treatment, Michigan, 2007;
    7. 7
      Dryer, D. J.; Korshin, G. V. Investigation of the reduction of lead dioxide by natural organic matter Environ. Sci. Technol. 2007, 41, 5510 5514
    8. 8
      Schock, M. R.; Scheckel, K. G.; DeSantis, M.; Gerke, T. L. Mode of occurrence, treatment and monitoring significance of tetravalent lead. Presented at the AWWA Water Quality Technology Conference, Quebec, Canada, 2005.
    9. 9
      Schock, M. R.; Harmon, S. M.; Swertfeger, J.; Lohmann, R. Tetravalent lead: A hitherto unrecognized control of tap water lead contamination. Presented at the AWWA Water Quality Technology Conference, Nashville, TN, 2001.
    10. 10
      Rajasekharan, V. V.; Clark, B. N.; Boonsalee, S.; Switzer, J. A. Electrochemistry of free chlorine and monochloramine and its relevance to the presence of Pb in drinking water Environ. Sci. Technol. 2007, 41, 4252 4257
    11. 11
      Lin, Y. P.; Valentine, R. L. The release of lead from the reduction of lead oxide (PbO2) by natural organic matter Environ. Sci. Technol. 2008, 42, 760 765
    12. 12
      Jafvert, C. T.; Valentine, R. L. Reaction scheme for the chlorination of ammoniacal water Environ. Sci. Technol. 1992, 26, 577 586
    13. 13
      Ozekin, K.; Valentine, R. L.; Vikesland, P. J. Modeling the decomposition of disinfecting residuals of chloramine. InWater Disinfection and Natural Organic Matter; ACS Symposium Series 649; American Chemical Society: Washington, DC, 1996; pp 115125.
    14. 14
      Vikesland, P. J.; Ozekin, K.; Valentine, R. L. Monochloramine decay in model and distribution system waters Water Res. 2001, 35, 1766 1776
    15. 15
      Duirk, S. E.; Gombert, B.; Croue, J. P.; Valentine, R. L. Modeling monochloramine loss in the presence of natural organic matter Water Res. 2005, 39, 3418 3431
    16. 16
      Valentine, R. L.; Brandt, K. I.; Jafvert, C. T. A spectrophotometric study of the formation of an unidentified monochloramine decomposition product Water Res. 1986, 20, 1067 1074
    17. 17
      Diyamandoglu, V.; Selleck, R. E. Reactions and products of chloramination Environ. Sci. Technol. 1992, 26, 808 814
    18. 18
      Vikesland, P. J.; Valentine, R. L. Reaction pathways involved in the reduction of monochloramine by ferrous iron Environ. Sci. Technol. 2000, 34, 83 90
    19. 19
      Corbin, J. F.; Teel, A. L.; Allen-King, R. M.; Watts, R. J. Reactive oxygen species responsible for the enhanced desorption of dodecane in modified Fenton’s systems Water Environ. Res. 2007, 79, 37 42
    20. 20
      Smith, B. A.; Teel, A. L.; Watts, R. J. Mechanism for the destruction of carbon tetrachloride and chloroform DNAPLs by modified Fenton’s reagent J. Contam. Hydrol. 2006, 85, 229 246
    21. 21
      Watts, R. J.; Sarasa, J.; Loge, F. J.; Teel, A. L. Oxidative and reductive pathways in manganese-catalyzed Fenton’s reactions J. Environ. Eng. 2005, 131, 158 164
    22. 22
      Smith, B. A.; Teel, A. L.; Watts, R. J. Identification of the reactive oxygen species responsible for carbon tetrachloride degradation in modified Fenton’s systems Environ. Sci. Technol. 2004, 38, 5465 5469
    23. 23
      Teel, A. L.; Watts, R. J. Degradation of carbon tetrachloride by modified Fenton’s reagent J. Hazard. Mater. 2002, 94, 179 189
    24. 24
      Watts, R. J.; Bottenberg, B. C.; Hess, T. F.; Jensen, M. D.; Teel, A. L. Role of reductants in the enhanced desorption and transformation of chloroaliphatic compounds by modified Fenton’s reactions Environ. Sci. Technol. 1999, 33, 3432 3437
    25. 25
      Valentine, R. L.; Ozekin, K.; Vikesland, P. J. Chloramine Decomposition in Distribution System and Model Waters; American Water Works Association Research Foundation: Denver, CO, 1998.
    26. 26
      Valentine, R. L.; Jafvert, C. T.; Leung, S. W. Evaluation of a chloramine decomposition model incorporating general acid catalysis Water Res. 1988, 22, 1147 1153
    27. 27
      Jafvert, C. T.; Valentine, R. L. Dichloramine decomposition in the presence of excess ammonia Water Res. 1987, 21, 967 973
    28. 28
      Valentine, R. L.; Jafvert, C. T. General acid catalysis of monochloramine disproportionation Environ. Sci. Technol. 1988, 22, 691 696
    29. 29
      Schock, M. R. Hot research topics in lead control. Presented at the EPA Inorganic Contaminant Issues Workshop, Cincinnati, Ohio, 2007;
    30. 30
      Lin, Y. P.; Washburn, M.; Valentine, R. L. Reduction of lead oxide (PbO2) by iodide and formation of iodoform in the PbO2/I/NOM system Environ. Sci. Technol. 2008, 42, 2919 2924
    31. 31
      Zhang, P. C.; Ryan, J. A. Transformation of Pb(II) from cerrusite to chloropyromorphite in the presence of hydroxyapatite under varying conditions of pH Environ. Sci. Technol. 1999, 33, 625 630
    32. 32
      Zhang, P. C.; Ryan, J. A.; Yang, J. In vitro soil Pb solubility in the presence of hydroxyapatite Environ. Sci. Technol. 1998, 32, 2763 2768
    33. 33
      Schock, M. R.; Wagner, I.; Oliphant, R. J. Corrosion and solubility of lead in drinking water. In Internal Corrosion of Water Distribution Systems; AWWA Research Foundation and American Water Works Association :Denver, CO, 1996.
    34. 34
      Vikesland, P. J.; Ozekin, K.; Valentine, R. L. Effect of natural organic matter on monochloramine decomposition: Pathway elucidation through the use of mass and redox balances Environ. Sci. Technol. 1998, 32, 1409 1416
    35. 35
      Renner, R. Chloramines again linked to lead in drinking water Environ. Sci. Technol. 2005, 39, 314A 314A
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    Additional table of NH2Cl autodecomposition reactions and figures of PbO2 SEM images, and second-order plots of monochloramine decay. This material is available free of charge via the Internet at http://pubs.acs.org.


    Terms & Conditions

    Electronic Supporting Information files are available without a subscription to ACS Web Editions. The American Chemical Society holds a copyright ownership interest in any copyrightable Supporting Information. Files available from the ACS website may be downloaded for personal use only. Users are not otherwise permitted to reproduce, republish, redistribute, or sell any Supporting Information from the ACS website, either in whole or in part, in either machine-readable form or any other form without permission from the American Chemical Society. For permission to reproduce, republish and redistribute this material, requesters must process their own requests via the RightsLink permission system. Information about how to use the RightsLink permission system can be found at http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

OOPS

You have to login with your ACS ID befor you can login with your Mendeley account.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect

This website uses cookies to improve your user experience. By continuing to use the site, you are accepting our use of cookies. Read the ACS privacy policy.

CONTINUE