You are on page 1of 234

American Water Works Association

.RESEARCH FOUNDATION

Chemistry of
Corrosion Inhibitors
in Potable Water

Subject Area:
Distribution Systems
CHEMISTRY OF CORROSION
INHIBITORS IN POTABLE WATER

Prepared by:
Mark M. Benjamin
Steve H. Reiber
John F. Ferguson
Elizabeth A. Vanderwerff
and
Mark W. Miller
Department of Civil Engineering
University of Washington
Seattle, WA 98195

Prepared for:
AWWA Research Foundation
6666 West Quincy Avenue
Denver, CO 80235

February 1990

Published by the AWWA Research Foundation


and the American Water Works Association
DISCLAIMER

This study was funded by the American Water Works Association Research
Foundation (AWWARF). AWWARF assumes no responsibility for the content of the
research study reported in this publication, or for the opinions or statements
of fact expressed i.n the report. The mention of tradenames for commercial
products does not represent or imply the approval or endorsement of AWWARF.
This report is presented solely for informational purposes.

Copyright 1990
by
AWWA Research Foundation
American Water Works Association
Printed in U.S.A.

ISBN 0-89867-506-5
TABLE OF CONTENTS

Page
TABLES...................................................... vi i
FIGURES..................................................... ix
ACKNOWLEDGMENTS............................................. xli i
FOREWORD.................................................... xv
EXECUTIVE SUMMARY........................................... xvii
I. INTRODUCTION................................................ 1
II. LITERATURE REVIEW........................................... 5
Introduction............................................. 5
Action of Corrosion Inhibitors........................... 6
Chemical Inhibitors in Potable Water..................... 7
Phosphate, Polyphosphates, and Zinc Phosphate......... 7
Silicate Inhibitors................................... 9
Combinations of Inhibitors............................ 10
Carbonate-Based Inhibitors: CaCO^ Deposition Indices.. 10
Scale Composition and the Siderite Model of Scale
Format ion.............................................. 12
Summary.................................................. 15
III. METHODS AND MATERIALS....................................... 17
Introduction............................................. 17
Water Quali ty i n the Exposure Loops...................... 17
Seattle Tap Water..................................... 17
Water Quali ty Adjustment.............................. 18
Loop-Water Monitoring................................. 22
Corrosion-Rate Measurement............................... 23
Introduction.......................................... 23
Long-Term Corrosion Rates............................. 24
Pipe-Loop Construction and Coupon Materials........... 27
Surface Preparation for Black- and Galvanized-Iron
Coupons............................................. 33
Surface Preparation for Copper Coupons................ 37
Technique Validation.................................. 40
Electrochemical Corrosion-Rate Measures.................. 40
Introduction.......................................... 40
Polarization Cell and Electric Circuit................ 42
General Polarization Measurements..................... 44
Corrosion-Rate Measurements by Monitoring Oxygen
Depletion.............................................. 50
Purpose of Oxygen-Depletion Corrosion-Rate
Measurements........................................ 50
Page
Oxygen-Depletion Piping Loop.......................... 51
Analysis of Oxygen-Depletion Corrosion-Rate
Measurements........................................ 53
Accuracy and Precision of DO Depletion
Measurements........................................ 55
Analysis of Corrosion Scale.............................. 57
Purpose of Corrosion-Scale Analysis................... 57
X-Ray Analysis of Corrosion Scales.................... 58
Chemical Analysis of Corrosion Scales................. 59
Scale Oxidation.................................... 59
Ferrous Iron Analys i s.............................. 59
Total-Metal Analysis of Dried Scales............... 61
Carbonate Analysis................................. 61
Phosphate.......................................... 62
Copper Composition of Scales....................... 62
IV. CORROSION OF MILD-STEEL PIPE................................ 65
Corrosion Rates Based on Weight Loss of Mild-Steel Pipe.. 65
Overall Weight-Loss Comparison........................ 65
Correlation of Water Quality Variable with Corrosion
Rates............................................... 76
Buffer Intensity................................... 77
Alkalinity......................................... 78
The Langlier Saturation Index...................... 79
pH................................................. 79
Phosphate.......................................... 80
Initial Corrosion Rates............................... 81
Comparison of Long-Term Corrosion Rates in pH 6.0
and pH 8.0 Seattle Tap Water........................ 82
Comparison with Previous Corrosion Research
Conducted Using Seattle Tap Water................... 83
Iron Corrosion Scale Analysis............................ 86
Physical Characteristics of Iron Corrosion Scales..... 86
Corrosion-Scale Weights............................... 88
X-Ray Diffraction Analysis of Iron Corrosion Scale.... 88
Chemical Analysis of Iron Corrosion Scale............. 91
Total Iron Content................................. 91
Ferrous Iron....................................... 92
Carbonate Content.................................. 95
The Siderite Model....................................... 97
Qualitative Correlation of This Research With the
Siderite Model of Iron Corrosion.................... 97
Relationship Between Water Quality Parameters
and the Siderite Model........................... 98
A Quantitative Iron Corrosion Model Based on
Siderite Formation.................................. 104
Response of Scale-Covered Mild-Steel Pipe to Changes
in the Alkalinity and pH of the Conditioning Water..... 107
Corrosion-Rate Measurements........................... 107
Scale Analysis........................................ 113

IV
The Effect of High Conductivity on Initial
Corrosion Rates..................................... 115
Summary.................................................. 117
V. CORROSION OF COPPER PIPES................................... 121
Copper Corrosion Rates................................... 121
Scale Accumulation and Composition....................... 128
Short-Term Water Quality Effects on Aged Copper
Corrosion.............................................. 134
Di ssolved Oxygen...................................... 136
Chlorine Residual..................................... 136
pH ................................................... 140
Short-Term Tests Using Fresh Coupons..................... 142
Summary.................................................. 149
VI. CORROSION OF GALVANIZED PIPES............................... 151
Cumulative Weight Loss................................... 151
Scale Characteristics.................................... 155
Scale Accumulation.................................... 155
Scale Mineralogical Identification.................... 159
Electrochemical Measurements on Galvanized Surfaces......... 161
Validation............................................... 161
Reduction Couples........................................ 164
Di ssolved Oxygen...................................... 164
Chlori ne.............................................. 167
pH ................................................... 167
Summary.................................................. 170
VII. SUMMARY, CONCLUSIONS, AND RECOMMENDATIONS................... 173
Mild Steel............................................... 173
Copper................................................... 175
Galvanized Steel......................................... 176
Recommendations.......................................... 177
VIII. REFERENCES.................................................. 179
I.. APPENDIX Al................................................. 185
Validation of Weight-Loss Corrosion-Rate
Measurement Techniques................................. 185
Summary of the Weight-Loss Technique..................... 191
.. APPENDIX A2................................................. 193
Electrochemical Corrosion-Rate Measurements:
Principles and Validation.............................. 193
Surface-Polarization Procedures.......................... 194
The Leroy Model.......................................... 199
Accuracy of Corrosion Measurements....................... 201
Surface Potential and Pitting............................ 207
Pipe Surface Electrochemistry............................ 208
.I. NOMENCLATURE................................................ 209
LIST OF TABLES

Table Page
1 Seattle Tap Water as Received at the University of
Washington During 1986 and 1987......................... 18
2 Characteristics of Conditioning Loop Water................ 19
3 Summary of Procedures for Black-Iron and Galvanized-
Iron Coupon Preparation/Restoration..................... 36
4 Summary of Copper Coupon Preparation/Restoration
Technique............................................... 39
5 Comparison of Corrosion-Rate Measurement Methods.......... 56
6 Cumulative Corrosion Rates Over Six Months Exposure....... 76
7 Water Quality Variables and Corrosion Rates............... 77
8 Initial Corrosion Rates Found on Mild-Steel Pipe.......... 81
9 Long-Term Corrosion Rates in pH 6.0 and pH 8.0 Seattle
Tap Water............................................... 83
10 Seattle Corrosion-Control Pilot-Plant Studies Average
Corrosion Rates Obtained on Mild-Steel Pipe............. 85
11 Iron Content of Dried Iron Corrosion Scales............... 91
12 Ferrous Iron Content of Iron Corrosion Scales............. 93
13 Carbonate Content of Iron Scales.......................... 96
14 Weight-Loss Data From Changing Water Chemistry
Experiments............................................. 112
15 Ferrous Iron and Carbonate Content of Iron Scales
Exposed to Changes in Water Chemistry................... 114
16 Response of Corrosion Rate to High-Conductivity Water..... 117
17 Composition of Copper Corrosion Scales.................... 133
18 Zinc Concentration in Zinc Galvanized Corrosion Scales.... 158
Al-1 Weight Loss Variation as a Function of Exposure
Duration................................................ 185
Al-2 Weight Loss Variation as a Function of Exposure
Duration Corrected for Variation due to
Surface-Preparation Techniques.......................... 190
A2-1 Comparison of Electrochemical Corrosion Rates with
Weight-Loss Measurements................................ 206

Vll
LIST OF FIGURES

Figure Page
1 Cross Section of Insert, Spacer, and Union, and
Assembled ISWS Corrosion Tester......................... 25
2 Cross Section of Coupon/Sleeve Headpiece Assembly
Used for Weight-Loss Test............................... 28
3 Black-Iron, Galvanized-Iron, and Copper Coupon/Inserts.... 30
4 Pi pe Coupon Exposure Loop................................. 32
5 Black-Iron Coupon With Polishing Tool..................... 34
6 Copper Coupon With Polishing Tool......................... 38
7 Polarization Flow Cell.................................... 43
8 Schematic of Closed-Loop Recirculation System for
Polarization Measurements............................... 45
9 Flow Diagram of Tafel Slope Fitting Routine............... 48
10 Evans Diagram of Polarization on a Copper Surface
Showing Theoretical and Observed Values................. 49
11 Schematic of Pipe Loop for Measuring Corrosion Rate by
Oxygen Depletion........................................ 52
12 Weight Loss of Iron Pipes Exposed to Seattle
Tap Water, pH 8.0 ...................................... 66
13 Weight Loss of Mild-Steel Pipe Sections Exposed to
pH 6.0 Water............................................ 67
14 Weight Loss of Iron Pipes Exposed to the Low-
Carbonate Water at pH 8.0............................... 68
15 Weight Loss of Iron Pipe Exposed to the Low-
Carbonate Water at pH 8.4.............................. 69
16 Weight Loss of Iron Pipe Exposed to the pH 7.2,
High-Carbonate Water................................... 70
17 Weight Loss of Mild-Steel Coupons in P04 -Inhibited
and Noninhibited Waters................................. 71
18 Change in Corrosion Rate With Length of Exposure to
pH 8.0 Water............................................ 72
19 Change in Corrosion Rate With Length of Exposure to
pH 6.0 Water ........................................... 73
20 Change in Corrosion Rate With Length of Exposure to
Low-Carbonate Waters.................................... 74
21 Change in Corrosion Rate With Length of Exposure to
High-Carbonate Water.................................... 75
22 Scale Weight Formed on Iron Pipe Exposed to Different
Water Qualities......................................... 89
23 Total Soluble Ferrous Iron in Equilibrium With Siderite... 99
24 Oxidation Rate of Ferrous Iron in Equilibrium With
Siderite................................................ 101
25 Calculated Buffer Intensity as a Function of pH for
Waters with the Indicated Concentration of Inorganic
Carbon and No Other Weak Acids or Bases................. 103
26 Corrosion Rate of High-Carbonate Scale Exposed to
pH 8.0 Tap Water........................................ 109
27 Corrosion Rate of High-Carbonate Scale Exposed to
pH 6.0 Tap Water........................................ 110

ix
jgure Page

28 Seattle Tap Water Scale Exposed to High-Carbonate Water... Ill


29 Weight Loss of Copper Pipe Exposed to pH 8.0 Seattle
Tap Water............................................... 122
30 Weight Loss of Copper Pipe Exposed to Low-
Carbonate Seattle Tap Water at pH 8.0................... 123
31 Weight Loss of Copper Pipe Exposed to High-
Carbonate Seattle Tap Water at pH 7.2 .................. 124
32 Weight Loss of Copper Pipe Exposed to Seattle Tap Water
Adjusted to pH 6.0...................................... 125
33 Weight Loss of Copper Pipe Exposed to Various Water
Qualities in the Absence of Phosphate................... 126
34 Variation in Corrosion Current and Tafel Slopes on an
Aging Copper Surface Exposed to pH 6.0 Tap Water........ 127
35 Weight-Loss Comparison of Copper Pipes in the Presence
and Absence of Phosphate................................ 129
36 Scale Weight on Phosphate-Inhibited and Noninhibited
Copper Pipe............................................. 130
37 Scale Weight on Copper Pipes Exposed to Different Water
Qualities............................................... 131
38 Comparison of Instantaneous Corrosion Rates on Copper
Coupon Surfaces Under Various Water Quality and
Inhibitor Conditions.................................... 135
39 Variation in Corrosion Current and Tafel Slopes
as a Function of Dissolved Oxygen....................... 137
40 Variation in Corrosion Current and Tafel Slopes on
Copper Surfaces as a Function of Free Chlorine
Residual................................................ 138
41 Chlorine Residual Effects on Well-Aged Phosphate-Based
Seales.................................................. 139
42 Variation in Corrosion Current and Tafel Slopes on
Copper Surfaces as a Function of pH..................... 141
43 Response of Copper Coupon to pH Transients PO^-P = 0 ppm.. 143
44 Corrosion Current for Coupon Exposed to 1 mg/L P04
at Indicated Time....................................... 144
45 Effect of Orthophosphate Inhibitor Addition on
Corrosion Current and Tafel Slopes for a Relatively
Fresh Copper Surface.................................... 145
46 Response of Copper Coupon to pH Transients;
P04 -P = 5 mg/L.......................................... 147
47 Effects of pH Shift on Phosphate Inhibitor Scale.......... 148
48 Weight Loss for Galvanized Pipe Exposed to Various
Water Qualities in the Absence of Phosphate............. 152
49 Weight Loss for Galvanized Pipe Exposed to Different
Water Qualities in the Absence of Phosphate............. 153
50 Weight Loss for Galvanized Pipe in the Presence and
Absence of Phosphate.................................... 154
51 Scale Weight Formed on Galvanized-Iron Pipe Exposed to
Phosphate-Inhibited and Noninhibited Waters............. 156
52 Scale Weight Formed on Galvanized Pipe Exposed to
Various Water Qualities in the Absence of Phosphate..... 157
Page

53 Scale Weight Formed on Galvanized Pipe in the Presence


of Phosphate............................................ 160
54 Electrochemical Measurement Variation as a Function
of Exposure Time........................................ 162
55 Comparison of Weight Loss and Electrochemical Corrosion-
Rate Measurements Under Different Water Quality
Conditions.............................................. 163
56 Electrochemical Measurement Variation on a Galvanized
Surface as a Function of Dissolved Oxygen............... 165
57 Evans Diagram for Polarization of a Galvanized Coupon
Surface................................................. 166
58 Electrochemical Measurement Variation on a Galvanized
Surface as a Function of Free Chlorine Residual......... 168
59 Electrochemical Measurement Variation on a Galvanized
Surface as a Function of pH............................. 169
Al-1 Coefficient of Dispersion as a Function of Exposure
Duration................................................ 187
Al-2 Modified Coefficient of Dispersion as a Function of
Exposure Duration....................................... 189
A2-1 Surface Potential Scan Across a One Centimeter Band
of a Black Electrode.................................... 196
A2-2 Surface Potential Scan Across a One Centimeter Band
of a Copper Electrode................................... 197
A2-3 Surface Potential Scan Across a One Centimeter Band
of a Galvanized Iron Electrode ......................... 198
A2-4 Evans Diagram of Measured and Predicted Polarization
Currents on a Black Iron Electrode ..................... 202
A2-5 Evans Diagram of Measured and Predicted Polarization
Currents on a Black Iron Electrode ..................... 203
A2-6 Evans Diagram of Measured and Predicted Polarization
Currents on a Copper Electrode.......................... 204
A2-7 Evans Diagram of Measured and Predicted Polarization
Currents on a Copper Electrode.......................... 205

XI
ACKNOWLEDGMENTS

The authors wish to express their appreciation for the financial and technical
assistance provided by the AWWA Research Foundation in general, and the
support and patience of Nancy McTigue and Rick Karlin in particular. We have
also benefited significantly from discussions with members of the project
advisory committee: Man/in Gardels of the USEPA, Steven Hubbs of the
Louisville (Ky.) Water Company, and Greg Reed of the University of Tennessee.
Finally, this document could not have been completed without the diligent
efforts and unfailing good humor of Marcia Buck and Pam Hjelm, word processors
par excellence.

Xlll
FOREWORD

This report is part of the on-going research program of the AWWA Research
Foundation. The research described in the following pages was funded by the
Foundation in behalf of its members and subscribers in particular and the
water supply industry in general. Selected for funding by AWWARF's Board of
Trustees, the project was identified as a practical, priority need of the
industry. It is hoped that this publication will receive wide and serious
attention and that its findings, conclusions, and recommendations will be
applied in communities throughout the United States and Canada.
The Research Foundation was created by the water supply industry as its center
for cooperative research and development. The Foundation itself does not
conduct research; it functions as a planning and management agency, awarding
contracts to other institutions, such as water utilities, universities,
engineering firms, and other organizations. The scientific and technical
expertise of the staff is further enhanced by industry volunteers who serve on
Project Advisory Committees and on other standing committees and councils. An
extensive planning process involves marty hundreds of water professionals in
the important task of keeping the Foundation's program responsive to the
practical, operational needs of local utilities and to the general research
and development needs of a progressive industry.
All aspects of water supply are served by AWWARF's research agenda:
resources, treatment and operations, distribution and storage, water quality
and analysis, economics and management. The ultimate purpose of this effort
is to assist local water suppliers to provide the highest possible quality of
water, economically and reliably. The Foundation's Trustees are pleased to
offer this publication as contribution toward that end.
The goal of this project was to improve the scientific and engineering
understanding of the corrosion process in household plumbing, and thereby
develop improved methods to reduce corrosion.

Richard P. McHugh // AJames F. Manwaring, P.E.


Chairman, Board of Trustees Hx^cutive Director
AWWA Research Foundation AWWA Research Foundation

XV
EXECUTIVE SUMMARY

INTRODUCTION
This research project investigated the corrosion rates of copper, mild-steel,
and zinc-galvanized pipe coupons as a function of the chemical
characteristics of the water they were exposed to. The coupons were short
sections of pipe typical of the kind that might be employed in a household
plumbing system. Several coupons of each material were placed in a holder
and exposed to water of controlled composition over a period ranging from a
few weeks to more than two years. After a predetermined exposure period, the
coupons were removed from the assembly and analyzed for weight loss.
Additionally, corrosion scales that had developed were characterized by
visual examination, by x-ray diffraction to identify their mineral forms, and
by chemical analysis.
The weight-loss analysis gave an indication of the corrosion rate over
periods of months to years. Other techniques were used to evaluate short-
term corrosion rates, particularly the rates that reflected short-term
responses to transient changes in the composition of the circulating water.
These other techniques included the rate of depletion of dissolved oxygen in
the water and electrochemical polarization methods.
Overall, more than 1,000 pipe coupons were subjected to corrosion testing.
The number of individual coupons analyzed, the range of analyses, the variety
of water qualities investigated, and the duration of the tests make this one
of the most extensive studies of corrosion in potable water systems ever
undertaken.

WATER QUALITY VARIABLES STUDIED


The water used in this project all originated as Seattle tap water. This
water contains very little natural alkalinity, hardness, or other dissolved
constituents, and it is highly undersaturated with respect to precipitation
of calcium carbonate. After treatment, its pH is near 8.0.
Six different waters, all prepared from Seattle tap water, were studied. One
of the waters studied was simple tap water, with minor additions of acid or
base as necessary to maintain a pH of 8.0. A second water was prepared by
acidifying tap water to pH 6.0 with hydrochloric acid.
Two waters were prepared by adjusting the carbonate content of tap water
upward and downward from its natural value. In one, tap water was acidified
and bubbled with Ct^-free gas to remove about 30 percent of the total
dissolved inorganic carbon before its pH was re-adjusted to 8.0. In another,
sodium bicarbonate was added to the solution, and (X^-enriched air was
bubbled through tap water. This latter treatment generated a water with
pH 7.2 and a much higher alkalinity (400 mg/L as CaC03 ) than any of the other
waters studied.

XVll
The final two waters were prepared by adding either 1 or 5 mg/L PO*-P to tap
water and adjusting the pH to 8.0. Overall, the six waters studied represent
a range of possible water quality alterations that a utility might consider
to avoid or mitigate a corrosion problem.

CORROSION OF MILD STEEL


The three pipe materials responded quite differently to the six waters. Of
the metals studied, mild steel (black iron) pipe sections corroded most
rapidly. For these coupons, the pH 7.2 high-carbonate water was dramatically
less corrosive than any of the other waters studied. Corrosion seemed to
correlate reasonably well with buffer intensity, less well with alkalinity,
and not well at all with other water quality measures such as pH or Langelier
Saturation Index. In particular, increasing the pH of the exposure water
toward 8.3 without changing the total dissolved inorganic carbon content
decreased the buffer intensity and increased the corrosion rate, even though
it increased the Langelier Index.
In accord with the model of Sontheimer et al. (1981), good corrosion
protection in iron pipes seems to be conferred by the presence, at least
transiently, of siderite (FeC03) in the protective scale. There was direct
or indirect evidence for the presence of this mineral in every protective
scale. Siderite participates in the formation of a hard, adherent scale on
the pipe surface that is very distinctive. Because siderite is formed by the
reaction of ferrous and carbonate ions, and because ferrous ions are unstable
with respect to oxidation to ferric ions in oxygenated solutions, the
development of a siderite layer requires conditions that limit the oxidation
rate of ferrous ion and allow it to build up in the surface environment.
Moderately alkaline, high-carbonate, highly buffered waters meet these
requirements. Even a relatively low content of siderite in the scale seems
to confer a good deal of corrosion protection.
A moderate maintenance dose of phosphate (1 mg/L P) had no significant effect
on corrosion of mild steel pipe sections. While 5 mg/L P04 -P did reduce
their corrosion rates, the absence of an effect at the lower inhibitor dose
makes control of iron corrosion by phosphate addition unattractive, at least
in the low alkalinity waters studied.

CORROSION OF COPPER
As expected, copper pipe sections had much lower corrosion rates than steel
pipe sections. Copper differed from steel in that even low doses of
phosphate inhibitor dramatically lowered the corrosion rate of copper
coupons. Somewhat surprisingly, changing pH from 6.0 to 8.0 did not reduce
the corrosion rate of copper, in contrast with results of previous
investigators.
Copper corrosion rates were fairly similar in all the phosphate-free loops,
especially after the first two months of exposure. Rates were steady
thereafter and were not responsive to carbonate content or pH. The mineral
cuprite was found in the scale from the pH 8.0 tap water exposure, but not in
xviii
the scale from lower-pH loops. It is possible that although the cuprite did
not form a continuous, protective layer during the course of this study, such
a layer might have formed after longer exposure periods.
The chemical composition of the scale in the high-phosphate loop was
consistent with formation of a mineral of the composition Cu^PO^)?.
However, the scale in the low-phosphate loop contained much less phosphate
than would be required to bond with all the oxidized Cu in the scale. Thus,
just as a small amount of siderite can offer corrosion protection to iron
pipe, it appears that copper pipe can be protected by the presence of a
relatively small fraction of the protective mineral in the scale. The
presence of such a scale also protected the pipe against the short-term
effects of exposure to a more corrosive water.

CORROSION OF GALVANIZED STEEL


Corrosion of zinc galvanized pipes seemed to proceed in two stages. The
first stage appears to end when most, if not all, of the zinc has been
oxidized. During the first stage, zinc corrosion rate responds most strongly
to pH. The pH 6.0 water corroded the coupons at a rate almost an order of
magnitude faster than the other water qualities. The pH 7.2 high-carbonate
loop was comparably corrosive to the pH 8.0 loops, suggesting that either
there is a critical pH somewhere between 6.0 and 7.2 at which the corrosion
rate changes dramatically, or else that the high-carbonate concentration
counteracted the corrosiveness of the pH 7.2 water somewhat.
The weight of accumulated scale was as large or larger in the pH 6.0 loop
compared to the other loops, and a zinc phosphate mineral, hopeite
(Zn3(P04 )2), was detected in the scale from the high-phosphate loop.
Together, these results confirm an obvious but sometimes overlooked fact:
that neither accumulation of large amounts of scale nor the presence of a
well-crystalized mineral can guarantee good corrosion protection.
Corrosion rates of aged galvanized coupons were unresponsive to changes in
solution residual chlorine or dissolved oxygen content, and only weakly
responsive to changes in pH. This contrasts with the results for aged copper
coupons, for which corrosion rates increased dramatically as the chlorine
residual increased.

RECOMMENDATIONS
Among the significant conclusions from this research is that no single action
can be expected to limit corrosion in all common types of plumbing, but that
there are useful measures that will restrict corrosion in any given system.
For steel pipes, it is important to recognize that the Langelier Saturation
Index is probably not a good indicator of corrosion protection. This is
certainly the case in soft waters, where buffer intensity is likely to be
much more important than the degree of saturation with respect to calcium
carbonate. Establishing conditions where siderite formation is favored would
be most effective in minimizing iron corrosion. A high dose of phosphate

XIX
will apparently be of some value in protecting iron pipes from corrosion, but
such a high dose is probably impractical in most systems.
On the other hand, copper corrosion in soft waters can be reduced by small
additions of phosphate. While our results did not show a dramatic effect of
pH on copper corrosion, other reputable studies have found such an effect.
Also, we investigated the beneficial effects of phosphate only at pH 8.0, and
so do not know the effectiveness of phosphate addition at lower pH values.
Where copper corrosion in soft waters is perceived to be a problem, addition
of phosphate and maintenance of a pH near 8.0 appears to be a very effective
corrosion control strategy. Copper piping is used primarily in premise
plumbing, and adjustment of water chemistry as the water enters the building
may be useful in these cases.
Protection of zinc galvanized pipe apparently requires addition of alkalinity
to soft waters. Adjusting the pH to 8.0 reduces corrosion dramatically
compared to maintaining it at pH 6.0, and raising the pH to 7.2 with addition
of large amounts of carbonate alkalinity is also effective. The exact
strategy that is appropriate or preferred in a given system will depend on
the characteristics of the particular system.

XX
INTRODUCTION

Corrosion is a multifaceted problem that can have economic, public health,


aesthetic, and political repercussions. The most significant problems caused
by corrosion include deterioration of water quality and the accumulation of
excessive concentrations of toxic heavy metals; an accumulation of corrosion
products that cause pumping costs to increase or flow rates to be reduced to
unacceptable levels; or structural deterioration of the pipe leading to leaks
or catastrophic pipe failure. Pipe failure is generally associated with
localized corrosion, and water quality deterioration and scale accumulation
are associated with more general or uniform corrosion. Regardless of whether
a corrosion problem is primarily health-related, economic, or aesthetic, it
is important to understand the causes so that appropriate actions can be
taken to combat them.

Corrosion of metallic pipes is an electrochemical process in which pipe metal


is chemically oxidized. Corrosion can occur whenever there is a difference
in electrical potential between two points on a metallic surface. The site
with the higher electrical potential becomes the anode (where the metal is
oxidized), and the site with the lower electrical potential becomes the
cathode (where some chemical species is reduced). There must also be an
electrical connection between the two sites (the pipe serves this purpose),
and a conducting solution between them.

Any heterogeneity in a pipe can lead to a difference in electrical potential


between two points and can initiate electrochemical corrosion. Because there
are likely to be minor metallurgical differences in the surface of any pipe,
this mechanism may be ubiquitous, although it may not be the dominant
mechanism.

Corrosion can also be initiated when the pipe surface is chemically attacked
by aggressive molecules in the water. For example, water in "dead spaces"
near pipe joints or cracks may have chemical characteristics somewhat
different from the chemical characteristics of the bulk water solution. The
difference in chemical composition between the two points leads to a
corresponding difference in the electrochemical potential, which can affect
1
the electrical potential on the pipe surface and cause a corrosion reaction.
Such a condition is referred to as a "concentration cell."

The electrical potential difference needed to initiate corrosion may also be


generated by electrical contact between dissimilar metals, which leads to
galvanic corrosion. Pitting is often associated with galvanic corrosion,
although pitting can result from other corrosive processes as well.

Regardless of the force driving the corrosion reaction, the practical ways to
reduce the corrosion rate are to interfere with the process by making it more
difficult for molecules to either approach or vacate the vicinity of the
anode and cathode, or to alter the electrical potential of the surface and
thereby lower the driving force for the reaction. This is, in essence, what
successful corrosion inhibitors do.

There is a large but relatively unintegrated body of knowledge concerning


corrosion and corrosion inhibition in water supply systems. That is, there
have been numerous studies of the effects of specific water quality variables
or water treatment processes on the corrosion rates of certain materials in a
given system, but there is relatively little in the way of unified theory
that has been developed, much less applied, to corrosion control in such
systems.

This research project represents a systematic investigation of the corrosion


of three common plumbing materials as a function of the chemical composition
of flowing water. The materials studied were copper and galvanized and
ungalvanized mild-steel pipes. Tap water at the University of Washington in
Seattle served as the baseline water. The water is of high quality,
containing very low concentrations of dissolved minerals. The chemical
composition of the water was altered by adding hydrochloric acid, sodium
bicarbonate, or sodium ortho-phosphate in various combinations, or by
stripping some of the dissolved C02 from the water.

The study differed from previously performed studies in significant ways.


First, both the design of the exposure loops and the analytical methods used
represent improvements in corrosion assessment protocols. The exposure
2
conditions were more representative of those in household plumbing systems
than conditions previously used in such tests. Additionally, the analytical
results are more accurate and their precision is better documented at low
corrosion rates than those previously reported.

Second, the tests were conducted using more than 1,000 individual coupons
that had been exposed to water of well-controlled and wel1-documented quality
over a period of up to 2 1/2 years. To our knowledge, such an extensive data
base has not previously been compiled in a manner that allows relatively
long-term corrosion rates to be compared as a function of water quality
variables.

Finally, this study is one of the first to correlate a variety of water


quality conditions with extensive information about the formation and
properties of corrosion scales. While the importance of such scales has long
been recognized, those few cases in which both water quality and scale
chemistry have been addressed have focused either on a single type of scale
or an evaluation of scales in pipes that had been exposed to uncontrolled
water quality over many years.

The relationships that have been established by this work and the data base
that the project makes available to other researchers will be of significant
value in cases where a corrosion problem currently exists and in the
development of the field of corrosion control.
LITERATURE REVIEW

INTRODUCTION

The use of chemical inhibitors to reduce corrosion rate is a relatively old


concept. Early information describing the effectiveness of inhibitors was
often based on such qualitative results as reduction in consumer complaints
about red water or fixture staining. Even in recent years, the problems
associated with conducting meaningful corrosion rate tests have often made
the results hard to interpret and too site-specific to extrapolate to other
water systems. Problems associated with the tests included an incomplete
understanding of the chemical interactions that occur among inhibitors, other
dissolved chemicals, and chemicals released by the corrosion reactions, and
an inability to evaluate the composition and quantity of protective scales
that form at the metal/water interface. Factors that are clearly important
in determining whether or not a particular inhibitor will be effective in a
particular system include initial water quality, type and concentration of
inhibitor used, material and condition of the pipes prior to the test, and
duration of the test.

A wide range of inorganic and organic inhibitors has been used in various
process industries and power-generating plants to protect plumbing from
highly corrosive waters. The vast majority of these inhibitors are
inappropriate for potable water supplies and are not discussed here. The
only chemicals commonly added to potable water for corrosion control are
bases to increase pH, carbonate-containing chemicals to encourage growth of
protective calcium carbonate or metal carbonate scales, orthosilicates and
polysilicates, orthophosphates and polyphosphates, and zinc (Zn). These
chemicals may be added individually or in various combinations. Evidence for
their relative effectiveness comes from a combination of short-term bench and
pilot-scale tests conducted by water utilities or engineering consultants,
observations made by these groups when a system-wide corrosion control
program was put into effect, and long-term investigations conducted by
various research groups.
ACTION OF CORROSION INHIBITORS

Most corrosion inhibitors used in potable water supplies act by forming a


protective scale over anodic or cathodic sites (some inhibitors can form a
scale over both types of sites). The scales are commonly inorganic
precipitates containing at least one of the ions added as inhibitor.
Although organic compounds are never used to control corrosion in a water
supply, natural organics can act as inhibitors by adsorbing to pipe walls,
complexing with Fe(II), and reducing the rate at which it is oxidized to
Fe(III).

The most effective and reliable form of corrosion inhibition involving scale
formation is passivation. In this process, an inorganic scale must first
grow on the metal surface. Typically, the scale first forms at the anode,
but grows to cover both anodic and cathodic sites (Ryder and Wagner, 1985).
Thus, it may act as a diffusion barrier to reactants and products of both the
corrosion half-reactions. In the presence of such scales, many surfaces
display a dramatic transition from being highly favorable to further
corrosion to being highly unfavorable to further corrosion as the electrical
potential at the interface between the scale and the underlying metal
increases. This transition occurs even though the thermodynamic driving
force for corrosion increases with increasing potential. The transition may
be related to a change in scale character from being an ionic conductor to an
electronic conductor (Bockris and Reddy, 1970).

The term "passivation" is used somewhat inconsistently in the literature.


Some authors use the term to describe any precipitated scale on the metal
surface that reduces corrosion rate. Other authors reserve the term for
cases where changes in the corrosion rate are caused by specific types of
changes in the surface electronic properties. Regardless of the mechanism
involved, this transition is a common one and can be used to great advantage
in corrosion control. The requirements are that a scale be formed and that
the potential be maintained at a sufficiently anodic value long enough for
passivation to occur.
CHEMICAL INHIBITORS IN POTABLE WATER

Phosphate. Polyphosphates. and Zinc Phosphate

Ortho-phosphate appears to be the most effective corrosion inhibitor


applicable to a wide range of plumbing materials. Although it is generally
not considered to be effective in preventing copper corrosion, there are
numerous reports of its ability to reduce the corrosion rates of zinc, iron,
lead, and galvanized piping. It is commonly applied either alone or in
combination with zinc, and is reported to be most effective for inhibition of
iron corrosion in a pH range of 6.0 to 7.5 (Ryder and Wagner, 1985; Hatch,
1973). At higher pH values there is some evidence that the presence of zinc
is particularly important for effective corrosion prevention.

Phosphate may be added as a soluble salt such as mono- or di-basic sodium


phosphate or in combination with zinc as zinc phosphate, "zinc bimetallic
phosphate," or "zinc glassy phosphate." When added in this latter form, it
may release either monomeric or polymeric phosphate species as it dissolves.
A typical dose might be 1 to 2 mg/L as phosphorus (P), with a dose three to
five times greater added for approximately one month to form the initial
scale (Karalekas et al., 1983; Richards and Moore, 1984; Muller and Ritter,
1980).

The specific action of ortho-phosphates in reducing corrosion has been


identified in only a few cases. In particular, when zinc is present, a zinc
phosphate scale can be inferred (Swayze, 1983) or has been observed by Auger
spectroscopy (Lumsden and Szklarska-Smialowska, 1978) in some systems where
corrosion inhibition was successful. The zinc may be added as zinc phosphate
or as a soluble salt, or it may already be present as a corrosion product.

Hatch (1973) reports that zinc does not alter the qualitative mode in which
ortho-phosphate acts as an inhibitor, and therefore the optimum conditions
for inhibition (e.g., pH range) are the same when applying phosphate alone or
with zinc. Nevertheless, he and others have found that zinc does allow use
of a lower concentration of phosphate for the same level of protection, and
causes the protection to develop faster for the same phosphate dose (Ryder
7
and Wagner, 1985; Hatch, 1973). However, there are also reports of systems
where zinc phosphate was ineffective in reducing corrosion of iron, copper,
or lead (Karalekas et al., 1983), or where zinc phosphate was effective but
bimetallic zinc phosphate (ZnNaPCfy) was not, even though the doses of Zn and
o-PO^ were quite similar (Muller and Ritter, 1980).

When zinc is not available to form a precipitate with the phosphate, a


protective scale may still form. On iron pipes, this most likely is a
ferrous or ferric phosphate. Although in theory phosphate ions could protect
a surface by sorbing onto a different, previously formed iron oxide phase,
Ryder and Wagner (1985) report that phosphate is most effective as an
inhibitor in the absence of prior scale formation, suggesting that the
phosphate must actually be part of the precipitate to be effective. In fact,
several workers report that phosphate will soften previously formed scales,
causing "red water" problems as the scale washes out of the system (Swayze,
1983). In these cases, the integrity of the scale may be compromised as the
mineral is converted from a hydrous iron oxide to an iron phosphate. Under
slightly different conditions, phosphate can reduce or eliminate red water
problems and simultaneously reduce head loss in corroded pipes by completely
solubilizing the scale (Swayze, 1983; Shull, 1980).

Since the tendency for a phosphate precipitate to form is pH dependent, it is


not surprising that pH has a significant effect on the effectiveness of
phosphate inhibitors. This has been observed qualitatively in a number of
studies, and the influence of pH on the competing precipitation and
complexation reactions has been demonstrated by Shull (1980). Shull's
research showed that when iron pipes were exposed to the same concentration
of bimetallic zinc phosphate at pH 7.2 and 8.2, the head loss through the
pipe, related to scale formation, was greater at the higher pH (precipitation
dominates), while the amount of iron in solution was greater at the lower pH
(complexation dominates). In neither case did they measure corrosion rate
directly.

Despite fairly widespread use of polyphosphates in the United States, there


appears to be no evidence in the literature that polyphosphates by themselves
inhibit corrosion of any common type of plumbing material. In those
instances where polyphosphates are considered useful, their effectiveness can
usually be attributed to their ability to complex and solubilize particulate
iron. Thus, polyphosphates can reduce pipe constrictions related to
excessive scaling and reduce or eliminate red water problems by dissolving
iron oxide particles. Also, under some circumstances and particularly at
temperatures greater than about 120 F, polyphosphates can degrade to
orthophosphate, which may be the active inhibitor in these systems (Ryder and
Wagner, 1985). For instance, Hatch (1973; 1952) reported that the
effectiveness of polyphosphates decreases significantly at pH greater than
about 7.0 due to a transition from a "thick electrodeposited protective
scale" to a "much lighter adsorbed scale." The latter scale was not only
nonprotective itself but also interfered with deposition of protective
calcium carbonate scales. Since iron forms soluble complexes with
polyphosphate, the protective scale that Hatch observed at pH less than 7.0
was most likely an oxide or a phosphate, not a polyphosphate.

Individually, there is no evidence that either phosphate or polyphosphate


reduces the corrosion rate of copper tubing. In fact, it is generally
thought that the only certain way to protect copper plumbing from corrosion
is by raising the pH. The pH needed to accomplish good corrosion control
depends on the other ions present in solution, but is generally in the range
8.0 to 9.0. However, Ryder and Wagner (1985) have recently shown that a
combination of orthophosphates and polyphosphates can reduce the rate of
general corrosion of copper dramatically, even though these compounds
individually are not effective inhibitors. The synergistic effect of these
two compounds has not been explained, but if it is a general phenomenon, this
may account for the very good performance of polyphosphates in preventing
corrosion in some systems where they may partially degrade to orthophosphate.

Silicate Inhibitors

The evidence for the effectiveness of silicates as corrosion inhibitors is


mixed, but recent investigations tend to indicate that their usefulness is
limited to relatively few materials and water quality conditions. The only
case where silicate inhibition of corrosion appears to be well documented is
in asbestos cement pipes, where its effectiveness may be attributed to a
surface-catalyzed conversion to quartz (Schock and Buelow, 1981).

Some of the older work on silicate inhibition suggested that it could protect
mild-steel pipes by formation of a continuous amorphous silica protective
scale, and that this scale could form only after other corrosion products had
formed an "anchoring" scale (Ryder and Wagner, 1985; Lehrman and Schuldener,
1952). However, Wagner (1985) claims that the beneficial effects often
ascribed to silicate can be explained solely by the fact that the silicates
raised the pH of the test water. He and his co-workers find that the same
corrosion reduction can be accomplished by raising the water pH with other
bases, and that the silicate itself provides no specific benefit. While this
may be the case at pH values greater than 8.0, which are generally used in
conjunction with silicate treatment, Hatch (1973) claims that there is little
support for using such high pH values, and Wood et al (1957) indicate that
silicates may in fact provide more protection at pH 7.0 than at higher
values. In any case, studies have shown silicates to have little or no
benefit in reducing corrosion of copper, zinc, iron, or lead under the
conditions in which they are commonly applied (Ryder and Wagner, 1985;
Wagner, 1985; Sontheimer et al., 1985).

Combinations of Inhibitors

As noted earlier, recent research has suggested that significant advantages


may accrue by combining various inhibitors. If one considers hydroxide ion
(strong base) to be an inhibitor, then the idea of using such combinations is
not a new one. Nevertheless, the idea of synergistic corrosion inhibition by
two or more chemicals has been minimally explored and warrants further study.

Carbonate-Based Inhibitors: CaC03 Deposition Indices

More than 100 years ago, Heyer (1888) described the reduction of lead
dissolution from plumbing systems by the formation of a thin protective
calcium carbonate scale on the pipe wall, induced by increasing calcium and
carbonate levels to the solubility level. Heyer's work led to the "marble"
test, still in use today, for the measurement of carbonate stability in water
10
supplies. A great deal of corrosion-control practice today is designed to
control CaCOj deposition, and it has become common to describe the
corrosiveness of a water in terms of its ability to form a protective calcium
carbonate scale on distribution pipes.

Langelier (1936; 1946) developed the concept of a saturation index (SI),


which he defined as the difference between the hypothetical equilibrium pH
for CaC03 solubility and the actual pH. The positive or negative magnitude
of the index indicates the degree of CaC03 supersaturation or
undersaturation. Largely because of the absence of a useful corrosivity
measurement, Langelier's concept of a saturation index gained widespread
acceptance and is frequently used today as an indicator of potentially
corrosive conditions. Several researchers have published diagrams that aid
in calculating the SI of a water (Caldwell and Lawrence, 1953; Lowenthal and
Marais, 1976; Merrill and Sanks, 1977), and others have developed alternative
approaches designed to give a more accurate approximation of the index.
Larson and Buswell (1942) improved the calculation by considering the effect
of ionic strength on the equilibrium constants. Ryznar (1944) modified the
Langelier approach and generated a stability index empirically related to the
degree of steel corrosion anticipated in a particular water. Dye (1952)
defined a saturation index based on a "momentary excess" (ME) in an attempt
to predict the amount of CaC03 scale deposition that could be expected from a
specific water supply. A qualitative and quantitative comparison of these
and other Si's has been provided by Rossum and Merrill (1983).

The maintenance of a positive saturation index is often advanced as a means


of reducing corrosion rates (Hudson and Gilcreas, 1976; Millete, 1980).
However, there is little evidence to support the conclusion that the SI
accurately predicts the corrosive properties of many waters. Langelier's own
work (1946) showed that the SI was of little value as a predictor when
applied to certain waters. Stumm (1960) and Larson and Skold (1957) explored
a variety of water conditions and found no consistent correlation between
corrosion rates and saturation index. Kuch and Sontheimer (1985) describe a
study indicating that when scales in iron-containing pipes have more than 10
percent CaC03 , the scales no longer protect the pipe from corrosion. Recent
work by Singley et al (1984) has also confirmed that saturation indices alone
11
cannot be directly related to the corrosion rate. As an alternative, they
proposed an empirical predictor for corrosion rates incorporating eight
variables, one of them a saturation index.

The historical inability to correlate CaC03 deposition indices in a


consistent and widely applicable way with corrosion rates raises the question
of why current treatment practice continues to place such reliance on them.
The answer appears to be simply the lack of a more accurate alternative, a
situation which reflects an inadequate understanding of the factors
controlling the formation and protectiveness of scales.

Recent research has focused more on the physical and chemical properties of
corrosion scales and has led to the development of a new model for corrosion
inhibition by the growth of protective scales in iron-containing pipes. This
research is reviewed briefly below.

SCALE COMPOSITION AND THE SIDERITE MODEL OF SCALE FORMATION

Scale formation is common in any system involving contact between metals and
natural water. The scale layer is often the principal protective agent
against internal corrosion in pipes, and its effectiveness determines the
useful life of the system (Hepburne, 1965). The major components of the
scale are corrosion products (metal and hydroxide ions), dissolved ions, and,
in many cases, calcium carbonate deposited from the water. The role that
corrosion products play in the formation of protective scales is not well
understood. The emphasis placed on CaC03 deposition demonstrates that it has
long been taken for granted that CaC03 by itself could form a protective
scale.

The calcium carbonate precipitation process is controlled, at least in the


early stages, by the electrochemical changes at the surface of the metal.
The corrosion process consists of the oxidation of the metal, coupled with
the cathodic reaction, predominated by the reduction of oxygen to OH". The
local pH increase at the pipe wall produced by the cathodic reaction
encourages CaC03 precipitation according to the following reaction:
Ca++ + HC03 " + OH' <=> CaC03 + H20
12
Other factors affecting the precipitation rate include the concentration of
dissolved Ca++ ions, the bicarbonate and total alkalinity, buffer capacity,
temperature, and liquid turbulence. Calcium carbonate precipitates in two
crystalline forms: calcite (hexagonal) and aragonite (orthorhombic).
Calcite is the more stable form at ambient temperatures and generally
comprises the precipitate in water piping. By contrast, aragonite forms a
considerable portion of the precipitate in hot water or boiler piping, the
critical temperature depending upon mineral content and magnesium
concentration (Pourbaix, 1973).

Baylis (1926) demonstrated that products of corrosion could be incorporated


into the common scale found on pipe walls. He also showed, through
theoretical calculations, that divalent iron produced in the corrosion of
steel pipes would precipitate primarily as FeC03 (siderite) if it is not
directly oxidized to Fe(III). Stumm (1956) followed this work with analyses
of scales formed in both soft and hard water. He reported that most of the
CaC03 was deposited near the pipe wall, while the total iron (Fe(II) plus
Fe(III)) concentration in the scale increased with distance from the wall.
Interestingly, a greater mass of CaC03 was deposited in the softer water than
in the hard water, but corrosion rates were higher in the soft water. In
both cases the bulk of the iron oxidized in the corrosion process was
retained within the scale.

Other researchers (Feigenbaum et al., 1978; McCauley and Abdullah, 1958) have
analyzed iron-pipe scales and identified a number of iron precipitates
enmeshed within the calcite, including siderite (FeCC^), goethite (FeOOH),
magnetite ^630^), and lepidocrocite (FeOOH). Kolle and Sontheimer (1985),
and Kolle and Rosch (1980) have examined scale morphology in detail and
determined that steel pipe scales form a readily identifiable "shell-like
layer" consisting of magnetite and goethite that divides the scale into
regions of differing composition. Beneath the shell-like layer (i.e., nearer
the pipe wall) siderite and goethite are found, while above the layer calcite
and goethite predominate. The authors suggest that a strong shell layer
encourages the production of siderite in the underlying layer (it shields the

13
Fe++ from rapid oxidation) and aids in fixing the scale components to the
metal surface.

Based on these observations, Sontheimer et al. (1981) proposed the siderite


model of scale formation. The model premise is that siderite (FeC03 ) is
critically important for the formation of a dense, uniform, and protective
scale on steel pipe walls, even though it is not a principal component of the
scale. By contrast, conditions hindering siderite formation tend to produce
nonuniform, thick scales that are both porous and readily sloughed. Such
scales are not protective and often lead to red water conditions.

Three reaction levels are defined in the model:


Primary Reactions
la. Fe -> Fe++ + 2e~
Ib. H20 + 1/2 02 + 2e~ -> 20H"
Ic. 2HC03 ' + 20H' -> 2C03 " + 2H20
Secondary Reactions
Ha. 2Ca++ + 2C03 " -> 2CaC03 (calcite)
lib. 2Fe++ + 2C03 " -> 2FeC03 (siderite)
lie. 2Fe++ + 1/2 02 + 40H" -> 2FeOOH (goethite) + H20
Tertiary Reactions
Ilia. 2FeC03 + 1/2 02 + H20 -> 2FeOOH (goethite pseudomorph) + 2C02
I lib. 3FeC03 + 1/2 02 -> Fe304 (magnetite) + 3C02

The primary reactions take place at the metal/scale interface and are usually
controlled by oxygen diffusion through the scale. The rate of diffusion,
coupled with the buffer intensity of the solution, determines the pH near the
pipe wall. This localized pH exerts strong influence on the competing
secondary reactions.

According to the reactions defined in the model, siderite formation will be


preferred over calcite if enough Fe++ ions are produced in the corrosion
reactions and if these ions are not further oxidized and precipitated by
goethite formation (reaction He). The controlling factor for goethite
formation appears to be buffer intensity. A low buffer intensity leads to a
high localized pH that favors this reaction, thus limiting the production of
14
siderite. The authors claim that field experience supports the hypothesis
that the siderite reaction path leads to a superior protective layer.

The tertiary siderite reactions involve the slow oxidation of the ferrous
iron in the siderite to the ferric form and the consequent production of
magnetite and a goethite pseudomorph. These two crystalline forms compose
the shell-like layer observed by Kolle and Sontheimer (1985). The shell
layer is formed directly from the siderite and has a dense impermeable
morphology. Although goethite crystal structure is generally less dense than
siderite, the goethite pseudomorph formed maintains the dense structure of
the siderite. This shell layer is important to the siderite model because it
retards the diffusion of oxygen to the metal surface, slowing the corrosion
rate and creating favorable conditions for further siderite formation. The
gradual conversion of the siderite to the ferric oxide form enhances the
shell layer and thus guarantees the integrity of the protective scale.

The siderite model contains many simplifications of complex equilibria


phenomena. However, it is the first corrosion model that incorporates
formation of different types of scales imparting differing degrees of
protection to steel pipes, and which is consistent with qualitative
observations of scale composition and morphology. The model also explains
the importance of buffer intensity in the scale formation process and
provides a framework for interpreting the possible mechanisms by which
corrosion inhibitors work.

SUMMARY

Clearly, a better understanding of the factors that control formation and


character of protective scales on corroding surfaces is needed, particularly
in cases where corrosion-inhibiting chemicals are added to a water supply.
Results of site-specific studies are fairly abundant and provide guidelines
for identifying the important parameters and possible outcomes when chemical
inhibitors are added. However, using these studies to draw categorical
conclusions about the effectiveness of inhibitors under a wide variety of
conditions is almost certain to fail, possibly leading to inappropriate

15
decisions about which inhibitor to use and ultimately to significant economic
and possibly public health costs.

16
METHODS AND MATERIALS

The majority of the corrosion-rate tests conducted for this research involved
the use of pipe coupons through which water was pumped. The water was stored
in reservoirs where its quality could be controlled. The following sections
describe the apparatus used in the tests and the preparation and analysis of
the pipe coupons.

INTRODUCTION

Two phases of experiments were conducted. In the first phase, pipe sections
were exposed to waters with a specific water quality in long-term
conditioning pipe loops. Pipe sections from the conditioning loops could be
removed after various exposure periods to determine weight-loss and to
analyze the corrosion scale. Corrosion scales were analyzed by powdered x-
ray diffraction and by wet chemical methods. X-ray diffraction was used to
determine the presence and identity of crystalline mineral compounds. Wet
chemical methods were used to determine the chemical composition of these
scales.

In the second phase of experiments, prescaled pipe sections were exposed to


changing water chemistry conditions. First, pipe sections were exposed to a
specific water quality for several months. Then, the water quality was
changed. In some cases, the pipes remained in the new water type for several
months. Before and after the water chemistry change, corrosion rates were
monitored using dissolved-oxygen-depletion or electrochemical corrosion-rate
measurements. At the end of the second set of experiments, the corrosion
scales on some pipe sections were analyzed to determine the effect of water
chemistry changes on the composition of scale.

WATER QUALITY IN THE EXPOSURE LOOPS

Seattle Tap Water

All of the waters studied were prepared from Seattle tap water. The source
waters for Seattle tap water are in mountainous regions receiving heavy
17
precipitation. The local geology in the watershed is largely granitic rock.
The source waters flow down steep streambeds, so there is little time for the
water to contact and dissolve minerals. These conditions create a water with
low alkalinity, low hardness, and low total dissolved solids. Seattle's
water treatment consists of chlorination, fluoridation, and the addition of
lime and sodium hydroxide for corrosion control. The aim of Seattle's
corrosion-control program is to raise the pH to 8.0 to counteract the water's
natural acidity and the acidity contributed by chlorination and fluoridation.
Without corrosion control, the pH would be about 6.0. The treatment does not
maintain a positive Langelier Saturation Index (SI); the SI after treatment
is still about -0.9.

The typical composition of the Seattle tap water used in this study is shown
in Table 1.

TABLE 1
Seattle Tap Water as Received at the University of Washington During
1986 and 1987.

Alkalinity 18.0 mg/L as CaC03


Hardness 28.0 mg/L as CaC03
Calcium 9.0 mg/L
Magnesium 1.2 mg/L
Sodium 1.8 mg/L
Potassium 0.3 mg/L
Chloride 3.7 mg/L
Sulfate 2.7 mg/L Sodium
Silica 12.0 mg/L
Conductivity: 50 micro mhos/cm
Langelier Index: -0.9
Buffer Intensity: 0.025 meq/L/pH unit

Water Quality Adjustment

The most important water quality parameters studied were pH and carbonate and
phosphate ion concentrations. Calcium ion concentrations were not varied.
Alteration of either the pH or the carbonate ion concentration (Cj) affects
the buffer intensity, saturation index, and conductivity. For example,

18
raising Cj increases the conductivity. Buffer intensity and SI are both
increased by increases in Cj at fixed pH. SI is also increased by increases
in pH. In contrast, when Cj is fixed, buffer intensity peaks at pH 6.3 and
then decreases as the pH is raised, going through a minimum at pH 8.3 and
then increasing again with further increases in pH (see, for example,
Figure 25).

Seven different waters, each with different water qualities, were studied
(Table 2). These water qualities were chosen in an attempt to isolate the
effects of ortho-phosphate and carbonate concentrations and pH variations on
corrosion rates.

TABLE 2
Characteristics of Conditioning Loop Water

Alkalinity Conduc^ Cone, in mq/1


pH mg CaC03/L tivity Ca Cl S04 o-P04 -
p
Water 1 6.0 10 70 7.6 17 2 0
Water 2 8.0 20 50 7.6 3 2 0
Water 3 8.4 20 50 7.6 3 2 0
Water 4 8.0 13 60 7.6 10 2 0
Water 5 7.2 400 650 7.6 3 2 0
Water 6 8.0 23 61 7.6 3 2 1
Water 7 8.0 30 81 7.6 3 2 5

Buffer Saturation Larson


Intensity-f Index Index

Water 1 0.310 -3.2 2.5


Water 2 0.025 -0.9 0.3
Water 3 0.024 -0.5 0.3
Water 4 0.016 -1.1 1.2
Water 5 96.5 -0.4 0.01
Water 6 -0.9 2.5
Water 7 -0.9 2.5

-{-Buffer intensity in meq/L/pH unit.


19
In the first water (Water 1 in Table 2), pH was adjusted to 6.0 using
hydrochloric acid, which lowered the alkalinity to 10 mg/L as CaCt^. As a
result, this water had the lowest pH, alkalinity, and saturation index of all
of the water types studied. The buffer intensity of this water, however, was
higher than that of three of the other waters studied, since its pH was close
to 6.3, where the buffer intensity goes through a local maximum.

The second conditioning loop water (Water 2) was generally unaltered Seattle
tap water with a pH of 8.0 and an alkalinity of 20 mg/L as CaCC^. Since this
water leaves the tap undersaturated with C02 , its pH drifts downward when it
contacts the ambient atmosphere. To compensate, small amounts of NaOH were
added as necessary to keep pH at 8.0 for this study.

The third conditioning loop water (Water 3) resulted from attempting to strip
water of carbonates by bubbling C02 -free air through Seattle tap water. This
raised the pH to about 8.4, while lowering the carbonate content by about 2
percent. Alkalinity was not changed. This high-pH water had a slightly
lower buffer capacity than unaltered Seattle tap water. Hydroxide ions
accounted for about 2.5 percent of the buffer capacity in this water.

This water was prepared so that the effect of dissolved carbonate species on
corrosion rates could be evaluated. The relatively small change in the
concentration of these species that occurred when the water was sparged with
C02-free gas made such an evaluation impossible. For a few months, the gas
sparging procedure was altered in hopes of increasing the amount of C02
stripped without altering other water quality parameters, but these efforts
were unsuccessful.

In order to facilitate C02 stripping, the loop-water pH was reduced to 6.1


using hydrochloric acid; lowering the pH converted carbonate alkalinity to
carbonic acid, which was then removed by sparging the water with C02 -free
air. After sparging overnight with C02 -free air, the pH was adjusted back to
8.0 using sodium hydroxide. This reduced the total dissolved carbonate by 30
percent compared to the unaltered tap water. Fewer equivalents of base were
needed to return the loop water pH to 8.0 than the equivalents of acid added
initially, because the carbonate concentration and buffer capacity had been
20
lowered by the gas stripping. After pH adjustment, the water (Water 4) was
ready for use in the conditioning loop. This procedure produced water with
an alkalinity of 13 mg/L as CaC03 and a slightly lower buffer intensity and
saturation index than Seattle tap water at pH 8.0.

Both the pH 8.4 water, which had been decarbonated at alkaline pH, and the pH
8.0 water, which had been decarbonated under acid conditions, had
significantly lower concentrations of the fully-protonated ^03 species than
any of the other loops. Coupons had been in the loops for various exposure
periods when the change in procedure occurred. Rather than discard these
coupons, an effort was made to determine whether the change affected their
corrosion rates. For the black-iron coupons, such an effect was noted, and
the corrosion rates for each low-carbonate exposure period are discussed
separately. However, for the other coupon types there was no significant
difference in corrosion rate between the two periods, and the data have been
grouped and discussed collectively. In these cases, the water quality during
both exposure periods is referred to simply as "pH 8.0 low carbonate" water.

The fourth conditioning loop water (Water 5) had a very high alkalinity
produced by adding sodium bicarbonate. In addition, the pH was reduced to
7.2 using carbon dioxide gas. The high alkalinity and lower pH resulted in a
large increase in buffer intensity; the buffer intensity was almost 4,000
times greater than Seattle tap water at pH 8.0. Even with the high
alkalinity, this water still had a slightly negative SI due to the extremely
low calcium content of Seattle tap water.

Two loops were used to investigate the effectiveness of orthophosphate as a


corrosion inhibitor. The solution in both loops was maintained at a pH of
8.0 with 1 mg/L o-P04 -P added to one loop (Water 6), and 5 mg/L o-P04 -P to
the other (Water 7). These concentrations were chosen to examine the effects
of ortho-phosphate at a typical "maintenance" dose and at a high, initial
conditioning dose. The phosphate was added as

In the low-carbonate and high-carbonate conditioning pipe loops, gas was


continuously sparged into the bottom of the reservoir to alter the dissolved
C02 concentration of the circulation water. In these reservoirs, a floating
21
cover was used to minimize the water surface area contacting the ambient
atmosphere. In the other reservoirs, the water was allowed to equilibrate
with the atmosphere.

Addition of hydrochloric acid to the pH 6.0 water and to the low-carbonate


water resulted in an increase in the chloride ion concentration. It was
hoped that this chloride addition would not appreciably affect corrosion
rates. Unfortunately, the acid addition resulted in a large relative
increase in chloride ions, since the initial chloride concentration of
Seattle tap water is so low. The Larson Index is an indicator, based on
empirical observations, that attempts to describe the corrosive effects of
aggressive ions like chloride. The Larson Index is calculated by dividing
the aggressive ion concentration (total equivalent concentration of chloride
and sulfate) by the bicarbonate ion equivalent concentration. As shown in
Table 2, the Larson indices of these two conditioning waters were
significantly changed.

Larson and Skold (1957) found that iron corrosion rates significantly
increased as this ratio was increased from 0 to 0.7. The waters they studied
had alkalinities between 75 and 260 mg/L as CaC03 , much higher than Seattle
tap water. Higher alkalinities, plus other differences between their study
and the current one, make it impossible to predict the exact effect of
increased chloride ion concentrations on the corrosion rates measured in this
study. The increased chloride concentration of these two conditioning waters
must be considered when analyzing the results.

When acid, base, and sodium bicarbonate were added to the conditioning
waters, the conductivity was increased. The high carbonate water had a much
higher conductivity than all of the other waters; its conductivity was almost
100 times that of Seattle tap water.

Loop-Water Monitoring

In order to prevent accumulation of corrosion products in the conditioning


loops, the Seattle tap water loop and the pH 6.0 loop were changed three
times a week, while the other loop waters were changed twice a week. The
22
twice-weekly change schedule for the other loop waters was chosen because of
the time-consuming procedure needed to obtain these water conditions.
Between water changes, the pH of all the waters was monitored and adjustments
with acid or base were made if necessary. The pH was monitored using an
Orion Research analog pH meter, Model 301.

As the conditioning waters equilibrated with the atmosphere, their pH


drifted. Seattle tap water is undersaturated with C02 as it leaves the
treatment plant due to the addition of corrosion treatment chemicals
that raise the pH. Therefore, the pH of the pH 8.0 water tended to
decrease as the water equilibrated with the atmosphere. Likewise, the
pH of the low pH water tended to drift upwards between pH adjustments
because it was supersaturated with C02 after acid addition. The pH of
loop waters sparged with C02 -free air drifted upwards between water
chemistry adjustments as C02 was stripped from the water. In the high
carbonate water, pH remained constant because it was sparged with a
constant composition CC^-plus-air mixture and was at equilibrium before
it was circulated. In all cases, pH varied by less than 0.2 units, so
the target pH values adequately represent the actual loop-water pH
values.

CORROSION RATE MEASUREMENT

Introduction

The scope of this project required the application of innovative techniques


to analyze instantaneous and long-term corrosion rates. One such technique
was a modified version of an ASTM standard for weight-loss measurements on
pipe coupons. The modified version used simplified hardware suitable for
sampling large numbers of coupons. Coupled with new handling and retrieval
procedures, the technique offers improved precision and a more accurate
statistical analysis of corrosion rates. For instantaneous electrochemical
rate measurements a polarization cell was developed that uses pipe coupons as
the test surface and simulates the conditions and flow characteristics of an
actual pipe section. The cell can be used to make corrosion-rate
measurements or to map variations in corrosion potential along a pipe
23
surface. A multipoint polarization procedure can be used to develop
estimates of electrochemical parameters that define the redox characteristics
of the corroding surface. A discussion of these and other corrosion
measurements follows.

Lono-Term Corrosion Rates

Of the different corrosion-rate measures, long-term weight loss, i.e., the


cumulative loss of metal over an extended period of exposure, is the
measurement of greatest interest to people operating distribution networks.
Such measurements generally take two forms. The first and oldest of these is
the flat-coupon approach, where a thin rectangular piece of the metal of
interest is inserted via an insulating stem into the flow stream of a
distribution line. The standardized procedure is described by ASTM D 2688-83
Method B (ASTM, 1983a). The point of insertion is usually an exposed tee,
elbow, or corporation stop. Although other materials can be used, the coupon
is most often stamped from mild carbon sheet steel and is intended to
represent the pipe wall of a black-iron (ungalvanized mild-steel) service
line.

A second approach is the machined-nipple test developed at the laboratories


of the Illinois State Water Survey (ISWS)(Larson, 1975). It has been adopted
as a standard and is described by ASTM D 2688-82 Method C (ASTM, 1983b). The
technique is a significant improvement over the flat-coupon technique since
it uses actual pipe sections as the corrosion test surface. The black-iron
pipe sections, generally of one-in. diameter and four-in. length, are
machined to fit snugly inside a PVC nipple (Figure 1). The exterior and ends
of the pipe insert are coated with an epoxy resin to prevent corrosion on
non-wetted surfaces. In operation, the nipple assembly can be plumbed into
any convenient delivery line or used within a closed-loop system for
laboratory testing. A particular advantage of the machined nipple is that
flow conditions within the test section reproduce the effect of pipe
hydrodynamics, hence the potential effects of impingement corrosion can be
studied, and the growth and erosion of oxide scales and scale on pipe walls
can be duplicated. The minimum recommended exposure period is 120 days. If
possible, exposures of 12 months are preferred. Singley and Lee (1984) have
24
1-'/4 in ' UNION

VVV^^S/x^^v/N/
K)
Ul 1-'/4 in ' PIPE NIPPLE
10 in (254 mm) LENGTH
1.'/4 in* x in" BUSHING

1.66 in or 42.16 mm O.D.


J in ' SERVICE LINE
"1.315 in or 33.40mm O.D.

- v s~~
1 , -i
f
t V - <j
J \^__ A
<ISWS CORROSION TESTER

Figure I. Cross Section of Insert, Spacer, and Union, and Assembled ISWS Corrosion Tester
developed a modified version of the ISWS approach in which multiple coupons
are exposed.

The ISWS procedures for surface preparation of the pipe section coupons
involve several steps that must be performed carefully in order to give
reproducible results. Proper restoration of the coupon surface following
exposure is also critical to the weight-loss measurement. The somewhat
cumbersome sequence for preparing pipe section coupons consists of drying for
24 hr in a desiccator, phenolic stripping of the epoxy-resin exterior
coating, scouring of the interior surface using abrasive powders, exposure to
an inhibited acid solution to loosen deposits and etch the surface, immersion
in an ultrasonic bath, water and acetone rinses, another 24 hr of drying, and
final weighing. Handling of each coupon requires several hours of effort
spread over a two-day period.

Given the number and nature of procedures in the sequence, the possibility of
measurement errors or weight loss during sample preparation is significant.
However, the issue of precision in corrosion-rate measurements of this type
is rarely addressed. Even ASTM standards do not estimate the precision of
the resulting measurement, stating instead that this must be considered a
function of the individual system. In the literature, it is rare to find
confidence intervals or other quantitative indicators of measurement
precision. One study reported a standard deviation of 0.64 milligram per
square decimeter per day (mdd) compared to absolute rates of about 1 to 6 mdd
for corrosion of galvanized steel specimens using ASTM D2688 Method C (Neff
et al., 1987).

This project used a modified version of the ASTM machined-nipple test that
simplified the hardware and preparation technique. To improve the precision
of the measurements, the system was designed to hold a relatively large
number of coupons (pipe inserts), yet still offer ease of access and
simplified handling procedures. The following items are specific advantages
of this modified version:

26
1. The system is capable of supporting coupons made from a variety of
common metal plumbing materials (copper, galvanized-iron and black-
iron pipes).
2. It can be easily duplicated in other laboratories, using readily
available materials.
3. Coupon preparation and restoration procedures are both simple and
reliable.
4. The use of multiple coupons increases the precision of the overall
measurement, extending its use to the analysis of short-term weight
loss tests (exposure periods of less than 14 days).

Pipe-Loop Construction and Coupon Materials

Figure 2 presents the cross section of a typical acrylic-sleeve/pipe-insert


assembly. It consists of two headpiece/plunger mechanisms that fit in
opposing ends of a flanged sleeve. The headpieces are attached to the sleeve
flanges by bolts and generate a compressive force on the pipe inserts held
between them. The coupons are electrically isolated and sealed end-to-end by
washers cut from hard red-rubber gasket material. The wetted area (corrosion
test surfaces) is restricted entirely to the interior of the coupons. Up to
10 coupons may fill the standard 20-in. sleeve. Longer sleeves holding more
coupons are not recommended, since more seals require a compressive force
greater than the sleeve wall can safely carry. When one or more coupons are
removed, they can be replaced either by fresh coupons or by sections of PVC
pipe cut to the appropriate length.

For 1/2-in. black-iron and galvanized-pipe coupons, both with 0.85-in. OD, a
7/8-in. ID acrylic tube makes a satisfactory sleeve. The annular gap of
approximately .01 in. between the coupon exterior and the sleeve wall
provides an expansion area for the gasket material under compression. The
slight gap also makes loading and unloading the sleeve easier, and allows air
circulation to carry away moisture that may condense on the coupon exteriors.
The transparency of the acrylic is especially useful when seating the coupons
end-to-end, or when checking for possible leaks. All major pieces of the
sleeve assembly are machined from readily available acrylic stock. Other
materials (unions and nipples) are standard 1/2-in. PVC fittings.
27
Headpiece x Plunger
Vvlt UIIIUII
Pipe Coupon -\ / s.
j j "-=--1
IwT^I u
Lr^1 ' 1 1KICHr- nHr
. -/ 4 * i j
2.0"
Gasket
r^ir^.,._jQc *
Flexible Pvc
Nipple j L B- \ Acrylic Sleeve
> <>
1.5* 1,0'

Figure 2. Cross Section of Coupon/Sleeve Headpiece Assembly Used for Weight Loss Test
Standard 1/2 or 3/4 inch copper tubing can also be used (types K, L, or M) as
the coupon material; however, the thin wall of standard copper tubing (type K
is the thickest at 0.05 in.) makes it difficult to stack and seal the ten
coupons in a single sleeve. Because of the minimal butt-end exposure,
alignment of a large number of coupons is a problem that often leads to water
leaks at the many gasket seals. This task proved so troublesome that a
thick-wall copper axle stock was substituted for the standard water tubing.
The thicker walls give a broader shoulder against which to seal; this tubing
is also more durable than thin-wall stock and can be reused. The axle stock
selected has an OD of 0.675 inches and a wall thickness of 0.14 in. This
tubing is a minimum 99.9 percent pure copper, with a maximum phosphorus
content of 0.05 percent. This is a composition typical of most copper tubing
and appurtenances, and is nearly identical to ASTM specifications for the
three common water tubing types, which must have a purity of 99.9 percent and
a phosphorus content less than 0.04 percent (Cruse et al., 1985).
Microscopically, the surface of the axle stock appears free of pits and any
obvious inclusions, identical to the degreased surface of copper water
tubing. The axle stock and tubing form similar corrosion scales when exposed
to the same water for equivalent durations. An acrylic sleeve with an ID of
0.75 in. works well for mounting coupons made of this material.

Black-iron and galvanized-iron pipe coupons are cut from standard 1/2-in.
pipe, ASTM A 120 and ASTM A 120-82, respectively. Figure 3 presents a
photograph of the three coupon types. The ends of the coupons are lathe-cut
to ensure a smooth, flat finish that seals tightly against the gaskets. The
standard length of coupon is 4 cm, producing a coupon weight of roughly 50 g,
a size easily handled on most precision balances. Black-iron and galvanized-
iron coupons have an interior surface area of 18.5 cm , while the thick-
walled copper coupons present an area of 12.5 cm2 . After machining, the
coupon exteriors are degreased to remove cutting oil, polished, stamped with
an identification number, and given an exterior coating of clear acrylic to
protect them during handling and to prevent moisture-induced corrosion. The
acrylic coating is applied to the butt ends of the coupon as well.

29
Figure 3. Black Iron, Galvanized Iron, and Copper Coupon/Inserts

30
Figure 4 is a diagram of the closed recirculation loop and manifold assembly
used to determine the effect of specific water quality conditions on
corrosion of different metal types. Three sleeves, each containing a
different coupon type, are held in the manifold. The sleeves are connected
to the manifold by single-union ball valves, making it simple to isolate and
remove individual sleeves/coupons during loop operation. In all cases, a
flow rate giving a fluid velocity of 1 foot per second (fps) was used,
producing a flow pattern well within the turbulent regime.

One test was conducted after the loops had been in use for approximately 10
months to determine the relative flow rates through the three arms of the
apparatus. In this test, water at the normal flow rate of 1.5 gallons per
minute was pumped through the pH 6.0 and pH 8.0 Seattle tap water loops, and
the flow through each arm was measured independently using a graduated
cylinder to collect the exiting fluid.

Visual inspection suggested that the diameter of both the galvanized and
ungalvanized steel coupons had been reduced by about 15 to 20% respectively,
corresponding to about a 40% reduction in cross-sectional area. The flow
distribution in both loops was about the same, with about half the total flow
(0.75 gpm) passing through the arm with the copper coupons, and about one-
quarter the total flow (0.37 gpm) passing through each of the other arms.
These flow rates correspond to linear velocities of 1.9 ft/sec through the
copper coupons and 0.45 ft/sec through the other two arms, if we ignore any
reduction in cross-sectional area due to scale growth. If we use the
estimate of 40% reduction in cross-sectional area for the steel coupons, the
linear velocity through them becomes 1.1 ft/sec.

Although there may have been some effect on mass transport of reactants and
corrosion products due to the differences in flow rates among the three
loops, these effects were probably small, since flow through all the loops
was well into the turbulent regime. The effect on the rate of scale erosion,
which is related to the linear velocity of flow through the pipes, may have
been more significant.

31
Pipe Coupon
Sleeves

Rotometer
(Flow Indicator)

Flow Control.
Valve

Pump

Figure 4. Pipe Coupon Exposure Loop

32
There were two practical ways to conduct the long-term exposure experiments.
In one, headloss could be equalized in the three arms, by hydraulic linkages
via the headers above and below the pipes. Alternatively, the flow rates
through the three loops could have been equalized using valves and flow
measuring devices such as rotameters. In neither case would the linear
velocities through the loops be constant over the exposure period. The
variation in linear velocity is expected to be less in the system chosen than
in the alternative, since scale growth would decrease the flow rate and
thereby lessen the effect of reduced cross-sectional area. In this work, the
apparatus was set up in such a way that the total flow through all the loops
was constant, but the relative flows through the three arms varied while the
headloss through each arm was equivalent.

Surface Preparation for Black and Galvanized Iron Coupons

Unlike the surface preparations recommended in ASTM procedures, the approach


used in this study relied entirely on mechanical surface conditioning. The
intent was to eliminate the time-consuming requirements of successive acid
rinses and other chemical treatments needed to remove the corrosion scale and
to restore the original metal surface. The use of mechanical brushing and
polishing procedures, which can be duplicated almost exactly from one coupon
to the next, improves the overall precision of the weight-loss measurements.
At the same time, mechanical cleaning greatly speeds the process, while
reducing the overall effort.

A hand-held high-speed polishing tool (Dremel Corp., Racine, WI) with


variable-speed motor (up to 10,000 rpm) was used to prepare all three types
of pipe coupons. Different toolpieces were used depending on the metal type
and inside diameter (ID) of the coupon. Figure 5 shows a photograph of a
1/2-in. black iron coupon and the angled rotating brush used to prepare its
interior surface. The brush (available from Dremel Corp.) has a flexible
stainless-steel bristle and a diameter just slightly less than the ID of the
coupon. The brush passes easily through the coupon center. As it rotates,
the slanted bristles bend outward and span the gap between brush and coupon
wall. Experience has shown that it is neither necessary nor desirable to
apply pressure against the coupon wall while cleaning. Excessive pressure or
33
Figure 5. Black Iron Coupon with Polishing Tool

34
rotational speed will shorten the life of the brush (which is approximately
25 cleanings) and possibly abrade noncorroded metal from the coupon wall. A
relatively low speed of 2,000 rpm was used with this toolpiece. All but the
thickest corrosion scales were easily removed with a few passes of the brush;
even heavily pitted surfaces could be quickly cleaned.

Galvanized iron coupons have nearly the same ID as black-iron coupons and
were worked with the same toolpiece. However, because zinc is a softer metal
than steel, it was especially important not to abrade the coupon surface by
applying pressure against the coupon wall. Fortunately, the zinc corrosion
scales are less adherent than iron and can be removed in one or two quick
passes of the brush without disturbing the underlying galvanized (zinc)
layer.

In cleaning a galvanized-iron or black-iron coupon, use of the rotating brush


did not restore the metal surface to a lustrous patina, although this was
possible with pressure and high rpm. A shiny surface requires the abrasion
of the underlying metal and adds to the imprecision of the weight-loss
measurement. The desired surface, whether new or restored, is smooth and has
the dull-grey luster characteristic of new pipe before it is put into
service. For good precision in short-term exposure studies, it is important
that the surface conditions of the different coupons be as nearly identical
as possible. The initial surface condition is somewhat less critical in
longer studies, since corrosion processes will "normalize" the surface,
regardless of the initial metal patina.

Whether the coupon test surfaces were being prepared for the first time or
were being cleaned following exposure, use of the rotating brush required a
thoroughly dry surface. When dry, the scales are friable and can be reduced
to a fine powder that brushing action quickly disperses. If wet, the scale
will form a gritty paste that is difficult to remove and causes excessive
abrasion to the underlying metal. For a thin scale, a few hours of drying in
a desiccator was sufficient. Thicker scales (greater than 1 mm depth)
required up to 24 hr. Drying ovens were not used, since they may soften the
exterior acrylic coating. However, rapid drying could be achieved using a

35
hot-air blower (the type used for paint stripping) if care was taken to avoid
concentrating the heat on the coupon exterior.

Following use of the rotating brush, the interior surfaces of the coupons
were given a final polishing using clean, dry cotton swabs. This removed
residual powdered scale and generally left the metal surface in a condition
close to its original form. Because of the acrylic coating, the coupon
exterior and butt-ends could be quickly cleaned with a lightly moistened
tissue. To avoid soiling the exterior surfaces, cotton gloves were worn when
handling the coupons. After final polishing, the coupons were blown dry,
then stored in a desiccator prior to weighing. Excluding the initial drying
time, a group of 10 exposed black-iron or galvanized-iron coupons could
easily be cleaned, polished, and weighed in 1 hr, compared to ASTM procedures
requiring 3-4 hr of work spread over a period of 48 hr.

The principal procedures used in the initial cleaning or restoration of the


interior surfaces of black-iron and galvanized-iron coupons are listed in
Table 3.

TABLE 3
Summary of Procedures for Black-Iron and
Galvanized-iron Coupon Preparation/Restoration

1. Thoroughly dry the coupons in desiccator, or blow dry with hot-air


gun. Do not dry in oven.
2. Use 2,000 rpm rotating brush to dislodge dried scale and to polish
surface. Avoid overpolishing.
3. Swab coupon interior clean with dry cotton-tipped applicators.
4. Clean acrylic-coated exterior with moist tissue.
5. Blow dry again; store in desiccator prior to weighing.
6. Weigh coupons to the nearest 0.1 mg.

36
Surface Preparation for Copper Coupons

Unlike zinc or iron, corroding copper forms a dense, uniform scale that is
tightly bound to the metal surface. The protection afforded by this scale is
largely responsible for copper's relatively low corrosion rates. The
tightness of the scale makes it more difficult to remove than either zinc or
iron scales, and the removal process is further complicated by the metal's
softness, which precludes the use of steel-bristle polishing tools.

Figure 6 is a photograph of the toolpiece used in cleaning the interior of


copper coupons. It is an elliptical cone of pressed felt (Dremel Corp.) with
a maximum diameter slightly smaller than the ID of the coupon. In contrast
to the other coupon types, copper was cleaned when wet, using a nonabrasive
detergent solution (Liquinox, Alconox Inc., NY) as a polishing aid. To
generate the friction required to dislodge the scale, the tool was turned at
high speed (5,000 rpm). The wetted surface helps to dissipate the heat and
avoid burnishing the metal. The abrasion necessary to dislodge the scale
comes from the hard copper oxide particles released during the polishing
process. The detergent solution keeps the particles in suspension and helps
to rinse the surface clean at the end of the process.

Because it is possible to overpolish the metal, some practice is required in


recognizing when the surface is sufficiently cleaned. The intent was to
return the surface to the bright luster of a new copper penny. As in the
case of mild-steel and galvanized-iron coupons, excessive polishing can
remove non-oxidized metal and decrease the accuracy of the weight-loss
measurement. Although different degrees of polishing are a potential source
of variability in this test, the variation can be minimized by practice. In
the current study, polishing was done by only three individuals, and
comparison of coupons polished by these three people established that they
polished the coupons to an extent that was visually indistinguishable.

As with the black-iron and galvanized-iron coupons, mechanical cleaning of


copper was quick, requiring only slightly more effort than the other two
types. However, more care was necessary to ensure that unoxidized metal was
not removed from the copper surface. Under equivalent conditions, the
37
Figure 6. Copper Coupon with Polishing Tool

38
corrosion weight loss from copper may be an order of magnitude less than that
of black iron; hence, careful polishing was critical to the precision of the
weight-loss measurement. The procedure is summarized in Table 4.

TABLE 4
Summary of Copper Coupon Preparation/Restoration Technique

1. Apply a small amount of nonabrasive detergent paste to a wetted


copper surface.
2. Using felt tool, polish at 5,000 rpm until underlying metal exhibits
a bright luster. This should not require more than 60 seconds for
even a highly tarnished coupon.
3. Rinse coupon interior with distilled water and swab dry with cotton
applicators.
4. Wipe acrylic-coated exterior clean with moistened tissue.
5. Dry in desiccator, or blow dry. Weigh coupons to the nearest 0.1
mg.

Shortly after the start of this research, it became obvious that the copper
coupons would not generate adequate corrosion scale in a reasonable time
period with which to perform x-ray analysis. To solve this problem, flat
copper plates similar in composition to the copper pipe coupons were exposed
to the same water qualities.

The plates were cut to a size that allowed mounting in the specimen holder of
the x-ray diffraction equipment. Prior to being placed in the exposure-loop
reservoir, the surface tarnish was removed using steel wool. Five plates
were installed on a plexiglass plate that was approximately 1 in. wide and 36
in. long. Each plate was held by plastic compression devices at specific
intervals such that the lowest plate was 6 in. below the water surface.

39
While the loop water was being changed during the thrice-weekly water quality
maintenance, the plates were kept submerged in loop water in a separate
basin.

Technique Validation

Imprecision in the weight-loss measurement comes either from handling and


preparation/restoration procedures or from natural variations in corrosion
rate produced by differences in metallurgical properties or chemical
conditions on the coupon surface during exposure. Aside from carefully
selecting and machining the coupon materials, little can be done to alter the
effects of the latter. These effects, however, are probably of minor
importance when compared to the potential impacts of inadequate handling
procedures. A series of tests was conducted to assess the reproducibility of
the preparation/restoration procedures and the precision of the overall
measurement. The results of these validation studies are presented in
Appendix Al. In most of the corrosion tests in this study, the weight loss
per coupon was on the order of a few tens of mg for copper coupons, and
hundreds of mg for the steel coupons (both galvanized and ungalvanized).
Based on the data in Appendix Al, the error associated with the coupon
preparation technique was at most a few percent of the total weight loss.

ELECTROCHEMICAL CORROSION-RATE MEASURES

Introduction

In addition to weight-loss measurements, corrosion rates of copper, black


iron and galvanized-iron pipes were evaluated by electrochemical techniques.
A cell configuration was chosen that uses pipe coupons as the working surface
and simulates the conditions and flow characteristics of the distribution
system. The pipe coupons that served as the electrode working surfaces (test
electrodes) are identical to those described in the weight-loss section
above. A six-point polarization procedure was used to estimate the two
relevant electrochemical parameters: the corrosion current and Tafel slopes.
Complete descriptions of electrochemical principles and validation tests are

40
presented in Appendix A2. The polarization cell and analytical procedures
are described in the following section.

Several different electrochemical techniques have been used to measure


corrosion rates in drinking water systems, the most common and successful of
which has been polarization resistance, also called linear polarization.
This technique has been widely used since it was first described by Wagner
and Traud (1938) and subsequently expanded by Stern and Geary (1957). The
basic relationship derives from the modified Butler-Volmer equation, which is
discussed in most electrochemistry textbooks.

Commercial polarization-resistance instruments are produced in several


electrode configurations. The most popular design is a three-electrode
system manufactured by Petrolite Corporation (Houston, TX), in which
cylindrical electrodes machined from the metal of interest are mounted
parallel to each other in a triangular pattern. Each electrode corrodes
freely and establishes its own corrosion potential. In operation, one serves
as the reference electrode, one as the counter electrode, and the remaining
one as the test electrode. The difference in freely corroding potential
between the reference and test electrodes is measured, and an offset
potential is applied to the test electrode. After a preset equilibration
time, the impressed current is measured and the corrosion current is computed
using the Stern-Geary equation (Stern and Geary, 1957). In a commercial test
unit, all of these steps are carried out automatically, and the output is a
single number that is a measure of the corrosion rate. This simplified
technique is appropriate for qualitative assessments of corrosion conditions
and for determining relative rates of corrosion in similar systems (Reiber et
al., 1987).

The polarization cell and procedures used in this study constitute a


significant improvement over the linear polarization procedures described
above. Developed in part through a research program sponsored by the U.S.
Environmental Protection
*
Agency, they offer several advantages over the more
conventional approach:

4!
1. The new polarization cell is easily constructed and can be used with
a variety of potentiostatic devices without having to rely on
expensive proprietary instrumentation.
2. The cell uses actual plumbing materials (black-iron, copper, and
galvanized-iron pipe sections) as the corrosion-measurement surface.
3. Cell geometry duplicates the hydrodynamics of pipe flow and its
impact on the corrosion surface.
4. The technique allows one to determine the values of fundamental
variables controlling the corrosion process (e.g., the anodic and
cathodic Tafel slopes) for the specific surface and water quality
conditions of interest.

Polarization Cell and Electric Circuit

A cross section of the polarization cell is presented in Figure 7. As in a


typical polarization assembly, it consists of test, reference, and counter
electrodes, but the configuration and geometry are unique. Flow is axial
through the test-electrode section and can be adjusted to simulate desired
velocities or Reynolds numbers. The test electrode itself is made from
sections of water pipe (1/2-in. black iron or galvanized pipe, 5/8-in. type K
hard copper tubing, or the copper axle stock previously described in the
weight-loss section). The counter electrode penetrates the axial center of
the cell along its length and is thus equidistant from all points on the
test-electrode surface. This prevents distortions in the current flow during
polarization extremes, even when ohmic resistance may be significant. The
reference electrode is a Ag/AgCl electrode encased in a narrow, flexible
polyethylene stem. A ceramic frit junction fits into the end of the stem and
can be recessed any desired distance. The junction frit is porous and has a
high leak rate, ensuring an exceptionally stable reference voltage. A
principal advantage of the flexible electrode is that the junction can be
positioned at any location in the polarization cell. This is particularly
useful for studying the variation of corrosion potential across the test-
electrode surface.

In this research, a modified RDE2 Potentiostat (Pine Potentiostats, Pine, PA)


was used to conduct polarization scans, although any high-quality
42
To Potentiostat

Ag-AgCl
Reference
Electrode

Flow Out

Polyvinyl tube (filled


" with 3M KC1)
Pipe Coupon Ceramic
Test Electrode Frit

Platinum Counter
Electrode

Flow In

To Potentiostat

Figure 7. Polarization Flow Cell

43
potentiometer capable of set-point polarizations could be substituted.
Analog output from the potentiostat was converted to digital form by a
Keithley Series 500 Data Acquisition System (Keithley Co., Cleveland, OH).
Data was then transferred to an IBM-PC-based microprocessor for statistical
analysis, graphing, and storage.

The principal function of the polarization cell was to measure instantaneous


corrosion rates and evaluate the effect of water quality transients on
corrosion of surfaces with preformed scales. To this end, tests were
conducted in a closed-loop recirculation system where precise water quality
conditions could be maintained (Figure 8). Mechanisms were provided to
control temperature, flow rate, and dissolved gases, including
instrumentation for the continuous monitoring of pH, temperature,
conductivity, and dissolved oxygen. Chemical addition ports for inducing
transients during monitoring, as well as sampling ports to test for the
release of corrosion products, were included.

General Polarization Measurements

In general, corrosion rates measured on plumbing materials are an order of


magnitude lower than those observed for iron in seawater, or the down-hole
tooling of an oil drilling rig. Most electrochemical measurements in these
environments deal with corrosion current densities measured in milliamps/cm^,
while current densities on common water distribution plumbing materials are
more likely to be in the microamp/cm^ range. It is usually easier to make an
electrochemical measurement on a rapidly corroding system for two reasons:
first, the current densities are higher and thus more accurately measured;
and second, a rapidly corroding surface returns more quickly to its freely
corroding condition following the application of a polarizing current.

Because distribution system corrosion rates are relatively low, maintaining


the freely corroding condition of the test-electrode surface is crucial for
an accurate measurement. The task is complicated because the freely
corroding surface is readily affected by mechanical action, such as scale
removal, or by electrochemical actions that favor either the anodic oxidation
44
SAMPLING pH FLOW TEMP COND DO
PORT
I I I I I
GAS
SPARGER
SENSORS

3 COUPON
SLEEVE

TO
POTENTIOSTAT

STIRRER
HEATER PUMP POLARIZATION
CELL

Figure 8. Schematic of Closed-Loop Recirculation System for Polarization Measurements

45
or the cathodic reduction processes. This means that polarizing the surface
by impressing a current density of several milliamps/cm^, as would be done
for a rapidly corroding surface, is not appropriate, since it alters both the
physical and electromechanical character of the electrode surface. The
problem is further complicated by large spatial and temporal variations in
the baseline corrosion potential (Ecor). An individual site may show a
potential shift of + or - 10 mV within a 5-min period. Shifts as great as 50
mV in a day have been observed, especially on iron. This supports Uhlig's
hypothesis that anodic and cathodic regions are constantly in motion,
establishing a "dynamic" surface equilibrium where the favorability of
specific half-cell reactions changes as the corrosion process proceeds
(Uhlig, 1971). Because anodic and cathodic sites may shift, measurement of
polarization resistance at a particular point on the surface must be
completed quickly before the baseline Ecor (from which the polarization
offset is measured) can change substantially.

Application of a polarizing current generally increases the baseline drift


since it tends to redistribute the anodic and cathodic sites (Ijsseling,
1986). If the baseline is not stable, the polarization offset, and hence the
polarization resistance, cannot be measured accurately. This problem,
combined with the sensitivity of the corrosion surfaces to perturbations
caused by impressed currents, militates against the use of the usual anodic
and cathodic continuous polarization scans (ASTM, 1982). A six-point
polarization offset procedure similar to those suggested by Daniel son (1982)
and Jankowski and Juchniewicz (1980) was used instead. The electrode was
polarized to offsets of 20, 40, and 60 mV in the anodic direction, and -20,
-40, and -60 mV in the cathodic direction. Each polarization step was
completed within a 2-min span, hence the opportunity for baseline drift was
minimized. Because the magnitude of the polarizations was small, the
potential disturbance to the character of the freely corroding surface was
lessened, yet the perturbations were large enough that the impressed current
could be accurately measured. Baseline Ecor was measured both before and
after each polarization to ensure that drift was minimal. The polarization
was repeated if drift produced an error greater than 10 percent of the target
offset.

46
Electrochemical parameters were calculated from the polarization data by
fitting each point to the modified Butler-Volmer equation and adjusting the
estimates of the Tafel slopes to minimize the overall error between observed
and predicted polarization currents. The procedure is summarized in the flow
diagram of Figure 9. In essence, it is a brute force approach that tests
every feasible combination of anodic and cathodic Tafel slope and ultimately
selects the pair that gives the best fit. As such, it lends itself readily
to a computer-based solution; a BASIC language program similar to that of
other researchers (Gerchakov et al., 1981) was developed to speed the
cumbersome calculations. After establishing appropriate search boundaries
and iteration increments, a corrosion current was calculated using the Leroy
Model and an initial selection of Tafel slopes. The Leroy Model (see
Appendix A2) is a statistical approach that predicts the corrosion current
density from a limited set of polarization data and a given set of Tafel
slopes.

Figure 10 is an Evans diagram constructed from polarization data collected on


a copper coupon exposed to the Seattle distribution system for seven days.
The lines of best fit (polarization curves) resulting from the parameter
estimation technique described above are also shown. In this case, the
maximum single error between measured and predicted values is less than 5
percent; it is generally possible to find a fit that falls within an error
range of 5-10 percent. The dashed lines intersecting at Ecor are
extrapolations of the Tafel slopes from the linear region of the polarization
curve. The current density at the point of intersection represents the
current density of the freely corroding surface and can be converted to a
penetration rate by applying Faraday's law.

Appendix A2 documents efforts to validate and assess the precision of the


electrochemical measures. Good correlation of the electrochemical technique
with weight-loss measurements can be readily achieved on the uniformly
corroding surface of copper pipes and the generally uniform corrosion surface
of galvanized iron. However, surfaces that exhibit thick or extremely patchy
scale, such as well-aged black-iron coupons, severely complicate the
measurement of surface potential and polarization resistance. Exposure
periods of as little as 30 days may produce extreme pitting with abundantly
47
Enter Polarization Data

Establish Search Bounds, Increment Steps,


Polarization Offsets and Error Limits

Increment Tafel Slopes and Initialize


Search Program

i
Calculate Corrosion Rate
(Leroy Model)

i
Calculate Predicted Corn Current for each Offset
(Modified Butler-Volmer Eqn.)

fail
i
Error Limit Test at Each Polarization Offset
Measured vs. Predicted Polarization Current

pass

Calculate SSE and Save Values

no
Exceed Iteration Bounds

yes
i
Select Minimum SSE - Report
(Goodness of Fit Test)

Figure 9. Flow Diagram of Tafel Slope Fitting Routine

48
Enlargement

Observed Polarization
Seattle Tap Water Values
dechlorlnated
pH = 7.2
Cond. = 50 uS
1100 _

Current Density CuA/cn;

Figure 10. Evans Diagram of Polarization on a Copper Surface Showing


Theoretical and Observed Values
scaled surfaces that cannot be evaluated in this fashion. Hence,
electrochemical evaluation of corrosion processes was largely restricted to
copper and galvanized-iron coupons.

CORROSION-RATE MEASUREMENTS BY MONITORING OXYGEN DEPLETION

Purpose of Oxygen Depletion Corrosion-Rate Measurements

Iron corrosion rates were also determined by monitoring the rate of oxygen
consumption in the corrosion reaction. Dissolved oxygen acts as the electron
acceptor in the corrosion reaction; for each mole of iron oxidized to ferrous
iron, one-half mole of 02 is consumed. Therefore, the relationship between
iron corrosion rate and oxygen-depletion rate is more straightforward than
electrochemical measurements, which can be affected by scale development and
changes in water chemistry. Electrochemical corrosion-rate measurements are
especially difficult on heavily scaled iron pipe because of uneven corrosion
rates over the metal surface and because of variation in scale thickness and
characteristics.

Weight-loss measurements can only be used to obtain long-term cumulative


corrosion rates. In contrast, dissolved oxygen (DO) depletion-rate
measurements reflect the corrosion rate occurring at the time of the test.
In this way, DO depletion measurements are similar to electrochemical
corrosion-rate measurements.

Oxygen-depletion measurements were used to monitor the effects of water


chemistry changes on the corrosion rate of pipes with preformed iron
corrosion scale. Since DO measurements do not alter the pipe surface,
corrosion rate could be measured before and after water chemistry changes
using the same pipe sections.

Use of the oxygen depletion rate as a corrosion-rate measurement is not new.


In 1924 Whitman used this method to measure the corrosion rate of steel
plates exposed to Boston tap water. In his experiments, dissolved oxygen was
monitored using the Winkler titration method. Much more recently, Sontheimer
et al. (1985) reported that Kuch refined dissolved-oxygen corrosion-rate
50
measurements by employing a membrane-type DO probe in a closed piping loop.
The apparatus used in the research presented here is similar to the system
used by Kuch.

Oxygen-Depletion Piping Loop

A schematic of the pipe loop system used for dissolved-oxygen-depletion


measurements is shown in Figure 11. Pipe sections from the long-term
conditioning piping loops could be installed in the oxygen depletion test
apparatus. This oxygen-depletion test apparatus will henceforth be referred
to as the "mini-loop", so named because of its small size compared to the
conditioning loops.

Except for the pipe sections to be tested, the mini-loop was composed
entirely of glass and plastic materials. The desired water chemistry was
established in a 4-L glass reservoir. Compressed gases such as C02 and N2
could be sparged into the bottom of the reservoir. After the desired water
chemistry was established, the reservoir was isolated from the circulation
loop. Oxygen-depletion corrosion rate measurements are faster when the ratio
of the pipe surface area to the circulation water volume is high. The
circulation loop contained only about 350 ml of water, which allowed for a
measurable DO decrease in a few hours when one to three pipe sections were
installed in the mini-loop.

DO was measured using a PTFE-membrane probe inserted into the circulation


loop and attached to a YSI (Yellow Springs Instruments, Yellow Springs, OH)
oxygen meter. The probe was calibrated by performing triplicate Winkler
titrations of loop water and then adjusting the oxygen meter reading to match
the Winkler titration results. The oxygen meter reading was somewhat
dependent on the flow rate in the circulation loop, so the meter calibration
was adjusted under identical flow conditions as those used during testing.
DO meter readings were continuously recorded using a strip chart. In a few
tests the final DO meter reading was also checked using the Winkler titration
method, and the titration results were found to match the final meter reading
closely.

51
fi) ) l ) )HD>-
Iron Pipe Sections
xA/W*
Heat Exchanger

l~/ Flow
Meter

4 Liter
Reservoir

Pump Magnetic Stirrer

Figure 11. Schematic of Pipe Loop for Measuring Corrosion Rate by Oxygen Depletion
Water was circulated using a nylon-lined centrifugal pump. The flow rate was
kept constant at a flow velocity of 0.2 m/s.

The circulation loop was maintained at a constant temperature by using a heat


exchanger that consisted of glass tubing submerged in a constant-temperature
water bath. Temperature was monitored using the temperature probe attached
to the DO probe. Temperature during all the tests was 21 1* C, and
temperature variation during the course of an individual run was generally
less than 1 C.

Certain precautions were taken to control factors that might interfere with
any of the measurements or interpretations. For example, bacteria present in
the iron scale or growing in the corrosion loop would cause higher DO
depletion rates than the rates for metal oxidation alone. Therefore, the
circulation loop was acid-washed between tests (after removal of pipe test
sections) in order to prevent bacterial growth. If any air leaks were
present, oxygen transfer into loop circulation water would have lowered the
apparent 02 depletion rate. To avoid this, the circulation loop was pressure
tested to detect leaks. Finally, since chlorine can act as an electron
acceptor in the corrosion reaction, water for these tests was dechlorinated.

Trapped bubbles can also bias the oxygen-depletion-rate measurements because


of oxygen transfer from the bubble into the circulation water. It was
frequently difficult and time-consuming to dislodge air bubbles caught in
certain parts of the circulation loop. Use of transparent acrylic tubing
allowed the detection of bubbles trapped in most parts of the loop; however,
bubbles trapped in the pump or in the pipe test section could not be
detected. The circulation loop was presumed bubble-free when no more bubbles
could be dislodged by shaking and tapping the pump and other hidden sections
of the circulation loop.

Analysis of Oxygen-Depletion Corrosion-Rate Measurements

Corrosion rates decreased as the dissolved oxygen decreased. In order that


consistent comparisons could be made, all rate measurements were initiated
and ended when the DO concentrations were 7.85 and 6.85 mg/L, respectively.
53
This range was picked because it was similar to the DO level in the
conditioning loops. In addition, higher D.O. levels increased the tendency
for bubble formation in the circulation loop.

The ratio of oxygen consumed to the quantity of metal oxidized depends on the
oxidation state of the corrosion products. Therefore, an exact conversion
between the oxygen depletion rate and the corrosion rate requires knowledge
of the iron oxidation state in corrosion scales.

In our experiments iron scales were analyzed for the percentage of iron
present in the ferrous state. The results of this analysis were used in the
conversion between oxygen depletion rates and corrosion rates. A sample
conversion is shown below for a scale in which 40 percent of the iron is in
the ferrous state and the remaining 60 percent is in the ferric state. In
this example it is assumed the three iron pipe sections each with 18.5 cm2 of
surface area are installed in the 0.350 liter mini-loop, and a 1 mg/L DO
decrease occurs over 60 min.

(0.4 x 2 e~) + (0.6 x 3 e") = 2.6 e" released per Fe atom oxidized

1 mg 07 56 mg Fe/mmole x 4 e~/mmole Oo
x 0.35 liters x
liter 32 mg 02/mmole 2.6 e"/mmole Fe
= 0.940 mg Fe oxidized.
0.940 mg Fe 525,600 min
Y _ __________ _ Y
1 cm3 Y
1,000
_
mils/2.54
___________ . _
cm
_ __ __________ _
60 min. year 7800 mg Fe 3 x 18.5 cm^
=7.5 mils/year (mpy).

If water quality conditions change, the oxidation state of iron in the scale
may be altered. If the oxidation state of iron in the scale is unknown or if
it is changing, the oxygen depletion rate cannot be correlated exactly with
corrosion rate. Thus, under transient conditions oxygen-depletion corrosion-
rate measurements must be viewed with reservations. Still, oxygen-depletion
corrosion-rate measurements can provide more information than can weight loss
alone about corrosion-rate changes under changing water chemistry conditions.
54
Accuracy and Precision of DO-Depletion Measurements

Dissolved oxygen-depletion corrosion-rate measurements correlated well with


weight-loss data. Table 5 shows a comparison between DO-depletion
measurements and weight-loss measurements obtained on mild-steel pipe exposed
to Seattle tap water (pH 8.0, alkalinity 20 mg/L as CaCOj). One problem with
this comparison is that weight-loss measurements are cumulative, and
dissolved-oxygen-depletion measurements represent the corrosion rate at a
point in time. The corrosion rate of a coupon after 600 days based on DO-
depletion is lower than the average corrosion rate of the same coupon
determined by weight loss. Since corrosion rates in this water type were
initially very high, it is not surprising that the long-term corrosion rates
based on cumulative weight-loss analysis, which include this high initial
weight loss, exceed the short-term corrosion rate occurring after long
exposure.

In an attempt to overcome this problem, incremental weight-loss corrosion


rates were calculated. The average weight loss at 35 days was subtracted
from the average weight loss at 234 days. The difference is assumed to be
the weight loss between 35 days and 234 days and is used to calculate .the
average corrosion rate during this time period. As Table 5 shows, the
incremental weight-loss measurements agree more closely with the dissolved-
oxygen-depletion corrosion-rate measurements. The problem with calculation
of incremental weight loss is that data scatter can lead to erratic results.
Because the incremental method relies on the calculation of differences
between the weight loss of two different pipe sections, it is more
susceptible than the other corrosion-rate measurement methods to errors
arising from the variable corrosion behavior of pipe sections.

The weight-loss standard deviation represents the weight-loss variation


between multiple pipe coupons with the same exposure history. By contrast,
the DO-depletion standard deviation represents the variation in repeat
measurements performed on the same group of pipe sections.

55
TABLE 5
Comparison of Corrosion-Rate Measurement Methods

Cumulative Weight-Loss Method (Long-Term)


Days Corrosion Rate Percent Std. Number of
Exposure in mpy Deviation Data Points
35 days 17.5 + 0.3 2% 2 pts.
234 days 8.0 + 0.4 5% 3 pts.
472 days 5.8 + 0.1 1% 2 pts.
699 days 3 5.0 ~ 1 pt.

Incremental Weight Loss Method (Intermediate-Term)


Days Corrosion Rate
Exposure in mpy
35 days 17.5
35-234 days 6.3
234-472 days 3.7
472-699 days 3 3.3

Dissolved-Oxygen-Depletion Corrosion-Rate Methods (Short-Term)


Days Corrosion Rate Percent Std. Number of
Exposure in mpy Deviation Data Points
47 days 11.6 + 0.7 6% 4 pts.
62 days 10.2 + 0.9 9% 3 pts.
117 days 9.7 + 1.2 12% 3 pts.
645 days 3 3.5 + 1.9 53% 4 pts.
aThe 699 day weight-loss measurement was performed on the same coupon used
for the 645 day DO-depletion-rate measurement. All of the other
measurements represent different sets of pipe sections.

56
The precision of DO-depletion corrosion-rate measurements appeared to depend
on the rate of oxygen depletion. In our experiments the rate of oxygen
depletion was fast when corrosion rates were high and when a greater number
of pipe sections were used during testing. As shown in Table 5, duplicate
measurements at high corrosion rates were fairly precise. The first three
DO-depletion corrosion-rate measurements shown in the Table were made when DO
depletion was rapid (1 mg/L decrease in about one hour). In these tests,
measurements were made with three pipe sections installed in the mini-loop.
Measurements made on the 645-day-old coupon, however, were much less precise.
For this coupon a 1 mg/L DO decrease took several hours, because the
corrosion rate was slow and because only one pipe section with this exposure
history was available for testing. Pipe sections with long periods of pre-
exposure were the least likely to have duplicates with the same exposure
history.

The dependence of measurement precision on the rate of oxygen depletion could


be related to oxygen meter drift. When tests take several hours, oxygen
meter drift could add error to measurements. Using a smaller DO range for
rate measurements shortened the test time but did not improve precision. The
failure of a smaller DO range to improve precision may be because of short-
term fluctuations or "noise" in meter readings. On a short-term cycle the
meter readings drifted up and down by about + 0.02 mg/L. These fluctuations
were not related to the meter calibration, since the reading fluctuated on
both sides of the calibration setting. This fluctuation is only 2 percent of
the 1 mg/L range used for measurements but is obviously more important as the
DO range used for measurements is decreased.

ANALYSIS OF CORROSION SCALE

Purpose of Corrosion-Scale Analysis

Corrosion scales were analyzed for their chemical composition, crystalline


structure, and physical appearance. The purpose of this analysis was to
identify the chemical compounds present in the scale and to assess structural
and compositional differences between scales formed under various water
chemistry conditions. Scale characteristics were correlated with the
57
corrosion rate of the scale-covered surfaces to determine which scale types
lead to lower corrosion rates. One goal of scale analysis was to correlate
the solid compounds identified in the scales with thermodynamic expectations.
As noted earlier, the composition of the bulk solution may differ from that
near the pipe wall, where the scales form. Nevertheless, it seems that
scales that are thermodynamically stable with respect to the bulk water are
more likely to form protective coatings; unstable scales are more likely to
dissolve into the bulk solution.

Visual examination of the scale reveals scale uniformity and the extent of
metal surface coverage. The presence of tubercles generally indicates
pitting corrosion as opposed to uniform metal loss. Large tubercles decrease
pipe carrying capacity and increase pumping costs.

The thickness, hardness, and adherency of the scale to the metal surface can
be judged as the scale is physically removed from the pipe surface. These
scale characteristics may indicate the degree of corrosion protection
afforded by the scale. These features can also be used to predict the
likelihood of metal release into the flowing water, since thick loose scales
can be easily sloughed off by changing water chemistry and flow conditions.

Different scale layers often exhibit different characteristics such as color,


hardness, porosity, and apparent crystalline structure. Each distinct scale
layer may be composed of dissimilar chemical compounds or may be different
manifestations of the same chemical compound. The identification of chemical
compounds in a scale is therefore only suggested by the scale's physical
appearance and must be confirmed through chemical or x-ray diffraction
analysis.

X-Rav Analysis of Corrosion Scales

Powdered scales were analyzed using x-ray diffraction analysis. X-ray


diffraction can identify the crystalline components present in the scale.
X-ray diffraction has two limitations: it cannot identify amorphous
compounds, and it gives only qualitative, not quantitative, results.

58
In this analysis, scale samples were ground to a fine powder and then sieved
through a 200-mesh screen to eliminate inadequately ground particles. The
powder was placed on an aluminum slide and exposed in a Phillips x-ray unit,
model 3005/230, with goniometer model 42202. The x-ray source was a copper
vacuum tube aligned to the K-alpha wavelength.

Chemical Analysis of Corrosion Scales

Corrosion scales were removed from the pipe surface and analyzed for total
metal (iron, zinc, or copper), ortho-phosphate, and carbonate content. The
scales from the iron coupons were also analyzed for ferrous iron content.

Scale Oxidation

Iron scale components generally oxidize to ferric hydroxide when dried in the
atmosphere (Sontheimer et al., 1985). Therefore, as the scale dries its
composition changes. In order to characterize scale composition as it exists
on the pipe surface, scales were analyzed before they were dry. After pipe
sections were removed from the conditioning pipe loops, the wet scale was
immediately scraped off and dissolved in a dilute hydrochloric acid (HC1)
solution. After dissolution, wet scales were analyzed for ferrous iron,
total iron, and carbonate. Some dried scales were also analyzed. The same
analysis methods were used for the dried scale as for the wet scale, except
that the dried scale was weighed before dissolution. The analysis of dried
scale was used to determine which scale components were most resistant to
oxidation.

Ferrous Iron Analysis

Ferrous iron was analyzed to evaluate the ratio of ferrous iron to total
iron. Iron scale was allowed to dissolve overnight in a mixture of
hydrochloric acid and deoxygenated water. After dissolution, the iron
solution was diluted using deoxygenated water to obtain an approximate
concentration of 0.25 mg/L ferrous iron. The solution was then analyzed in a
light spectrophotometer using 1,10-phenanthroline as the coloring agent.
Hach reagents and a Milton Roy Company Spectronic 1001 spectrophotometer were
59
used in both the ferrous iron and total iron tests. The 1,10-phenanthroline
will only react with ferrous iron. To analyze for total iron, the 1,10-
phenanthroline is added along with sodium hydrosulfite, a reducing agent
which reduces all the iron in the solution to ferrous iron. Solutions were
first analyzed for ferrous iron and then diluted again (if necessary) and
analyzed for total iron.

This analysis cannot detect small amounts of ferrous iron in the presence of
large amounts of ferric iron. The 1,10-phenanthroline will cause some ferric
iron reduction even if no other reducing agent is added. The analysis of
iron standards prepared entirely with ferric iron indicated the presence of
about 2 percent ferrous iron. Therefore, the analysis of scales containing 2
percent or less ferrous iron were not reliable.

Another problem with the ferrous iron analysis was that some ferrous iron
might have oxidized to ferric iron during dissolution and dilution, even
though the pH was kept low and deoxygenated water was used for dilutions.
Iron solutions were prepared using known amounts of ferric ammonium sulfate
and ferrous ammonium sulfate, and these controls were analyzed using the same
procedures as were used for scale samples. One ferrous/ferric control was
run during each ferrous iron scale analysis. The ferrous/ferric control
solutions lost an average of 12 percent of their ferrous iron during
preparation and analysis. This 12 percent average loss is based on nine
ferrous/ferric controls; ferrous loss ranged between 9 percent and 14
percent. To compensate for this loss, ferrous iron scale sample results were
increased by 12 percent as well. For example, a result of 32 percent ferrous
iron is reported as 36 percent (1.12 x 32 percent) ferrous iron to compensate
for the oxidation suspected to occur during analysis.

Error in the ferrous iron analysis and in all the other iron scale analyses
could be caused by scraping off an unrepresentative scale sample. Soft scale
components were removed much more easily than the hard tenacious scale layers
that formed on some of the iron pipe surfaces. On the other hand, very wet
soft scales often clung stubbornly to the pipe surface and scraping utensil,
and it was difficult to collect the resulting mass for analysis. As reported
by Sontheimer et al. (1985), Kolle and Kuch used dilute acid to remove iron
60
scale for analysis. The dilute acid might have been preferable for
completely removing wet soft scale components from pipe surfaces. However,
the dilute acid removal method was not used in this study in order to be
consistent with the scale removal methods used for weight-loss analysis, and
also because of the potential for dissolving iron metal along with the scale.

Total-Metal Analysis of Dried Scales

Dried scales were analyzed for their metal content. The scale was removed
from the pipe surface, air dried, and then ground into a powder. Next, the
scales were dried in a 105 C oven for 4 hr, cooled in a desiccator, and
finally weighed into a dilute hydrochloric acid solution. These solutions
were analyzed for total metal by atomic absorption spectrophotometry.

Carbonate Analysis

The carbonate analysis used in this research was developed by soil scientists
for determining the carbonate mineral content of soils (Bundy and Bremner,
1972). In this method, a potassium hydroxide (KOH) solution is used to
collect the C02 that evolves when carbonate minerals are dissolved in acid.

Corrosion scale was placed in a flask along with a potassium hydroxide trap.
The KOH trap consisted of a beaker containing exactly 5 ml of KOH. The KOH
was freshly prepared using ^-free water and protected from atmospheric
contamination. This beaker was suspended inside the flask, and the top of
the flask was sealed with a rubber stopper. A small-diameter glass tube
inserted into the rubber stopper was sealed with a soft rubber septum. Next,
a hydrochloric acid solution was injected through the rubber septum using a
syringe. The acid solution flowed down the glass tubing and covered the
scale at the bottom of the flask. As the scale dissolved overnight in the
acid solution, the carbonate present in the scale was evolved as C02 gas and
was collected in the KOH trap. Although it was assumed that all the
carbonate scale dissolved, it was not possible to test this assumption
independently. The next day, the KOH solution was removed from the flask,
and its carbonate content was determined by analysis in a total-carbon
analyzer. The carbon concentration in the KOH was proportional to the
61
carbonate in the scale. The acidic iron solution in the flask was then
analyzed for total metal. The results obtained were expressed as the mole
ratio of carbonate ion to metal.

Contamination with atmospheric C02 is an important source of error in this


technique. Analytical blanks were run to correct for atmospheric C02
contamination occurring during the analysis procedure. Some atmospheric
contamination was unavoidable during transfer of the KOH into the flask and
during KOH analysis for total-carbon content. Contamination limited the
sensitivity of the technique for detecting very low carbonate concentrations.
This limitation could be partially counteracted by using large quantities of
scale. Often, however, scale quantities of a particular scale type were
limited. In addition, obtaining complete scale dissolution was more
difficult when large quantities of scale were analyzed.

Phosphate

The ortho-phosphate content of scales was analyzed by dissolving the scales


from mild steel or galvanized steel coupons in acid and the scale from copper
coupons in an ammonia solution, and analyzing the resulting solution
spectrophotometrically according to the method of Strickland and Parsons
(1972).

Copper Composition of Scales

The weight loss due to corrosion of copper coupons could be evaluated with
acceptable accuracy and precision by weighing the coupons before and after
the polishing process described previously. However, it was difficult to
collect and characterize all the powdery scale that was released during
polishing. Also, acid dissolution of the scale directly from the coupon led
to dissolution of unacceptable amounts of metallic copper. Therefore, an
alternative approach was used for evaluating the oxidation state and
composition of this scale.

The approach used followed the method of Budesinsky (1977) for the
differentiation of oxidized and metallic copper in ores. In this method, the
62
oxidized copper is selectively dissolved by exposure to a solution comprising
36 g ammonium chloride plus 280 ml ammonia hydroxide, diluted to 1.0 L with
distilled water. Budesinsky added sodium pyrophosphate to his etching
solution to mask interferences from iron (III). Since the copper coupons in
this study were not expected to have such interferences, and since it was
desirable to analyze some of the scales for their phosphorus content,
pyrophosphate was not included in these solutions.

During the etch procedure, the interior of a copper coupon is filled with the
ammonia etch solution (approximately 3 ml). After a predetermined contact
period, the etch solution is drained into a volumetric flask. The procedure
is then repeated, refilling the coupon with fresh etch solution and, after an
equivalent contact period, adding the solution to the first volume collected.
The collected solutions are then diluted to a known volume and analyzed for
copper and, in some cases, ortho-phosphate.

The etching contact period varied from 10 to 30 min per solution. Two 10-min
etches of a clean coupon caused a 6.0 mg weight loss, of which 5.0 mg was Cu.
Two 30-min etches caused a weight loss of 11.8 mg, of which 10.3 mg was Cu.
It is not clear whether these weight losses represented dissolution of
oxidized Cu which was on the surface of the clean coupons prior to the test,
or whether the etching process led to some oxidation. Weight loss from
coupons that had been exposed to loop water was always at least three times
that of the corresponding clean coupon, except for one coupon from the 5 mg/L
P loop that lost a total of only 2.3 mg in two 10-min etches. It was
concluded that the procedure can provide a reasonable qualitative and
semiquantitative estimate of the copper content of collected scales.

63
CORROSION OF MILD-STEEL PIPE

CORROSION RATES BASED ON WEIGHT LOSS OF MILD-STEEL PIPE

Overall Weight-Loss Comparison

Iron weight-loss data are shown as a function of exposure duration for the
various water types in Figures 12 through 17. The corresponding cumulative
average corrosion rates are shown in Figures 18 through 21. In all of these
figures, each data point represents one mild-steel pipe section. Multiple
data points on the same day represent analysis of several individual coupons.
Average corrosion rates were calculated by dividing weight loss by days of
exposure and then converting the results into mils per year (mpy) (1 mil =
1/1000 in.). Thus the rates shown were averaged over the total exposure
period. If the corrosion rate of a pipe section changed during the exposure
period, then the actual corrosion rate occurring at the end of the exposure
period would be different from the cumulative corrosion rate shown.

In the pH 6.0 and the pH 8.0 waters with 0 and 1 mg/L o-P04 -P, weight-loss
data were obtained for pipe sections with almost 700 days of exposure. Few
other corrosion studies have obtained such long-term weight-loss data under
controlled water quality conditions. The oldest pH 8.0/5 mg/L P coupons had
560 days exposure; the oldest pH 7.2/high-carbonate pipe section had 460 days
exposure; the oldest pH 8.4/low-carbonate pipe section had 255 days exposure;
and the oldest pH 8.0/low-carbonate pipe section had 180 days exposure.

A six-month test period has frequently been used in previous potable water
corrosion studies. Table 6 shows average corrosion rates for pipe sections
with approximately six months exposure.

*1 mil/yr = 0.0254 mm/yr; for the coupons used in these experiments, 1


mil/yr = 5.43 mg/cnr-day = 54.3 g/nr-day.

65
5000

4000-

3000-

Weight Loss o
(mg)
o
2000-
o
o

1000-

200 400 600 800


Exposure (days)

Figure 12. Weight Loss of Iron Pipes Exposed to Seattle Tap Water,
pH 8.0 (Water 2 from Table 2)
5000

4000-

3000-
o
Weight Loss
(mg) 8

2000- o
ON o

8
1000-
o

200 400 600 800


Exposure (days)

Figure 13. Weight Loss of Mild Steel Pipe Sections Exposed to


pH 6 Water (Water 1 from Table 2)
5000

4000-

o
o

3000-

Weight Loss
(mg)

2000-

1000-

200 400 600 800


Exposure (days)

Figure 14. Weight Loss of Iron Pipes Exposed to the Low Carbonate
Water at pH 8.0 (Water 4 from Table 2)
5000

4000-

3000-

Weight Loss
(mg)

2000-1
vO

1000-

200 400 600 800


Exposure (days)

Figure 15. Weight Loss of Iron Pipe Exposed to the Low Carbonate
Water at pH 8.4 (Water 3 from Table 2)
5000

4000-

3000-
Weight Loss
(mg)

2000-

1000-

200 400 600 800


Exposure (days)

Figure 16. Weight Loss of Iron Pipe Exposed to the pH 7.2,


High Carbonate Water (Water 5 from Table 2)
5000

4000-

o pH= 6.0 (Water 1)


3000-
pH= 8.0 (Water 2)
Cumulative
Weight Loss pH= 8.0 PO4= 1 mg/l (Water 6)
(mg)
2000 o pH= 8.0 PO4= 5 mg/l (Water?)


I
B
1000-
B

200 400 600 800


Exposure (days)

Figure 17. Weight Loss of Mild Steel Coupons in PO4 - Inhibited and
Non-Inhibited Waters.
20-

18- o
o
16-
o o
o
14-

12-

Corrosion
Rate (mpy) 10

8- o
8
8 o
6-

4-

2-

0
0 200 400 600 800
Exposure (days)

Figure 18. Change in Corrosion Rate With Length of Exposure


to pH 8.0 Water (Water 2)
20-

18-

16-

14-

o
12-

Corrosion 8 ^
Rate (mpy) 10 '
o ft 8
0 O
8- o
o o o 0 o
6- 0 0
o o
o
0
4-

2-

n-
200 400 600 800
Exposure (days)

Figure 19. Change in Corrosion Rate With Length of Exposure


to pH 6.0 Water (Water 1)
22 '

20- i :

18-
o
16-

14-
0
o 0
12 *
Corrosion
o pH 8.4. Alk. 20 mg/l (Water 3)
Rate (mpy)
10- pH 8.0. Alk. 13 mg/l (Water 4)

8-

6-

4-

2-

A -

50 100 150 200 250 300


Exposure (days)

Figure 20. Change in Corrosion Rate With Length of Exposure


to Low Carbonate Waters.
15

10- o
o

Corrosion
Rate (mpy)

5-

8 o
o
o o
o
8
100 200 300 400 500
Exposure (days)

Figure 21. Change in Corrosion Rate With Length of Exposure to High


Carbonate Water (Water 5)
TABLE 6
Over Six Months Exposure
ates Ove
Cumulative Corrosion Rates
(Rates calculated by weight loss method)

Number of
Water Type Corrosion Rate Data Points

Low Carbonate, pH 8.4 19.2 0.2 mpy 2 points


Low Carbonate, pH 8.0 13.3 mpy 1 point
Seattle Tap Water, pH 6.0 10.0 1.2 mpy 8 points
Seattle Tap Water, pH 8.0 8.6 - 2.2 mpy 3 points
High Carbonate, pH 7.2 2.0 - 0.9 mpy 4 points
5 ppm o-P04 -P, pH 8.0 1.9 - 0.2 mpy 2 points

Corrosion rates were calculated by averaging all data points with an


exposure period of 182 days + 20 days. The only exception is the pH 8.4
water where, due to limited data, weight loss from a 213-day-old pipe section
was used. No coupons from the 1 mg/L o-P04 -loop were analyzed after
appropriate exposure times for inclusion in the table.

The water quality characteristics of all the waters are reproduced in


Table 7. The pH 7.2 high-carbonate water and the 5 mg/L o-P04 -P water were
by far the least corrosive. The 1 mg/L o-P04 -P Seattle tap water at pH 8,0
(based on Figure 17), and the pH 6.0 water produced the next lowest corrosion
rates; corrosion rates in these two waters were similar.

The pH 8.4 low-carbonate water appeared more corrosive than the pH 8.0
Seattle tap water and the pH 6.0 water; however, limited data were collected
for this water type. The pH 8.0 low-carbonate water was the most corrosive,
producing corrosion rates more than double those in Seattle tap water at the
same pH. Corrosion in the pH 8.0 loop slowed considerably after 9 months.

Correlation of Water Quality Variable with Corrosion Rates

Table 7 shows water quality parameters and the average corrosion rates for
pipe sections with six months of exposure in the loops with no phosphate
inhibitor.
76
TABLE 7
Water Quality Variables and Corrosion Rates

Water Corrosion^ Buffer ,


Type Rate (mpy) pH Alk.t Intensity? SI

** -1.1
pH 8.4, low Cj 19.2 + 0.2 8.0 13 0.016
pH 8.0, low Cj 13.3 8.4 20 0.024 -0.5
pH 6.0 10.0 + 1.2 6.0 10 0.310 -3.2
pH 8.0 8.6 + 2.2 8.0 20 0.025 -0.9
pH 8.0, high CT 2.0 0.9 7.2 400 96.5 -0.4

Water Corrosion Larson


Type Rate (mpy) IndexfT
Log(OH') Log(C03 2 ") :H:

pH 8.4, low CT 19.2 + 0.2 1.2 -6.0 -5.9


pH 8.0, low CT 13.3 0.3 -5.6 -5.3
pH 6.0 10.0 + 1.2 2.5 -8.0 -8.5
pH 8.0 8.6 + 2.2 0.3 -6.0 -5.7
pH 8.0, high C 2.0 + 0.9 0.01 -6.8 -5.2

Averaged over six months of exposure.


|mg/L as CaCOo
Tmeq/L/pH unit
^jLangelier Saturation Index
CT = Total Dissolved Inorganic Carbon
tt[(cr) + (S042 -)]/(HC03 -)
rrDissolved Carbonate Concentration

Buffer Intensity

Of the several water quality parameters considered, buffer intensity had the
strongest correlation (an inverse one) with long-term corrosion rates. The
high-carbonate water had a much higher buffer intensity and a much lower
corrosion rate than all of the other water types. The pH 8.0 low-carbonate
water had the lowest buffer intensity and the highest corrosion rates. The
pH 8.4 low-carbonate water had a slightly smaller buffer intensity and a
greater corrosion rate than the pH 8.0 Seattle tap water. The higher
77
corrosion rate in the pH 8.4 water compared to the pH 8.0 water is consistent
with the minimum in buffer intensity at pH 8.3. Hedberg and Johansson (1984)
also found that iron corrosion rates increased when the pH was raised above
8.0.

The only exception to the inverse correlation between buffer intensity and
corrosion rates was the pH 6.0 water, which had a higher buffer intensity
than pH 8.0 Seattle tap water but was slightly more corrosive at long
exposure periods. This exception might have been due to the high chloride
concentration in the pH 6.0 water. The high chloride concentration of this
water is reflected in Table 7 by its high Larson Index. Chlorides could also
be partially responsible for the high corrosion rate in the pH 8.0 low-
carbonate water.

The fact that buffer intensity did not predict the relative long-term
corrosivity of pH 6.0 water compared to pH 8.0 water points to the complexity
of the relationship between water chemistry and iron corrosion behavior. It
is difficult to find any one water quality parameter that will always succeed
in predicting relative corrosivity. In general, however, these results
showed a significant inverse correlation between buffer intensity and
corrosivity. This correlation agrees with the research results of Stumm,
Sontheimer, Hedberg, and others.

Alkalinity

Alkalinity correlated weakly with corrosion rates. The pH 7.2 high-carbonate


water had the highest alkalinity and the lowest corrosion rates of all the
water types. The other four lower alkalinity waters were significantly more
corrosive. Among the low-alkalinity water types, however, alkalinity did not
correlate well with corrosion rates. The pH 8.4 low-carbonate water had the
same alkalinity as the pH 8.0 Seattle tap water, but was much more corrosive.
The pH 6.0 water had the lowest alkalinity of all the water types, but was
not the most corrosive. For these waters, buffer intensity was a better
predictor of corrosivity than was alkalinity.

78
The Langelier Saturation Index

The Langelier Saturation Index (SI) proved to be a very poor corrosivity


indicator for the low-alkalinity waters. Raising pH from 8.0 to 8.4 at
constant alkalinity increased the SI, but it also increased corrosion. The
pH 6.0 water had the lowest SI, but also a lower corrosion rate than the
high-pH and low-carbonate waters.

pH

There was also no apparent correlation between pH and corrosion rates. The
pH 8.0 low-carbonate water had the same pH as Seattle tap water, but was more
corrosive. The pH 6.0 water had the lowest pH, but was definitely not the
most corrosive. The pH 6.0 water also had the highest chloride concentration
and the highest Larson index; however, it still was less corrosive than the
higher pH, low-carbonate waters. Perhaps the higher buffer intensity gave
some protection against the aggressive chloride ion and the low alkalinity in
the pH 6.0 water.

Determining the effect of pH on iron corrosion rates is complicated because


the effect appears to change with duration of exposure and because of
possible interferences from chloride and sulfate when acid is used to adjust
the pH. Results from previous research are also conflicting.

Many previous researchers have found pH 8.0 water to be more corrosive than
pH 6.0 water. Larson and Skold (1957) and Stumm (1960) both found lower
corrosion rates in waters near pH 6.0 compared with waters at pH 8.0 and
above. To lower the pH, those studies used CC^, which does not increase the
chloride concentration or lower the alkalinity. In the research presented
here, the solutions were acidified with HC1, which raises the chloride
concentration and lowers the alkalinity. The corrosive attack of chloride
ions is a plausible explanation for the difference between this study and the
two previous studies, because chlorides have been found to promote iron
corrosion. On the other hand, both of the previous studies lasted only 100
days. At 100 days exposure, corrosion rates were comparable in the pH 6.0
and the pH 8.0 waters in the research presented here (see Figures 12 and 13).
79
If Stumm or Larson and Skold had continued their studies for a longer time,
the results of their studies may have agreed more closely with the research
presented here.

Eliassen et al (1956) also studied the effect of pH on iron corrosion. They


found more pronounced pitting corrosion at pH 6.5 compared to pH 8.0,
although overall weight loss by corrosion was the same. In that study, 30-
to-40 day flow tests were conducted using pipe specimens installed in a
circulation pipe loop. They used fairly low alkalinity water (57 mg/L as
CaC03 ) and adjusted the pH to 6.5 using hydrochloric acid. In one test using
a flow velocity of 0.6 m/s, corrosion rates at both pH 6.5 and pH 8 were
11 mpy. In another test using a flow velocity of 0.3 m/s, the corrosion rate
at pH 6.0 was 18 mpy, while at pH 8.0 the corrosion rate was again 11 mpy.
Eliassen's results contrast with this study, since it was found here that
30-day corrosion rates were lower in the pH 6.0 water than in the pH 8.0
water.

In summary, of the water quality variables studied, buffer intensity showed


the best qualitative correlation with the corrosion rates of mild-steel pipe.
Alkalinity correlated weakly with corrosion rate, in that the water with an
alkalinity of 400 mg/L as CaCOg produced much lower corrosion rates than
waters with alkalinities of 20 mg/L as CaC03 and less. In contrast, pH did
not have any apparent correlation with corrosion rates. The Langelier
Saturation Index, although commonly used to make corrosivity predictions,
also did not correlate well with corrosion rates.

Phosphate

The weight-loss data for the loops containing 1 mg/L P04 -P have quite a bit
of scatter. In general, they suggest that addition of this much ortho-
phosphate had no effect on iron corrosion rates when compared to the PO^-free
control water at pH 8.0. The addition of 5 mg/L P04 -P did reduce the
corrosion rate by about two-thirds compared to the control. While
concentrations as high as 5 mg/L may be added to new pipe systems for a few
weeks to generate a protective scale layer quickly, common practice is to add
80
no more than about 1 mg/L PO^-P as a maintenance dose. The reason for this
is to avoid a dramatic increase in the cost and the P load on sewage
treatment plants receiving wastewater from an affected community. Indeed, in
localities such as Seattle with open water supply reservoirs, even 1 mg/L
PO^-P would cause algal growth problems in the water distribution network.
Because an unrealistic maintenance dose of PO^ seemed to be required to
significantly reduce iron corrosion rates in the systems in this study, the
discussion of iron corrosion and its inhibition will emphasize the results
from the phosphate-free loops.

Initial Corrosion Rates

Initial corrosion rates are not always indicative of long-term corrosion


behavior; however, many other investigators have used short-term corrosion
tests in their research. The initial corrosion rates found in this study are
summarized in Table 8.

TABLE 8

Initial Corrosion Rates Found on Mild-Steel Pipe


(Calculated by Cumulative Weight-Loss Method)

Water Type Exposure Corrosion^Rate


(days) (mpy)
Seattle Tap Water, pH 6.0 36 5.24
36 4.75
Seattle Tap Water, pH 8.0 35 17.30
35 17.75
Low Carbonate, pH 8.0 42 18.86
High Carbonate, pH 7.2 47 11.31

*mils per year

81
Table 8 shows that reliance on short-term corrosion tests can lead to errors
when the results are used to predict long-term corrosion behavior. Until 60
days exposure, the pH 7.2 high-carbonate water appeared more corrosive than
the pH 6.0 water; however, after long-term exposures, the low-pH water was
clearly more corrosive. Pipe sections exposed to pH 8.0 Seattle tap water
also had much higher initial corrosion rates than those exposed to pH 6.0
water. Later, corrosion rates in the pH 8.0 water appeared to be the same as
in the pH 6.0 water.

These results agree with those of both Larson (1957) and Stumm (1960), who
found that initial corrosion rates did not predict corrosivity at longer
exposures. In those studies, corrosion rates measured after a week were not
well-correlated with corrosion rates measured after several months of
exposure.

The corrosion rates in the pH 6.0 water were very low initially (less than 60
days exposure), but then seemed to increase. After about 60 days exposure,
corrosion rates stabilized and then slowly declined. A plausible explanation
for this trend is that corrosion initiation sites form more slowly in the pH
6.0 water than in the higher-pH waters, but that once sites are initiated,
corrosion increases rapidly until a protective scale develops, at which time
corrosion rates decline once again.

Comparison of Long-term Corrosion Rates in pH 6.0 and pH 8.0 Seattle Tap


Water

Seattle tap water at pH 8.0 caused a weight loss similar to pH 6.0 Seattle
tap water until about a year of exposure. Thereafter, almost all of the pH
6.0 pipe sections had higher weight loss than the pH 8.0 pipe sections with
similar exposure periods. Table 9 shows a comparison of corrosion rates in
the pH 6.0 and pH 8.0 waters after a year or more of exposure.

82
Corrosion rates in the pH 8.0 water declined steadily during the entire test
period. No similar consistent decline was apparent for the iron pipe
sections exposed to the pH 6.0 water.

TABLE 9
Long-term Corrosion Rates in pH 6.0 and pH 8.0 Seattle Tap Water
(Calculated by Cumulative Weight-Loss Method)

Water Type Exposure Corrosion^Rate


(days) (mpy)

Seattle Tap Water 425 6.45


pH 6.0 483 6.35
528 6.89
528 8.14
680 7.21
Seattle Tap Water 382 7.04
pH 8.0 382 6.76
416 6.42
432 6.08
472 5.87
472 5.78
699 5.00

mils per year

Comparison with Previous Corrosion Research Conducted Using Seattle Tap Water

Previous corrosion research to evaluate the effectiveness of corrosion-


control water treatments has been conducted using Seattle source waters. In
1978, pilot-plant corrosion studies were conducted by Kennedy Engineers for
the City of Seattle Water Department. Kennedy Engineers (1978) used nine-
month flow tests to determine the effect of a lime/soda ash treatment on the
corrosion rates of black-iron pipe. The lime/soda ash treatment was
evaluated in Seattle's two main source waters, the Cedar River and the Tolt
River. In that study, the final pH of the treated waters was about 9.0.

83
Hoyt et al. (1982) and Herrera and Hoyt (1984) also evaluated the effect of
corrosion-control water treatment in Seattle source waters. They used
six-month-long pilot-plant studies and slightly different chemical treatment
methods than those used in the Kennedy study. In the 1982 study, Cedar river
water was dosed with lime. In the 1984 study, Tolt river water was treated
by two methods: one test was conducted by adding lime and soda ash, and
another test was conducted with lime and sodium bicarbonate. In the Hoyt et
al. and Herrera and Hoyt studies, the final pH of the treated waters was
about 8.0.

Water quality characteristics and iron corrosion rates from the corrosion-
control pilot-plant studies are shown in Table 10. As shown in the table,
the corrosion-control treatment methods used in these studies produced only
minor changes in iron corrosion rates. In the Kennedy studies, treatment
slightly decreased mild-steel corrosion rates in both the Tolt and Cedar
river waters. In the Hoyt and Herrera studies, treatment increased mild-
steel corrosion rates slightly in the Tolt river water and decreased them
slightly in the Cedar river water.

The reason this type of corrosion treatment had little effect on mild-steel
corrosion rates may have been that the water chemistry changes produced by
treatment had opposing effects on iron corrosion rates. In all of the
studies, alkalinity and hardness were added as a corrosion-control treatment.
The treatment raised the pH and lowered the buffer intensity of the test
waters. According to the siderite model, and in concordance with the results
obtained in this study, the lower buffer intensity could have caused higher
corrosion rates. On the other hand, these treatment methods also resulted in
increases in alkalinity and calcium ion concentrations and decreases in the
Larson Index. Since increases in alkalinity and calcium ion concentrations
and decreases in the Larson Index generally lower iron corrosion rates, these
changes may have partially compensated for the lower buffer intensity. In
this study, corrosion rates in the pH 8.0 Seattle tap water were also not
very different from the corrosion rates of iron pipe exposed to the pH 6.0
water.

84
TABLE 10
Seattle Corrosion-Control Pilot-Plant Studies
Average Corrosion Rates Obtained on Mild-Steel Pipe

Ctdar River Hater for 9 months - Kennedy


Alk.
<lfl Buffer Larsoo
Treatment PH g/L Intensity* Indext
None 7 19 7.6 0.15 0.4
Lime 1 Soda Ash 9 34 9.0 0.09 0.2

Corrosion Rate Corrosion Rate


Treatment 6 months 9 months
None 6.5 mpy 5.6 mpy
Lime & Soda Ash 6.2 mpy 4.4 mpy

Cedar River Hater for 6 months - Hoyt (1982) ,


Alk. Hardness Buffer Corrosion
Treatment pH Intensity* Rate
None 7.2 16 19 0.08 6.7 mpy
Lime 7.9 19 20 0.03 6.3 mpy

Tolt River Hater for 9 months - Kennedy <,,7|,


Alk. Buffer Larsop
Treatment pH mg/L Intensity* Indext
None 6.2 1 7.6 0.03 12
Lime & Soda Ash 8.9 10 9.0 0.03 0.6

Corrosion Rate Corrosion Rate


Treatment 6 months ' 9 months
None 9.1 mpy 8.7 mpy
Lime & Soda Ash 8.0 mpy 7.7 mpy

Tolt River Hater for 6 months - Herrera and Hoyt (1984)


Alk. L1me Buffer 4 Larsop
Treatment PH Dose5 Intensity* Indext
None 6 2 0.06 6.3 mpy
Lime & Soda Ash 7.85 14.5 1.5 0.02 8.3 mpy
Lime & Bicarbonate 7.95 21 4.5 0.02 7.5 mpy

'Alkalinity & Hardness: mg/L as CaCo3


tfiuffer Intensity: meoyL/pH unit
fLarson Index: [(CT) 4 (S04 '*)]/(HC03 ')
L1me Dose: mg/L CaO
Results obtained by exposing mild-steel pipe to water In a continuous-flow
piping loop.

85
IRON CORROSION SCALE ANALYSIS

Physical Characteristics of Iron Corrosion Scales

All of the low-alkalinity waters produced iron scales that were similar in
physical appearance and very different from the scale produced in the pH 7.2
high-carbonate water.

The low-alkalinity waters all produced thick, loose scales. The exterior
scale layer was a dark orange-brown crust. This outer layer also contained
flecks of yellow-colored scale. The crust was rigid but could easily be
broken through to reveal a soft porous interior. The interior was composed
of fine granular material that was black or greenish-black. This scale could
be removed by gentle scraping. On pipe sections exposed to pH 6.0 and pH 8.0
Seattle tap water, well-developed scale was approximately 1/8 in. thick. In
the low-carbonate, high-pH waters, scales were slightly thicker. In all of
the low-alkalinity waters, scale thickness was very uneven; scales consisted
of tubercles (mounds of scale) overlying wide shallow pits in the metal
surface.

In all water types, corrosion generally occurred first near metal stress
points such as the machine-cut edges and the butt weld. Machining and
welding stress the metal's crystal lattice and cause these areas to be more
susceptible to corrosive attack (Evans, 1960). Therefore, the extent of
metal deformation found on a particular pipe section could have influenced
initial weight loss. Pipe-to-pipe variations in welding and machining
explain some of the variation in weight loss among pipes with the same
exposure history.

At longer exposure periods, corrosive attack extended over a wider area of


the pipe surface. In the low-carbonate, high-pH waters, corrosion extended
over the entire pipe surface after two or three months of exposure. In the
less corrosive pH 6.0 and pH 8.0 Seattle tap waters, corrosion extended over
the entire pipe surface after four to six months of exposure. In the pH 7.2

86
high carbonate water, corrosion never completely covered the pipe surface,
even in the oldest pipe section with 460 days of exposure.

In contrast to the thick, loose scales produced in the low-alkalinity waters,


scales produced by the high-carbonate water were thin, hard, and tightly
bound to the metal surface. This scale layer was very dark brown and
somewhat resembled weld spatter. It did not cover the entire metal surface
and, like the other scale types, was often found at metal stress points such
as the butt weld. Removing this scale by mechanical methods was very
difficult; it could not be removed by wire brushing with the Dremel tool, so
it had to be chipped off using a stainless-steel spatula. After the scale
was removed, shallow pits were exposed. Some softer, easily removable scale
occurred on top of the hard tenacious scale layer. The total scale
thickness, however, was still much less than the other iron scales. In
general, this scale was about 1 mm or less in thickness. This scale did not
have the tubercle growths of the other scales; although the scale occurred in
patches, these patches had a relatively even thickness.

After long exposure periods, very small amounts of a hard, tenacious


underlying scale were found in some of the pipe sections exposed to the low-
alkalinity waters. This scale resembled the scale produced by the high-
carbonate water. In the low-alkalinity waters, this hard scale was only
found in pipe sections with approximately a year or more of exposure, and it
covered less than 5 percent of the pipe surface. Unfortunately, none of the
hard tenacious scale found on low-alkalinity water pipe sections was saved
for x-ray diffraction or wet chemical analysis.

The possible identity of scale components is suggested by comparisons with


the results published by other researchers. Previous corrosion researchers
have frequently found iron corrosion scales that resemble the scales produced
in the low-alkalinity water types. In general, the soft, dark,
greenish-black interior scale layers are thought to be either magnetite,
green rust, or a mixture of amorphous ferrous and ferric oxides, whereas the
orange-brown exterior crust is often postulated to be a fully-oxidized iron
oxide such as goethite. The hard tenacious scale found in the high-carbonate

87
pipe and in older pipes exposed to low-alkalinity waters resembles the
photographs of siderite from a study by Kolle and Rosch (1980).

Corrosion-Scale Weights

The weights of corrosion scales that formed on iron pipe sections are
shown in Figure 22. The low-carbonate, high-pH waters produced the thickest
and heaviest scales. These waters also produced the highest corrosion rates,
thereby emphasizing the fact that thick scale formation does not necessarily
provide corrosion protection. In fact, although a thick scale developed,
corrosion rates in the pH 8.0 low-carbonate water did not decline at all
during the six months of testing.

On the other hand, scale formation had a very dramatic effect on corrosion in
the pH 7.2 high-carbonate water. The formation of a protective scale during
the first 60 days of exposure correlated with a sharp decline in corrosion
rate. In this water type, the total weight loss at 56 days was not
appreciably different from that at 200-300 days. Similarly, the scale weight
at 60 days is approximately the same as the scale weight at 200 days. These
data suggest that little or no corrosion occurred after formation of this
protective scale layer.

Scale weight in the 1 mg/L PO^P loop was comparable to that in the pH 8.0
loop with no P04 . In the 5 mg/L P04 -P loop, scale weight was about one-third
of that in the loops with 0 or 1 mg/L-P04 -P. These scale weights correlate
reasonably well with the relative corrosion rates in the three loops. There
was no evidence that scale weight or corrosion rate had reached a plateau in
either of the loops with P04 inhibitor.

X-Rav Diffraction Analysis of Iron Corrosion Scale

Air-dried iron scales were analyzed using powder x-ray diffraction. The
results of this analysis are shown below.
7000

6000-

5000-
o pH = 6.0 (Water 1)
pH = 8.0 (Water 2)
4000- ;; PO4 - 1 mg/l (Water 6)

A Low Carbonate (Water 3,4)
Scale D PO4 - 5 mg/l (Water 7)
8
3000 H M High Carbonate (Water 5)
0
a
B a
2000- D
o
o

1000-

200 400 600 800


Exposure (days)

Figure 22. Scale Weight Formed on Iron Pipe Exposed to Different Water Qualities.
(Water 3 and 4 Data are not separated)
Water Type Iron Compound Detected

Seattle Tap Water pH 6.0 Goethite (a-FeOOH)


Seattle Tap Water pH 8.0 Goethite (a-FeOOH)
Low-Carbonate Water pH 8.0 & 8.4 Goethite (a-FeOOH)
High-Carbonate Water pH 7.2 Goethite (a-FeOOH) and
Siderite (FeC03 )
1 or 5 mg/L o-P04 -P Amorphous

Goethite was identified in all of the iron scales in the P04 -free loops,
while siderite was found only in the high-carbonate scale. These results
suggest that siderite was a constituent of the hard tenacious scale layer
that composed much of the high-carbonate loop scale. Scale x-ray diffraction
patterns were checked for other iron compounds such as hematite, magnetite,
other crystalline forms of ferric hydroxide, green rust, and for calcium
solids such as calcite and aragonite; however, no other compounds were
identified.

X-ray diffraction analysis only identifies crystalline compounds. Other


compounds may exist in these iron scales, but may be amorphous or too poorly
crystallized to be detectable. In addition, compounds containing ferrous
iron may oxidize to hydrous ferric oxides during air drying and therefore go
undetected. According to Sontheimer et al. (1985), hydrous ferric oxides
were considered for many years to be the only iron corrosion products, since
these are the only products remaining when scales are air dried. In the
opinion of Sontheimer, most in situ iron corrosion products actually contain
substantial amounts of ferrous iron.

90
Chemical Analysis of Iron Corrosion Scale

Total Iron Content

Dried, powdered iron scales were analyzed for their total iron content, and
the results are shown in Table 11. The iron content was higher in scales
that were stored a long time before analysis than in scales that were more
recently sampled. Oven-drying the scales at 105 C also appeared to increase
the iron content, probably by driving off bound water.

TABLE 11
Iron Content of Dried Iron Corrosion Scales

Weight % Iron in Dried Scale


Sample No.
Water Type 12345

Seattle Tap Water, pH 8.0 53 59 60 63 66


Seattle Tap Water, pH 6.0 46 59 61 61
Low-Carbonate, pH 8.0 and 8.4 46 61 62 64
High-Carbonate, pH 7.2 46 56 55 59 62

Average Standard Deviation of Duplicate Samples: + 0.5 percent


Treatment of corrosion scales prior to analysis:
Note: In all cases, scales were removed from the pipe surface,
placed in small plastic bags, and stored in a desiccator prior
to analysis.
#1 Air dried and stored for 2 months.
#2 Air dried and stored for 4.5 months.
#3 Stored 1 week and oven dried at 105 C for 4 hr.
#4 Stored 1.5 months and oven dried at 105 0 C for 4 hr.
#5 Stored 4.5 months and oven dried at 105'C for 4 hr.

The compositional changes that occurred during storage and air-drying show
that analyses of dried scale cannot be used to draw definitive conclusions

91
about the scale as it exists on the pipe surface. After oven-drying and
storage, the iron content of most scale samples approached the 63 percent
iron content of goethite, and goethite was found in the x-ray analysis of all
the dried scales.

In the scale formed in the pH 7.2 high-carbonate water, siderite as well as


goethite was found by x-ray diffraction analysis. The iron content of pure
siderite is 48 percent, which is less than the iron content of most of the
dried high-carbonate scale samples. However, as will be discussed in
subsequent sections, siderite probably composed only a small portion of this
scale, and therefore siderite did not have a large effect on its total iron
content.

Ferrous Iron

The results of ferrous iron analyses are shown in Table 12. Unless otherwise
noted, the results.in this table are for wet scale samples. Some iron
corrosion scales appear to oxidize quickly when air dried. For instance,
less than 1 percent ferrous iron was found in a Seattle tap water scale
removed from a pipe surface and then oven dried for 4 hr. Appreciably more
ferrous iron was found when this scale was dissolved in dilute hydrochloric
acid and analyzed immediately after the pipe section was removed from the
conditioning pipe loop.

For several measurements, scale samples were removed from two different pipe
sections with the same exposure history. These duplicate analyses are shown
as two entries in Table 12. With the exception of one high-carbonate scale
sample, the duplicate ferrous iron results were quite close. The differences
found in these duplicate samples may have been due to either analytical error
or to actual pipe-to-pipe variability.

In Table 12, iron scales which formed in pH 8.0 and pH 6.0 Seattle tap water
are grouped together. In general, the ferrous iron content of these two
scale types was similar and tended to increase with the length of the
conditioning period. One exception to this trend occurred in a pipe section
92
TABLE 12

Ferrous Iron Content of Iron Corrosion Scales

WET SCALES
% Fe+2
Water Type Exposure Period (% of Total Iron)
(days)
Seattle Tap Water
pH 8.0 172 16
pH 8.0 203 40
pH 6.0 351 27
pH 8.0 473 30 and 32
pH 6.0 528 32 and 36
Low-Carbonate Water
pH 8.0 and 8.4 139 41 and 46
256 48 and 48
285 36
394 36
High-Carbonate Water
pH 7.2 198 days 6 and 24
203 days 7
205 days 5 and 7
284 days 7

DRIED SCALESt

% Fe+2
Water Type Exposure Period (% of Total Iron)
(days)

Seattle Tap Water


pH 8.0 203 <1
High Carbonate Water
pH 7.2 56 13
205 6

*Analysis done immediately after scale removal from the pipe surface;
scales were not allowed to air dry.
^Analysis done on dried iron corrosion scales.

93
exposed for 203 days. Since scale thickness also increased with exposure,
increased ferrous iron correlated with thicker scales. Likewise, scales
formed in the low-carbonate waters were the thickest and contained the
highest percentage of ferrous iron. In contrast to the pH 8.0 and pH 6.0
scale samples, the ferrous iron content of the low-carbonate scales appeared
to decrease slightly after about nine months exposure.

One explanation for these results is that ferrous iron is produced as


interior corrosion products are formed. At a certain point, however, scale-
thickness growth slows down. Oxidation of ferrous iron in the exterior
corrosion layers may exceed ferrous iron production, decreasing the overall
fraction of ferrous iron in the scale.

The pH 7.2 high-carbonate scale contained a substantially smaller percentage


of ferrous iron than the scales formed in low-alkalinity waters. All of the
"wet" high-carbonate scale samples contained between 5 percent and 7 percent
ferrous iron, with the exception of one 24 percent outlier.

The most dramatic difference between the high-carbonate scales and the other
scales is that the small amount of ferrous iron in the high-carbonate scales
was apparently resistant to oxidation. There was no detectable loss of
ferrous iron when high-carbonate scale samples were air-dried. This
oxidation-resistant ferrous iron may have been present as the siderite
mineral detected by x-ray analysis. The resistance of siderite to oxidation
has been previously reported by Ghosh et al. (1967), who found no measurable
oxidation of siderite solids that were submerged in oxygenated water for two
weeks. The amount of ferrous iron detected in the high-carbonate scale
samples suggests that 5-10 percent of the total iron is in the form of
siderite in these scales.

In the research presented here, the ferrous iron content of scales was
inversely related to the scale's protectiveness. The pH 7.2 high-carbonate
water produced a scale that contained the least ferrous iron and provided the
most corrosion protection, whereas the low-carbonate waters produced a scale
that had the most ferrous iron and provided the least corrosion protection.
This relationship is in direct contrast to the results reported by Sontheimer
94
et al. (1981). In that study, the scales containing the highest percentage
of iron in the ferrous state were the most protective ones. In protective
scales, ferrous iron was about 90 percent of the total iron, while other less
protective scales were only about 40 percent ferrous. The difference between
the results presented here and Sontheimer's results may have been caused by
the different age of the scales that were analyzed. In the Sontheimer study,
scales were analyzed after only a week or two of exposure. In the study
presented here, scales were several months old. Since the ferrous iron
content of nonprotective scales seemed to increase with the length of
exposure, the initial fraction of ferrous iron in these scales might have
been very low.

At any rate, the results presented here support the thesis that oxidation-
resistant ferrous iron in the form of siderite is an important component of
protective scale. The easily oxidizable ferrous iron formed in low-
alkalinity waters was not as protective. Even though the oxidation-resistant
ferrous iron (presumably as siderite) composed only 5-10 percent of the high-
carbonate scale, this scale was highly protective.

Carbonate Content

Both the freshly sampled iron scales and the dried, powdered iron scales were
analyzed for their carbonate content. The results are shown in Table 13.
Results are given as the molar ratio of carbonate to iron. If all the
carbonate found in the scale was present as FeC03 , then the molar ratio of
carbonate to total iron represents the fraction of total iron present as
FeC03 .

The carbonate results exhibited a similar pattern to those for ferrous iron.
In scales formed in pH 8.0 or pH 6.0 Seattle tap water, or in the low-
carbonate water, the carbonate disappeared when the scales were dried. In
the scale formed in pH 7.2 high-carbonate water, the carbonate was retained,
and dried scales had approximately the same carbonate content as wet scales.

In the scales formed in the low-alkalinity waters, the mineral form of the
carbonate present was not determined, since no carbonate minerals remained in
95
TABLE 13

Carbonate Content of Iron Scales

WET SCALE
C03 - 2/FeT C03 2 YFe+2
Water Type Exposure Mole Ratio Mole Ratio
(days)

Seattle Tap Water 172 0.032 0.20


pH 8.0 203 0.72 0.18
Seattle Tap Water 351 0.034 0.13
pH 6.0
Low-Carbonate Water 139 0.48 0.11
pH 8.0 and 8.4 285 0.52 0.14
NA*
High-Carbonate Water 157 0.055
pH 7.2 203 0.54 0.8
205 . 0.086 1.7
205 0.092 1.3
284 0.111 1.6

DRY SCALE
C03 - 2/FeT C03 2 7Fe+2
Water Type Exposure Mote Ratio Mole Ratio
(days)

Seattle Tap Water 203 <0.01 NDt


pH 9.0 418 <0.01 NA
Seattle Tap Water 474 <0.01 NA
pH 6.0
Low-Carbonate Water 139 <0.01 NA
pH 8.0 and 8.4
High-Carbonate Water 56 0.055 0.4
pH 7.2 100 0.030 NA
203 0.11 1.8

NA - Ferrous iron analysis was not done on this sampl e.


*ND - Ferrous iron was below detection limits in this sample.

96
scales that were dried for x-ray diffraction analysis. The transient nature
of the carbonate may indicate that it was not in a crystalline mineral form.
Similar to the ferrous iron content, the carbonate content of wet scales
seemed to increase with length of the scale-formation period.

Table 13 also shows the mole ratio of carbonate to ferrous iron. As shown by
this ratio, scales formed in the high-carbonate water contained approximately
the same molar amount of ferrous iron and carbonate. This may mean that all
the ferrous iron present in the high-carbonate scales was present as FeC03
(siderite).

Wet scales contained a small amount of conditioning water. The conditioning


water contained alkalinity that could be detected as carbonate present in the
scale; however, calculations showed that the carbonate contribution from
conditioning water was not significant.

THE SIDERITE MODEL

Qualitative Correlation of This Research With the Siderite Model of Iron


Corrosion

To summarize the above results, siderite was identified only in the most-
protective iron scales. The ferrous iron and carbonate detected in these
scales were not lost by oxidation or decomposition when the scale was removed
from the pipe surface and dried. The scale was thin, dark, hard, and
adherent. Scale analysis suggests that siderite composed 5-10 percent of the
scale formed in the high-carbonate water. The siderite formed in a highly
buffered, high-alkalinity water at near neutral pH. Siderite was not
detected in the less-protective scales formed in low-alkalinity waters.

Scale analysis failed to identify the mineral species present in the


less-protective scales. Goethite was the only compound detected by x-ray
diffraction analysis of dry samples. However, goethite cannot be the only
compound present, since both ferrous iron and carbonate were detected in wet
samples of the less-protective scales. In the scales produced in the low-
alkalinity waters, the mineral form of the carbonate and ferrous iron is
97
unknown, and these components disappeared when the scales were air dried.
The loss of carbonate and ferrous iron during air drying may mean that these
components were not present as a crystalline mineral such as siderite.

Most of these results are consistent with the siderite model proposed by
Sontheimer et al. (1981). In that model, high buffer intensity and
alkalinity are thought to promote siderite formation, and the presence of
siderite is thought to promote the formation of protective iron scales. Only
the detection of moderate amounts of ferrous iron and carbonates in the
less-protective scales fails to correlate with this model.

Relationship Between Water Quality Parameters and the Siderite Model

Both carbonate and ferrous iron are necessary for siderite formation and are
therefore important elements of the siderite model. The thermodynamic
equation governing siderite solubility is: Ks = io~ 10 - 68 = (Fe+2 )(C03 ~ 2 ) at
o o
25"C, where (Fe ) and (C03 ~^) are the molar activities of ferrous iron and
carbonate in solution.

The carbonate concentration depends on the solution pH and total carbonate


content {Cj = (H2C03 ) + (HC03 ~) + (C03 ~ 2 )}. An increase in either pH or Cy
causes an increase in the carbonate ion concentration. An increase in the
carbonate ion concentration decreases the amount of soluble ferrous ion in
equilibrium with siderite. Figure 23 shows the soluble ferrous iron
concentration in equilibrium with siderite as a function of pH. This
relationship is shown both for the Cj present in the high-carbonate water
(9.1 x 10" 3 moles/L) and at the Cj present in Seattle tap water (4.0 x 10" 4
moles/L).

Dissolved ferrous iron can form aqueous complexes with both hydroxide and
bicarbonate anions. The primary complexes that form under normal drinking
water conditions are FeOH+ , FeHC03+ , and Fe(OH) 2 (aq). Since only
uncomplexed ferrous iron affects siderite solubility, these complexes
increase the concentration of dissolved ferrous iron necessary to form
siderite. Soluble ferrous hydroxide and carbonate complexes were included in
the total soluble ferrous iron shown in Figure 23. The contribution from
98
-2

Log Ferrous
Iron Cone. 4
(M/L)

-6-

pH

Figure 23. Total Soluble Ferrous Iron in Equilibrium with Siderite. Data Points
Represent Water 2 and Water 5 on Upper and Lower Lines, Respectively
ferrous iron complexes was slight; uncomplexed ferrous iron is the dominant
ferrous iron species throughout most of the pH region shown. The ferrous
hydroxide complex, FeOH+ , begins to predominate above pH 9.3.

From a thermodynamic perspective, siderite will be preferentially formed over


ferrous hydroxide in the pH and Cj regions normally encountered in drinking
water. At the Cj of Seattle tap water, solid ferrous hydroxide becomes less
soluble than siderite only at pH values above 10.3. At the Cj of the high-
carbonate water, solid ferrous hydroxide does not become less soluble than
siderite until the pH is above 11.

In the high-carbonate water at pH 7.2, the soluble ferrous iron concentration


necessary for siderite precipitation is only one-third of that in Seattle tap
water at pH 8.0. Therefore, even at the lower pH, the high-carbonate water
appears to provide more favorable conditions for the formation of siderite
than the conditions in Seattle tap water.

If the thermodynamic solubility constant was the only factor governing


siderite formation, then raising the pH would promote siderite formation.
However, another important element of the siderite model is the kinetics of
ferrous iron oxidation. At the oxygen concentrations normally encountered in
surface waters, ferrous iron is thermodynamically unstable. Ferrous iron can
oxidize to ferric iron and form ferric solids or mixed ferrous/ferric solids.
The oxidation rate of ferrous iron increases with the square of the hydroxide
ion concentration. Therefore, as pH increases, the oxidation rate increases
rapidly.

Figure 24 shows the oxidation rate of ferrous iron as a function of pH, as


given by Stumm and Lee (1961). The ambient oxygen partial pressure of 0.21
atmospheres was used to calculate the oxidation rate. The oxidation rate of
ferrous iron also increases as the ferrous iron concentration increases. The
ferrous iron concentrations used to calculate the oxidation rates were the
concentrations in equilibrium with siderite as shown in Figure 23. At a
given pH, a higher ferrous iron concentration is necessary to form siderite
in the low-carbonate Seattle tap water, so the oxidation rate is higher. In
Seattle tap water at pH 8.0, the ferrous iron concentration in equilibrium
100
0.18

0.15-

0.12-

Ct = 4.0 E-4 M/L


Oxidation Rate 0.09 J
(mM/L/min)

0.061

0.03-

Ct = 9.1 E-3 M/L


0.00

Figure 24. Oxidation Rate of Ferrous Iron in Equilibrium with Siderite. Data Points
Represent Water 2 and Water 5 on Upper and Lower Lines, Respectively
with siderite is 1.1 x 10"^, and the oxidation rate is 1.8 x 10" 4 mol/L-min.
In the high-carbonate water at pH 7.2, the ferrous iron concentration in
equilibrium with siderite is 3.3 x 10~ 6 , and the oxidation rate is 1.4 x 10" 6
mol/L-min, almost 100 times less than in pH 8.0 Seattle tap water. In the pH
7.2 high-carbonate water, both the higher carbonate concentration and the
lower pH contribute to a lower iron oxidation rate and therefore promote
siderite formation.

In the siderite model, a high buffer intensity is thought to suppress iron


corrosion. In the corrosion process, the reduction of oxygen at the cathode
produces hydroxide ions by the following reaction:
H20 + 1/2 02 + 2 e- 2 OH"
Hydroxide production raises the pH in the area of the cathode, increasing
ferrous iron oxidation rates and therefore inhibiting siderite formation.
Highly buffered water, however, resists these pH changes.

Figure 25 shows buffer intensity as a function of pH at the total carbonate


concentrations of the high-carbonate water and of Seattle tap water. The
buffer intensity in the high-carbonate water at pH 7.2 is many times greater
than that in pH 8.0 Seattle tap water. Both the lower pH and the higher
carbonate level contribute to the high-carbonate water's greater buffer
intensity. In the low-alkalinity Seattle tap water, even the buffer
intensity peak at pH 6.3 is lower than the minimum buffer intensity in the
high-carbonate water. Therefore, in very low alkalinity waters, even the
maximum buffer intensity obtainable by pH adjustment may be too low to
provide for corrosion protection.

In summary, higher total carbonate concentrations encourage siderite


formation by decreasing the ferrous iron concentrations necessary for
siderite precipitation and also by increasing the water's buffer intensity.
Increases in pH encourage siderite formation by converting bicarbonate ions
to the carbonate form, but the higher pH also counteracts siderite formation
by increasing the ferrous iron oxidation rate and decreasing the buffer
intensity. In the research presented here, high-carbonate concentrations and
moderate pH levels appeared to promote siderite formation.

102
5- Ct = 9.1 E-3 M/L

4-

Buffer Intensity 3-
mM/L-pH unit

2H

Ct = 4.0 E-4 M/L

Figure 25. Calculated Buffer Intensity as a Function of pH for Waters With the
Indicated Concentration of Inorganic Carbon and No Other Weak Acids or Bases.
Data Points Represent Water 2 and Water 5 on Lower and Upper Lines, Respectively
A Quantitative Iron Corrosion Model Based on Siderite Formation

A complete model of iron corrosion and scale formation would be very complex.
Such a model would include the oxidation reactions of iron metal and ferrous
iron, the reduction reaction of oxygen, the precipitation reactions of all
potential iron solids, and the changes occurring in already-formed iron
solids such as the dehydration and oxidation of precipitated ferrous solids
to form goethite. Such a model would also have to account for transport of
reactants to the metal surface and of products away from the surface. This
transport is influenced not only by the flow of water through the pipe but
also by the properties of the developing scale and by the nature of the
anodic and cathodic reactions.

On some metals, equilibrium modeling can be used to predict the water quality
conditions that will cause protective scales to form. Unfortunately, the
equilibrium modeling of iron corrosion scale formation is severely limited by
incomplete knowledge of the identity and crystalline structure of solids that
exist in the scale (Snoeyink and Kuch, 1985). Even with x-ray diffraction
analysis and wet chemical analysis, it was not possible to determine the
complete composition of the corrosion scales. The siderite model suggests a
general explanation for how water quality affects the formation of protective
scale, but it does not account for all of the reactions that may occur in the
iron corrosion process.

Because the complexity of the process precludes formulation of a complete


model, a very simple model was constructed to examine the effects of iron
oxidation rates more closely. In this model, iron corrosion at the metal
surface is assumed to produce a uniform concentration of ferrous iron that is
trapped beneath a scale of uniform composition and thickness. These
assumptions would be valid if diffusion from the metal surface was fast
compared with subsequent reactions involving the ferrous ions, and if the
ferrous ions were trapped by a barrier to diffusion such as a developing
scale layer.

104
Using these assumptions, the production rate of ferrous iron from the
corrosion reaction was compared with the oxidation rate of ferrous iron.
From this comparison, it was determined whether it was possible for the
ferrous iron concentration to build up enough for siderite to precipitate.

Because reliable data was not available for the chemical conditions at the
metal surface, the chemical conditions were assumed to be the same as those
in the bulk solution. The pH and total carbonate concentration of the bulk
solution were used in calculations. The oxygen concentration was assumed to
be the saturation value in equilibrium with an atmospheric oxygen
concentration of 0.21 atmospheres partial pressure.

As a first approximation, the ferrous iron was assumed to be trapped in a


layer equivalent to a 2-mm-thick 100 percent void layer. At an iron
corrosion rate of 6 mils/year, ferrous iron would be produced in the void
space between the metal surface and the scale surface at a rate of 2 x 10" 5
moles/L-min. At ferrous iron concentrations in equilibrium with siderite,
the ferrous iron oxidation rate in pH 8.0 Seattle tap water is 1.8 x 10
moles/L-min. At concentrations high enough to produce siderite, ferrous iron
oxidizes much faster than it is produced, which is impossible. Therefore,
this model suggests that it is impossible to precipitate siderite in the pH
8.0 Seattle tap water, because ferrous iron cannot build up to a high enough
concentration before it is oxidized to the ferric form.

At ferrous iron concentrations that are in equilibrium with siderite in the


pH 7.2 high-carbonate water, the ferrous iron oxidation rate is 1.4 x 10" 6
moles/L-min. Assuming a similar ferrous iron production rate of 2 x 10" 5
moles/L-min, the oxidation rate is much less than the production rate.
Therefore, ferrous iron can exist at concentrations high enough to produce
siderite.

In the above calculations, the ferrous iron production rate was about an
order of magnitude different than its oxidation rate for water qualities
similar to both Seattle tap water and the high-carbonate water. Since the
difference between the two rates was large, the conclusions drawn from the
model are fairly insensitive to small variations in the assumed values of the
105
corrosion rate~and scale thickness. The assumptions of a 2-mm-thick scale
layer and a corrosion rate of 6 mpy are reasonable approximations based on
the results of this research. However, the conclusions from the model would
be the same even if slightly different values had been used.

These rough calculations illustrate how oxidation rates could limit the
potential for siderite formation by not allowing the ferrous iron to build up
to concentrations high enough for siderite precipitation. Although this
crude model cannot represent the real iron corrosion and scale formation
mechanisms because important details such as diffusion and concentration
gradients are ignored, its correspondence with observed trends is satisfying.

The scale formed in Seattle tap water contained ferrous iron; however, this
ferrous iron oxidized as the scale dried. It is possible that this ferrous
iron was incorporated into a mixed ferrous/ferric solid like FegC^ or
amorphous green rust. Much lower ferrous iron concentrations are necessary
for formation of these mixed ferrous/ferric solids than are required for
siderite formation. The ferrous iron produced by corrosion in Seattle tap
water is therefore either oxidized or precipitated as a mixed ferrous/ferric
solid before it can build up to concentrations high enough for siderite
precipitation. The formation of these mixed ferrous/ferric solids might
explain why the ferrous content of scales in our study did not correlate with
scale protectiveness as they did in the experiment reported by Sontheimer et
al. (1981). Although ferrous iron was present in the scales, none was
present in the siderite form.

Corrosion scales may hinder the diffusion of oxygen to the metal surface and
cause reducing conditions to develop beneath scale layers. If siderite was
present in the small amounts of hard, tenacious scale formed in the low-
alkalinity waters after long exposure periods, it may have been formed by
this mechanism. When less oxygen is present, ferrous iron oxidation rates
decrease and higher concentrations of ferrous iron can result. This may
cause the formation of siderite even when conditions in the bulk solution are
unfavorable. Even though siderite only composed 10 percent or less of the
high-carbonate scale, this scale was still protective, presumably because the
other components of the scale retained the dense structure of siderite. In
106
the siderite model, scales do not have to be entirely composed of siderite to
be protective. Siderite is thought to form first near the metal surface and
then promote the formation of protective scale layers, but the protective
layers may be composed of other minerals such as ferric hydroxides that
retain the siderite structure. In the absence of siderite, precipitated
ferrous and ferric minerals form loose, amorphous, and nonprotective scales.
In this study, 90 percent to 95 percent of the iron in the protective scales
which formed in the high-carbonate water was in the ferric form. X-ray
diffraction analysis suggests that the ferric iron was present as goethite.
This goethite may have retained the dense structure of siderite.

RESPONSE OF SCALE-COVERED MILD-STEEL PIPE TO CHANGES IN THE ALKALINITY AND pH


OF THE CONDITIONING WATER

Corrosion-Rate Measurements

A series of experiments was conducted to determine how changes in water


chemistry affect corrosion rates on already-scaled iron pipe. Mild-steel
pipe sections were exposed to flowing water in the circulation loops for
about 90 days to allow development of a corrosion scale. After this
conditioning period, the pipe sections were installed in the mini-loop and
the corrosion rate was measured using the dissolved-oxygen-depletion method.
This initial corrosion-rate measurement was made in the same water as the
pipes already had been exposed to in the conditioning loop. Then water of a
different pH and alkalinity was circulated through the pipe sections. The
pipe sections were left in contact with the new water type for several weeks.
During this period, changes in corrosion rate were monitored using the
dissolved-oxygen-depletion method.

The following three experiments were conducted. In the first experiment,


three pipe sections initially exposed to the pH 7.2 high-carbonate water were
monitored after exposure to pH 8.0, Seattle tap water. In the next
experiment, three other pipe sections initially exposed to the pH 7.2 high-
carbonate water were monitored after exposure to the pH 6.0 water. In the

107
third experiment, two pipe sections initially exposed to the pH 8.0 Seattle
tap water were monitored after exposure to pH 7.2 high-carbonate water.

The results of these experiments are shown in Figures 26 through 28. The
initial oxygen-depletion corrosion rate is shown on the left-hand edge of
each graph. The initial rate shown on the graph is the average of three
measurements taken on the pipe sections shortly before their exposure to the
new water type.

The high-carbonate pipe sections responded differently to the pH 6.0 water


than they responded to the pH 8.0 water. In the pH 6.0 water, the corrosion
rate increased dramatically after a few weeks of exposure, whereas no
increase was observed in the pH 8.0 Seattle tap water. Apparently, the high-
carbonate scale is more vulnerable to attack by the pH 6.0 water because of
the water's lower pH, lower alkalinity, or higher chloride ion content.

When Seattle tap water pipe sections were placed in the pH 7.2 high-carbonate
water, the corrosion rate initially increased. After several weeks, however,
the corrosion rate decreased below the rate measured before exposure to the
high-carbonate water.

At the end of the experiments, weight-loss determinations were performed on


these pipe sections, and the results were compared to the dissolved-oxygen
depletion data. As discussed in the previous section on corrosion-rate
measurements, comparing weight loss and dissolved oxygen depletion data is
not straightforward; weight loss is a cumulative measure of corrosion rates,
while dissolved-oxygen depletion measures the corrosion occurring during a
time period of a few hours or less. Weight-loss results for the high-
carbonate pipe sections exposed to pH 8.0 Seattle tap water and for the
Seattle tap water pipe sections exposed to the high-carbonate water are shown
in Table 14. Also shown are representative weight-loss data from high-
carbonate and Seattle tap water pipe sections of the same total exposure
period. Weight-loss data were not obtained for the high-carbonate pipe
sections exposed to pH 6.0 water because, unfortunately, the outside surfaces
of these pipe sections became corroded. The protective acrylic finish on
these pipe sections was damaged by repeated handling.
108
3.0

2.5-

2.0-
Corrosion
Rate, mpy

o
vO

20 40 60 80 100 120
Days exposed to pH 8.0 Water

Figure 26. Corrosion Rate of High Carbonate Scale (from Water 5)


Exposed to pH 8.0 Tap Water (Water 2)
12

10-

Corrosion g-
Rate, mpy

40 60 80

Days Exposed to pH 6.0 Water

Figure 27. Corrosion Rate of High Carbonate Scale (from Water 5)


Exposed to pH 6.0 Tap Water (Water 1)
16

14-

12-

10 H
8
Corrosion
Rate in 8-
Seattle
Tap Water
61

4-

2- Q

0
0 10 20 30 40 50 60 70 80 90 100
Number of Days in High Carbonate Water

Figure 28. Seattle Tap Water Scale (from Water 2) Exposed to High Carbonate Water (Water 5)
TABLE 14
Weight-Loss Data From Changing Water Chemistry Experiments

Exposure Period Weight Loss


(mg)

High-Carbonate Water for 120 days and


pH 8.0 Seattle Tap Water for 113 days 893 170
for a Total Exposure of 233 days

High-Carbonate Water for 227 days 502 187

pH 8.0 Seattle Tap Water for 120 days and


High-Carbonate Water for 110 days 1964 + 228
for a Total Exposure of 230 days

pH 8.0 Seattle Tap Water for 234 days 1895 87

High-carbonate pipe sections exposed to pH 8.0 Seattle tap water had a higher
weight .loss than pipe sections exposed only to high-carbonate water for the
same total exposure period. In contrast, corrosion-rate measurements by the
dissolved-oxygen-depletion method showed no increase in corrosion rates after
the high-carbonate pipe sections were exposed to pH 8.0 Seattle tap water.
These conflicting results may be explained by referring to Figure 16, the
cumulative weight loss graph for pipe sections exposed to pH 7.2 high-
carbonate water. From this graph, weight loss in the high-carbonate water
does not appear to increase with exposure period. This may be because very
little corrosion occurred on high-carbonate pipe sections after the formation
of a protective scale layer. Thus, if these pipe sections had been left in
the high-carbonate water instead of being transferred into Seattle tap water,
their dissolved-oxygen-depletion corrosion rates would probably have
112
decreased to near zero. Therefore, even though the corrosion rates of the
high-carbonate pipe sections did not increase in pH 8.0 Seattle tap water,
they also did not decrease significantly, and so they still experienced more
weight loss. Unfortunately, two factors increase the difficulty of comparing
the two types of corrosion-rate measurements in this experiment: the
significant pipe-to-pipe weight-loss variations in the high-carbonate water,
and the large variations that occur when measuring very low corrosion rates
by the dissolved-oxygen-depletion method.

Seattle tap water pipe sections exposed to the high-carbonate water


experienced approximately the same weight loss as regular Seattle tap water
pipe sections of the same exposure period. This may have been because the
initially high corrosion rate experienced when the pipe sections were first
exposed to the high-carbonate water was balanced by a low corrosion rate
occurring toward the end of the experiment.

Scale Analysis

After the experiments were complete, the pipe sections were examined
visually, and in some cases the scale was removed for analysis. Over the
course of the experiment, the appearance of the scales did not change
appreciably. However, the high-carbonate pipe sections exposed to the pH 6.0
water and the pH 8.0 Seattle tap water did seem to thicken somewhat. They
also appeared to gain a higher percentage of soft dark material than that
which was present in regular high-carbonate scales. The pH 8.0 Seattle tap
water scale exposed to the high-carbonate water still had the lumpy, thicker
structure indicative of tubercle formation. However, the tubercles were less
pronounced than those on regular pH 8.0 Seattle tap water pipe sections of
the same exposure period.

The removed corrosion scales were analyzed twice for their ferrous iron and
carbonate contents; immediately upon removal (while still wet) and after oven
drying. The results of these analyses are shown in Table 15. For
comparison, Table 15 also shows representative values found in regular pH 7.2
high-carbonate and pH 8.0 Seattle tap water scales.

113
TABLE 15

Ferrous Iron and Carbonate Content of Iron Scales Exposed


to Changes in Water Chemistry

Ferrous Iron Content: Results given as the percentage of iron in the ferrous
state.
Seattle tap water scales; 203 days of exposure.
Wet Scale: 40
Dry Scale: < 1
Seattle tap water for 4 months; transferred to the high-carbonate water for 4
months; 230 days of total exposure.
Wet Scale: 18
Dry Scale: 9 and 10
High-carbonate water for 4 months; transferred to Seattle tap water for 4
months; 233 days of total exposure.
Wet Scale: 19
Dry Scale: 11
High-carbonate water scale; 203 and 205 days of exposure.
Wet Scale: 5-7
Dry Scale: 6

Carbonate Content: Results given as the mole ratio of carbonate to iron metal.
Seattle tap water scales; 203 days of exposure.
Wet Scale: 0.072
Dry Scale: <0.01
Seattle tap water for 4 months; transferred to the high-carbonate water for 4
months; 230 days of total exposure.
Wet Scale: 0.097
Dry Scale: 0.072 and 0.087
High-carbonate water for 4 months; transferred to Seattle tap water for 4
months; 233 days of total exposure.
Wet Scale: 0.095
Dry Scale: 0.092
High-carbonate water scale; 203 and 205 days of exposure.
Wet Scale: 0.054, 0.086, and 0.092
Dry Scale: 0.011

114
The most striking result of this analysis is that when the coupons from the
pH 8.0 Seattle tap water loop were exposed to the pH 7.2 high-carbonate
water, new scale containing ferrous iron and carbonate formed, and these
components remained after the scales were dried. Oxidation-resistant ferrous
iron and carbonate had previously been found only in high-carbonate scale.
This may indicate that siderite was formed in these scales after exposure to
the high-alkalinity water. The formation of siderite correlates with the
decreased corrosion rates measured toward the end of the exposure period
involving the high-alkalinity water.

The high-carbonate scale exposed to pH 8.0 Seattle tap water retained its
oxidation-resistant ferrous iron and carbonate content. The wet scale,
however, had more ferrous iron than the dry scale and had a higher ferrous
iron content than most of the regular high-carbonate wet-scale samples. This
high-carbonate scale appeared to contain some of the easily oxidized ferrous
iron normally found only in the Seattle tap water scales. This may mean that
some of the ferrous iron in these scales was incorporated into less-
protective, mixed ferric/ferrous solids instead of into the more-protective
siderite layer.

The Effect of High Conductivity on Initial Corrosion Rates

The corrosion rates of pipe sections exposed to pH 8.0 Seattle tap water
increased immediately after the coupons were exposed to the pH 7.4 high-
carbonate water, and later decreased. The initial increase may have been
caused by either the higher conductivity of the water or its lower pH. To
investigate this issue, a few short-term experiments were conducted where the
conductivity of the exposure water was increased while the pH was held
constant.

In these experiments the corrosion rate of prescaled pipe sections from the
Seattle tap water loop was first measured by the dissolved-oxygen-depletion
method. The coupons were then exposed to one of two higher-conductivity
waters, and the corrosion rate was measured again. In one solution,

115
conductivity had been increased by addition of sodium chloride, and in the
other solution the same increase in conductivity had been induced by addition
of a mixture of hydrochloric acid and sodium bicarbonate. The final pH (8.0)
and conductivity (350 /imhos/cm) of both solutions was the same, but the
former had a higher chloride content and the latter a correspondingly higher
bicarbonate content.

As shown in Table 16, the corrosion rate increased in both cases, and the
magnitude of the increase was comparable. The absence of a significant
difference between the corrosion rates in the two high-conductivity solutions
suggests that the chloride ion was not acting primarily as a specific
promoter of corrosion; rather, the conductivity seemed to act as a
nonspecific corrosion promoter, with chloride and bicarbonate contributing
equally to the effects.

Based on these experiments, the transient increase in corrosion rate when


coupons from the pH 8.0, tap water loop were exposed to high-carbonate water
could at least plausibly be attributed to the higher conductivity of the
latter water. The subsequent decrease in corrosion rate would then be
attributed to the specific corrosion-inhibiting effect of the high-carbonate
content, allowing formation of siderite once enough ferrous iron had
accumulated in the scale.

These experiments provide an excellent example of the type of data that can
be collected using the DO depletion method for evaluating corrosion rate.
For these experiments, weight loss would have been too small to detect and,
as noted earlier, electrochemical measurements on iron surfaces are very
imprecise due to variations in electrical potential along the surface.
Understanding the strengths and limitations of each approach allows the
experimenter to choose the best method for the particular experiment and
thereby obtain the most accurate results.

116
TABLE 16
Response of Corrosion Rate to High-Conductivity Water

Sodium Chloride Addition


Seattle Tap Water After NaCl Addition
pH 8.0, cond. = 50 pino/cm pH 8.0, cond. = 350 /jmho/cm
Trial #1* 10+1 mpy 17.5 + 1 mpy

Sodium Bicarbonate Addition


Seattle Tap Water After NaHC03 Addition
pH 8.0, cond. = 50 /who/cm pH 8.0, cond. = 350 /imho/cm
Trial #2} 12 + 1 mpy 16 +1 mpy
Trial #3^ 10 1 mpy 16.5 1 mpy

Trial #1: Pipe sections were exposed to pH 8.0 Seattle tap water for 62
days before testing.
'Trial #2: Pipe sections were exposed to pH 8.0 Seattle tap water for 47
.days before testing.
tTrial #3: Pipe sections were exposed to pH 8.0 Seattle tap water for 62
days before testing.

Stumm (1960) also showed that a higher solution conductivity can cause higher
initial corrosion rates. After an initial conditioning period, Stumm found
that the effect of high solution conductivity was masked by the formation of
protective scales. In the study presented here, corrosion rates decreased
after the Seattle tap water pipe sections were exposed for several weeks to
the high-carbonate water, presumably because a more protective scale
developed. Like Stumm's experiments, the protective scale counteracted the
corrosion-rate increase caused by the increased conductivity.

SUMMARY

To summarize this part of the work, the corrosion rates of mild-steel pipes
and the composition and structure of corrosion scales were related to water
quality variables.

117
The most protective iron corrosion scale was formed in a high-alkalinity,
moderate-pH water (alkalinity 400 mg/L as CaC03 , pH 7.2). Pipe sections
coated with this scale had very low corrosion rates (about 2 mpy or less).
The protective scale was thin, hard, dark, and very adherent to the metal
surface. Siderite was found in this scale by x-ray diffraction analysis.
This was the only scale where siderite was detected. Wet chemical analysis
of this corrosion scale before and after air-drying showed that 5-10 percent
of the iron was in the ferrous state. Small amounts of carbonate were also
found in both wet- and dried-scale samples; the ratio of carbonate to total
iron was approximately the same as the ratio of ferrous iron to total iron.
This suggests that most of the ferrous iron in this scale was present as
FeC03 , i.e., siderite.

Waters with alkalinities of 20 mg/L as CaC03 or less produced significantly


higher corrosion rates than the high-alkalinity (400 mg/L as CaC03 ) water.
Iron corrosion rates were generally between 10 and 20 mpy in all of the low-
alkalinity waters studied.

Scales formed in the low-alkalinity waters were thick and loose, consisting
of large tubercles covering shallow pits. After nine months to one year of
exposure to low-alkalinity water, small amounts of hard, tenacious scale were
formed under the soft, loose scale layers. Goethite (a-FeOOH) was the only
mineral detected in the x-ray diffraction analysis of dried low-alkalinity
scales. However, goethite was not the only compound present in these scales,
since ferrous iron and carbonate were detected in scale samples before they
were air dried.

Addition of 5 mg/L o-P04 -P at pH 8.0 reduced corrosion rates and scale


accumulation by about two-thirds compared to o-PC^-free controls. Solutions
containing 1 mg/L o-P04 -P were about as corrosive as the controls.

Corrosion rates in several of the exposure loops decreased continuously for


at least six months, and often longer. Therefore, six-month exposure tests
may be inadequate to fully characterize corrosivity of water to steel pipes.

118
In general, increases in corrosion rates correlated with decreases in buffer
capacity. In two waters, the buffer capacity of Seattle tap water was
decreased by bubbling the water with CC^-free air. Both of these waters were
more corrosive than regular pH 8.0 Seattle tap water. The pH 7.2 high-
carbonate water had by far the highest buffer capacity and was much less
corrosive than any of the low-alkalinity waters. The pH 6.0 Seattle tap
water was the only exception to the correlation between buffer capacity and
corrosivity; it had a higher buffer capacity than Seattle tap water, but was
also more corrosive. At first, the pH 6.0 water produced similar corrosion
rates to the pH 8.0 Seattle tap water, but after a year of exposure the pH
6.0 water was more corrosive.

Increased conductivity may have accelerated corrosion in those waters where


HC1 was added to adjust pH. The increased conductivity in these waters may
also explain why buffer capacity failed to predict the relative corrosivity
of the pH 6.0 water. Likewise, the high corrosion rates in the pH 8.0 low-
carbonate water may have been caused by its higher conductivity, since its
buffer intensity was only slightly less than the buffer intensity of pH 8.0
Seattle tap water.

119
CORROSION OF COPPER PIPES

COPPER CORROSION RATES

Copper coupons were exposed to water of various qualities in the


recirculating loops described in Chapter 3. After various exposure periods,
they were taken out of the loops and weighed, both before and after the scale
was removed. The results will be discussed in two groups: systems of varying
pH and carbonate content, but containing no ortho-phosphate in one group, and
systems at fixed pH but with varying ortho-phosphate content in the other.

Figures 29-32 show the results for weight loss in the four loops where ortho-
phosphate was not present, and Figure 33 presents the data for all these
loops on a single graph. Considering that the weight loss represents about
0.02 percent to 0.4 percent of the total weight of the coupon, and that every
data point reflects an independent experiment, the results for each of the
four data sets are remarkable for their internal consistency. In fact, in
each loop the cumulative weight loss not only increases quite steadily with
duration of exposure, but the data are reasonably linear. It appears that
the corrosion rates in the pH 6.0 Seattle tap water and the pH 7.2 high-
carbonate water may be somewhat greater during the first few weeks of
exposure than thereafter, but the corrosion rate in each of the loops is
nearly constant from about 50 days of exposure until the end of the test,
which lasted between 400 and 700 days.

The approximate constancy of the corrosion rate for a water quality


intermediate between that of the pH 6.0 and pH 8.0 tap water loops was
confirmed by an independent analysis using linear polarization. This
electrochemical procedure gives an indication of the instantaneous corrosion
rate of a coupon, as opposed to the integrated, average rate over the
duration of the test, which is the value derived from the weight-loss
measurements. Figure 34 shows the electrochemically derived corrosion rate
of coupons after exposure to pH 6.0 tap water for periods of 2 hr to 200
days. The initial rate is substantially higher than the rate at later times;
in fact, after only 20 hr the rate drops to 50 percent of its value after 2
hr. There is a slight additional decrease in rate between days 1 and 7, and
121
200

150-

Weight Loss
(mg)

o
o e
50- o 8

200 400 600 800


Exposure (days)

Figure 29. Weight Loss of Copper Pipe Exposed to


pH 8.0 Seattle Tap Water (Water 2)
150'

100-

Weight Loss
(mg)

N) o
OJ
o
50-
o

8
o

100 200 300 400


Exposure (days)

Figure 30. Weight Loss of Copper Pipe Exposed to Low Carbonate


Seattle Tap Water at pH = 8.0 (Water 4)
250

2(XH
o
o
8o
150-

Weight Loss
o
(ing)
0 8
100-

50-
08
o

100 200 300 400 500


Exposure (days)

Figure 31. Weight Loss of Copper Pipe Exposed to High Carbonate


Seattle Tap Water at pH= 7.2 (Water 5)
250

200-

150-

Weight Loss
(rag) 8
o
100-
(O

00

8
o
50-

200 400 600 800


Exposure (days)

Figure 32. Weight Loss of Copper Pipe Exposed to Seattle Tap Water
Adjusted to pH= 6.0 (Water 1)
250

200-

g
D

150-
Dn e o pH= 6.0 (Water 1)
Weight Loss pH= 8.0 (Water 2)
(mg) 8
o pH= 8.0 Low CO3 (Water 4)
100-
n pH= 7.2 High COS (Water 5)
NJ

50-
cflo
i _
*
ol

200 400 600 800


Exposure (days)

Figure 33. Weight Loss of Copper Pipe Exposed to Various Water


Qualities in the Absence of Phosphate
260

220 -
i
180- Q Cathodic Tafel Slope
* Anodic Tafel Slope
Ba,Bc
(mV/dec)
100 -I

60

20

2.0

1.5:
lo i.o:
(uA/cm2)

o.s:
1
o.o
.01 i 10 100 1000
Surface Exposure (days)

Figure 34. Variation in Corrosion Current and Tafel Slopes on an Aging Surface Exposed to
pH 6.0 Tap Water (Water 1). Error Bars Represent One Standard Deviation of
Replicate Measurements on Different Surfaces
virtually no further change during the next 200 days. The figure also shows
the changes in Tafel slopes over the 200-day period. The increase in both
slopes indicates that both the oxidation and reduction reaction are being
inhibited over time, which is consistent with the formation of a thin scale
gradually covering the pipe surface, blocking both anodic and cathodic sites.

The corrosion rates during the majority of the test period can be estimated
from the slopes of the graphs. Within the limits of precision attainable
from such an analysis, the corrosion rates in the loops with pH 6.0 or 8.0
tap water and with pH 7.2 high-carbonate water were identical from day 50
forward. Almost all the data points for the pH 8.0 tap water loop indicated
that less corrosion had occurred in this loop than in the other two loops
after equivalent exposure periods, but these differences seemed to reflect
different rates during the first 50 days of testing, rather than
substantially different rates thereafter. The corrosion rate in the pH 8.0
low-carbonate loop was about one-third of that in the other three throughout
the test period.

Addition of orthophosphate to the water had a very dramatic effect on copper


corrosion rates. Figures 35 and 36 present data for weight loss and scale
weight in these systems, along with the pH 8.0 tap water loop, which had the
same pH but no P04 . Although the cumulative weight loss was definitely
greater in the low-phosphate than in the high-phosphate system, the corrosion
rates in both systems were extremely low and were dramatically less than in
the control loop from the start of the test. The differences between the
loops with and without ortho-phosphate grew even larger with longer exposure
times, because while corrosion continued steadily in the phosphate-free loop,
it seemed to cease altogether in the high-phosphate (5 mg/L P) loop and
proceeded very slowly if at all in the low-phosphate loop after a few months.

SCALE ACCUMULATION AND COMPOSITION

Figure 37 displays the results for the accumulated scale on the coupons in
the phosphate-free systems after various exposure histories. Whereas the
weight loss is a measure of the overall corrosion rate, scale weight reflects
combined effects of corrosion rate, chemical composition of the scale, and
128
250

200-

1501
o pH= 6.0 (Water 1)
Weight Loss
(mg) 8 pH= 8.0 (Water 2)
o
100- pH= 8.0 PO4= 1 mg/L (Water 6)

n pH= 8.0 PO4= 5 mg/L (Water 7)


00

50-
o
<p >
ol
D C

200 400 600 800


Exposure (days)

Figure 35. Weight Loss Comparison of Copper Pipes in the Presence


and Absence of Phosphate
80

70-

60-

50-
o pH= 6.0 (Water 1)

Scale pH= 8.0 (Water 2)


Weight 40 ":
(mg) pH= 8.0 PO4= 1 mg/L (Water 6)
30- Q pH= 8.0 PO4= 5 mg/L (Water 7)

20- o
o
t
10-
a
a

0 100 200 300 400 500 600 700 800


Exposure (days)

Figure 36. Scale Weight on Phosphate-Inhibited and Non-Inhibited


Copper Pipe
80

70-

60-

50-
o pH= 6.0 (Water 1)
Scale
Weight 40 ~ pH= 8.0 (Water 2)
(mg)
pH=8.0LowCO3(Water4)
30-
B a pH= 7.2 High CO3 (Water 5)

o DO
20-
- S-'
10-

0 100 200 300 400 500 600 700 800


Exposure (days)

Figure 37. Scale Weight on Copper Pipes Exposed to Different Water Qualities
the tendency of the scale to adhere to the pipe walls and resist erosion.
Although the pH 6.0 and 8.0 tap water loops and the high-carbonate loop had
similar corrosion rates, coupons exposed to pH 8.0 Seattle tap water had much
more adherent scale than any of the other coupons.

The composition of the scale in the copper coupons was analyzed by dissolving
the scale in an ammonia solution. In the four loops under discussion, copper
represented between 69 percent and 84 percent of the scale that dissolved
(Table 17). The quantity of scale which could be collected from the usual
polishing procedures was insufficient to use for x-ray diffraction analyses.
Rather, flat copper plates were exposed to the water in the pH 6.0 and 8.0
tap water loops to generate enough scale for this purpose. The scales formed
in this way in both loops contained a cuprous oxide mineral known as cuprite,
Cu20. This mineral is 89 percent Cu, 11 percent 0 by weight, which is
somewhat more copper-rich than the analyses of the scale indicated. The wet
chemical analysis is more consistent with the scale being either a cupric
oxide, CuO, or a hydrated cuprous oxide, Ci^O't^O, which are 80 percent and
79 percent copper by weight, respectively. It'is possible that the scale
contained solids such as these that were amorphous and therefore were not
detectable by x-ray diffractometry. No plates were exposed to water in the
low- and high-carbonate loops.

An estimated mass balance on Cu can be computed after a representative period


of 400 days exposure in the four P04 -free loops, assuming all the scale
formed in all four loops was cuprite. Based on such an analysis, only about
one-tenth of the copper that had corroded remained attached to the coupon as
scale in the loops at pH 6.0 and 7.2, and one-fourth to one-third remained
attached in the loops at pH 8.0. The remaining copper either dissolved
directly without forming scale or formed scale that was eroded from the
coupon surface and entered the water.

The one time that copper was analyzed in the recirculating water, essentially
100 percent of it was present as particulate matter. This may have been
either cuprite or a cupric oxide (CuO), since oxidation of Cu(I) to the
cupric form would have been thermodynamically favorable in the oxygenated

132
TABLE 17
Composition of Copper Corrosion Scales

Copper Weight Copper


Age Loss Detected Percent
(days) (mg) (mg) Copper

Loop 1
Coupon #103 500 21.4 16.6 78

Loop 2
Coupon #118 500 57.0 39.1 69

Loop 3
Coupon #124 445 24.1 17.6 73
Loop 4
Coupon #212 71.0 55.0 78
Coupon #211 53.0 41.0 81

LoopB
Coupon #207 330 2.3 1.1 48

Loop 6
Coupon #328 190 29.6 20.9 71

Clean Coupon
Coupon #557 6.0 5.0 83

133
water. Efforts to quantify the fraction of copper in each oxidation state in
either the attached or suspended scale were unsuccessful.

There was a good deal of imprecision in the scale weight data for the
phosphate loops because the total amount of scale present was so small.
Overall, the order of increasing scale accumulation was the same as the order
for increasing corrosion rate, with apparently a somewhat larger fraction of
the scale remaining attached to the coupon in the phosphate loops than in
those that contained no ortho-phosphate. For instance, after 400 days
exposure, the cumulative weight losses in the control, low-phosphate, and
high-phosphate loops were in the approximate ratio 14:2.4:1. The
corresponding ratio for scale weights was 5.8:1.8:1.

Chemical analysis of the scale after several hundred days exposure indicated
that the scale in the low phosphate loop was about 73 percent copper, which
is in the same range as the copper content of the scales in the phosphate-
free systems. Only 3 percent of the scale weight could be accounted for as
o-P04 , the remainder apparently being oxide ions or bound water. The scales
from the high phosphate loop contained only 48 percent copper by weight.
Cupric phosphate, ^(PO^, is 50 percent copper, so this appeared to be a
likely candidate for the major compound forming the scale. However, the
scale was only 4 percent P04 by weight, compared to a theoretical value of 50
percent for ^(PO^. It is not clear what compound other than bound water
could make up the remainder of the scale, if these analyses are correct.

SHORT-TERM WATER QUALITY EFFECTS ON AGED COPPER CORROSION

As noted above, electrochemical measures can give an indication of the


corrosion rates at a given instant in time, making them ideal for studies of
the short-term effects of changes in water quality. The qualitative
conclusions based on weight-loss data about the effect of exposure time and
ortho-phosphate addition on copper corrosion were equally evident using
electrochemical tests (Figure 38). In the next series of tests,
electrochemical methods were used to investigate the sensitivity of copper
corrosion rates to transients in dissolved oxygen, pH, and ortho-phosphate
concentration.
134
1.0-

o pH 6.0 (Water 1)
pH 8.0 (Water 2)
0.8- a pH 8.0, PO4 1 mg/l (Water 6)
pH 8.0, PO4 5 mg/l (Water 7)

0.6-

Instant.
Corrosion
Rate (MPY)
0.4-
U)

0.2-

0.0
50 100 150 200 250
Exposure Period (Days)

Figure 38. Comparison of Instantaneous Corrosion Rates on Copper Coupon Surfaces Under
Various Water Quality and Inhibitor Conditions
Dissolved Oxygen

The dependence of copper corrosion on the available oxygen supply is


displayed in Figure 39. Here the relationship between dissolved oxygen
concentration and corrosion current is demonstrated for a relatively young
surface. Increases in oxygen concentration up to about 4 mg/L caused the
corrosion rate to increase, but further oxygenation of the water supply had
little additional effect.

In most distribution systems, the dissolved-oxygen concentration of delivered


water is near the saturation value. Controlling this parameter is neither
desirable from a water quality standpoint, nor practical from an operational
viewpoint. Thus, for all practical purposes, copper corrosion can be
considered insensitive to dissolved-oxygen values in most water supply
systems.

Chlorine Residual

Similar experiments were carried out where the free chlorine residual was
varied over the range of values that may be encountered in a water supply
system. The pH was maintained at 7.0 in all these experiments. The results,
displayed in Figures 40 and 41, indicate that even small chlorine residuals
can dramatically increase copper corrosion currents on both new and aged
surfaces. However, if 1 mg/L o-P04 -P is present during aging, the effect of
free Cl is dramatically diminished. These tests also showed that, as
expected, the corrosion potential is sensitive to chlorine residual, becoming
more oxidizing with increasing chlorine. For both fresh and aged surfaces,
the effect of as little as 0.1 to 0.2 mg/L free chlorine was noticeable, and
the corrosion current continued to increase approximately linearly with
increasing chlorine throughout the range of residuals typically applied in
water supply systems (up to about 1 mg/L).

136
260
m Cathodic Tafel Slope

* Anodic Tafel Slope

Ba.Bc
(mV/dec.)

20

0.5

0.4.

lo 0-3
((iA/cm2)
0.2

0.1.

0.0
0 4 6 10
Dissolved Oxygen (mg/L)

Figure 39. Variation in Corrosion Current and Tafel Slopes as a Function of Dissolved Oxygen
(pH 7.5, Temp 23 C, Conductivity 80 |iS, Flow Velocity 1 fps, Surface Exposure 5 Days)
220.

180-

Re
(niV/dcc.) 140 -"

100-

60

160

120-

Ra 80 <
(mV/dec.)

40-
00

0.

4-

3-
lo
2.

1, G> Fresh Cu Surface


* Aged Surface

0.00 0.25 0.50 0.75 1.00 1.25 1.50


Free Chlorine Residual (mg/L)

Figure 40. Variation in Corrosion Current and Tafel Slopes on Copper Surfaces as a Function of Free
Chlorine Residual (Conductivity 100 ^S, DO 10 mg/l)
220.

180.
Be
(mV/dec.)
140.,

100

140,

100.
Ba
(mV/dec.)
60.
u>
vO

20

3.0
2.5-
Aged Surface, Water 2
260 Day Exposure
2.0 :
Aged Surface, Water 6
lo 1.5-i Phosphate Inhibited
(u.A/(cm2) 220 Day Exposure
1.0-

0.5
0.0-9

0.00 0.25 0.50 0.75 1.00 1.25 1.50


Free Chlorine Residual (mg/L)
Figure 41. Chlorine Residual Effects on Well Aged Phosphate Based Scales (pH 7.0,
Conductivity 100 |iS, DO at Saturation)
Short-term effects of pH on corrosion rates were also studied
electrochemically. Although the hydrogen ion is not electrochemically active
in the corrosion process (i.e., hydrogen reduction is never expected under
drinking water conditions), pH can influence copper corrosion in several
ways.

First, pH alters the distribution of species in solution, thereby altering


the equilibrium potential of all the redox reactions that are taking place.
For instance, by changing the activity of hydroxide ion and the speciation of
dissolved copper and hypochlorite, a shift in solution pH changes the
potentials of the major anodic and cathodic corrosion reactions.

Second, pH is likely to affect the stability and protective qualities of the


passivating scales that form on copper surfaces. There is no doubt that
significant decreases in pH can cause the protective scale to dissolve, and
although the effects of small changes in pH are more difficult to predict, it
is reasonable that they might affect the scale's properties in some way.

The short-term response of copper corrosion currents to changes in pH is


shown in Figure 42. The experimental progression was from higher to lower pH
values, with a neutral pH relaxation period between successive measurements.
Corrosion currents start increasing significantly at pH values below about 8
on fresh surfaces, increasing by at least a factor of 2 with each unit
decrease in pH to 6. The increase in corrosion current of the aged sample is
much less dramatic.

Because the readings were made after only a few hours of equilibration to the
new solution conditions, it is not surprising that any protective scale that
had developed on the aged surface retained its protectiveness during the
test. The slight increase in corrosion current that was observed in this
system may have been due to partial loss of the scale or to the fact that the
driving force for the oxygen-reduction reaction is increased as pH is
decreased. By contrast, the response of the relatively young, unprotected
surface, which had been exposed to Seattle tap water for approximately three
140
220
B Fresh Cu Surface
ISO- Aged Surface, Waier 2

Be 140- a a + a
(mV/dec.)

100-

60
160.

120-

80- * * * *
Ba Q D D
(mV/dec.)
40-

3-

lo 2-
(\i\lcm2)
1-

0.
5 6 7 pH 8 9 10
Figure 42. Variation in Corrosion Current and Tafel Slopes on Copper Surfaces as a Function of pH
(Conductivity 100 nS, Dissolved Oxygen 10 mg/1)
days, approached that of a fresh surface. This suggests that if low pH
conditions persist in a system any protective scale that formed previously
might dissolve, leading to dramatically increased corrosion currents.

SHORT-TERM TESTS USING FRESH COUPONS

Several short-term experiments were conducted using fresh coupons. All these
experiments started with a short exposure to pH 7.5 water containing no
inhibitor, after which either pH or o-P04 concentration was changed.

In two experiments, the initial exposure to pH 7.5-water containing no


inhibitor lasted for 15 to 20 hr. During this time, the corrosion current
appeared to decrease from about 1.0 to about 0.5 /jA/cnr (Figure 43). When pH
was then decreased to 5.5, there was an immediate, dramatic increase in
corrosion current by a factor of 5 or more. This increase persisted for at
least several hours. After about 15 hr at this condition, pH was raised to
6.0 for about 15 hr, and then to 6.5 and finally 7.0. The corrosion current
decreased more or less steadily after the pH had increased to 6.0, and
attained a steady value of about 0.5 /iA/cm2 at pH 7.0 for at least 100 hr.

In another experiment, 1 mg/L o-PO^-P was added to the circulating solution


after 16 hr, and pH was maintained at 7.5 for an additional 100 hr (Figure
44). Within 10 hr after this change, the corrosion current had decreased by
about 40 percent, and it remained steady or decreased slightly for the
remainder of the experiment. After 100 hr exposure to 1 mg/L P, the
corrosion current was approximately 0.3 /iA/cm2 .

The short-term effect of ortho-phosphate addition on corrosion at pH 7.5 is


somewhat clearer in Figure 45. In this case, the initial exposure to a
phosphate-free solution was for nine days, long enough for a reasonably
stable corrosion current of about 0.5 /iA/cm2 to be established. The coupon
was then exposed to 1 mg/L o-P04 -P for seven days, after which the ortho-
phosphate was once again removed from the water supply. Within a day of
adding the phosphate, the corrosion current decreased dramatically, and by
the fourth day the current had decreased by about 75 percent, where it
appeared to stabilize. When the inhibitor was eliminated from the water, the
142
5.5 6.0

7.5 6.5 7.0

2-
Corrosion
Current
(uA/cm2)
1-
.
0
0 50 100 150 200 250
Exposure (hours)

Figure 43. Response of Copper Coupon to pH Transients PO4 - P = 0 PPM.


Values Above Graph Indicate pH.
1.5

lmg/I P-PO4 Added

1.0 -

Corrosion
Current
(uA/cm2)

0.5 -

0.0 1 50 100 150


Hours

Figure 44. Corrosion Current for Coupon Exposed to 1 mg/l PO4 at Indicated Time (pH 7.5, Initial PO4 0 mg/l)
260

220 -
System Flush to Remove
Orthophosphate Inhibitor
180 -
CD
El

Ba, Be
(mV/dec.)
Addition of 1 mg/L of
Orthophosphate
100 -

60- EI Anodic Tafel Slope


* Cathodic Tafel Slope
20-

0.6

0.5-

0.4-

lo Oo "
(HA/cm2)
0.2-

0.1 -

0.0 i 1
0 10 15 20 25 30
Experimental Progression (exposure days)
Figure 45. Effect of Orthophosphate Inhibitor Addition on Corrosion Current and Tafel Slopes for a
Relatively Fresh Copper Surface (pH 7.5, Temperature 23 C, Conductivity 80 [iS,
DO 8.2 mg/I, Flow Velocity 1 fps)
corrosion current began to increase once again, but not as rapidly as it had
decreased. One week after the inhibitor was eliminated, the corrosion
current had increased to about half of the initial, pretest value.

In another experiment, the initial exposure to inhibitor-free pH 7.5 water


lasted only 1 hr, after which 5 mg/L o-P04 -P was added. Over the next 35 hr
the corrosion rate decreased from about 0.75 to about 0.5 /iA/cm2 , comparable
to the change in the system where only 1 mg/L o-P04 -P had been added (Figure
46). When the pH of this system was then lowered to 5.5, there was an
immediate increase in the corrosion rate. However, in contrast to the case
when the pH had been lowered in the absence of phosphate, in this case the
corrosion rate started decreasing again within a few minutes. Within an
hour, even though the pH was still 5.5, corrosion had slowed considerably,
and it continued to abate over the next 60 hr. From this time on, corrosion
rate decreased gradually, and the pH was raised to 7.0 in increments.
o
Eventually, the rate had decreased to less than 0.5 /iA/cm . It is not clear
how much of this change can be attributed to the pH change and how much
simply to exposure time. However, it is clear that the continuous exposure
to ortho-phosphate inhibitor significantly mitigated the short-term effect of
exposure to an acidic solution.

This latter observation was confirmed in an experiment investigating the


effect of a transient pH change on corrosion rates of coupons with pre
existing scales. In this case, two coupons that had been exposed to
solutions containing 5 mg/L o-P04 -P were subjected to a pH change from 7.0 to
5.9. In one experiment, the ortho-phosphate inhibitor was simultaneously
removed from the solution (Figure 47). In both cases, the corrosion rate
increased, and in both it appeared to approach a steady value between 1 and 2
/iA/cm2 after one day. However, the pattern of the approach to this stable
value was very different in the two experiments. When the pH was changed
while ortho-phosphate addition was continued, the increase was very gradual.
By contrast, when the phosphate inhibitor was eliminated at the same time the
pH was changed, the corrosion rate increased rapidly and peaked at a value of
nearly 4 /iA/cm2 after 2 hr. Thereafter, it gradually decreased to the stable
value.

146
I 7.5 5.5 5.8 6.6 7.0
2.0

1.5-

Corrosion
Current 1.0-
-~j (uA/cm2)

0.5-

0.0
50 100 150 200 250
Exposure (hours)

Figure 46. Response of Copper Coupon to pH Transients; PO4 - P - 5 mg/L


220

180-

Be 140-
(mV/dec.)

100-

60

160

120-

Ba 80-
(mV/dec.) D
*

40-

Preformed Scale
PO4 soln. (5 mg/L)
Preformed Scale
lo No Inhibitor
(|iA/cm2)

-12 -10 -8 -6-4-202468 10 12 14 16 18 20 22 24 26


Experimental Progression (hours)

Figure 47. Effects of pH Shift on Phosphate Inhibitor Scale (pH Maintained at 7.0 until Time 0 then
Shifted to 5.9, (Conductivity 100 nS, Dissolved Oxygen 10 mg/l)
Taken together, these results suggest that a significant acid transient will
cause a corresponding increase in copper corrosion in any system. This may
be related to partial or complete dissolution of any protective scale layer
that had previously built up. However, the extent and duration of the
corrosion response can be mitigated substantially by the presence of ortho-
phosphate. On fresh coupons, the pH change caused a rapid initial increase
in corrosion rate regardless of the presence or absence of phosphate.
However, in the presence of of ortho-phosphate, this initial increase was
rapidly reversed; when phosphate was absent, this rapid increase persisted.
On coupons with pre-existing protective scales, lowering pH in the presence
of ortho-phosphate led to a gradual increase in corrosion, while lowering it
in the absence of phosphate caused a much larger short-term increase in
corrosion and a comparable long-term change.

SUMMARY

To summarize these results, copper corrosion rates in our tests were


remarkably steady. The only loop in which they changed substantially during
the course of the experiments was the pH 6.0 loop, in which corrosion rate
decreased significantly after about 50 days exposure. Corrosion rates in the
pH 7.2 high-carbonate water and the pH 8.0 tap water were about equal to each
other, and to the post-50-day rate in the pH 6.0 water. The absence of a
strong dependence of copper corrosion rate on pH in the range 6.0 to 8.0 is a
surprising result and is generally at odds with prior literature results.
The pH 8.0 low-carbonate water had a significantly lower corrosion rate than
any of the other three phosphate-free waters.

One or five mg/L o-PO^-P lowered copper corrosion rates very dramatically.
The effect was apparent from the start of the test, and after a few months of
exposure, corrosion of copper in these loops nearly ceased altogether.

The only mineral that could be positively identified in the scales in any of
the phosphate-free loops was cuprite, Cu0. A small fraction of the corroded
copper remained attached to the coupon walls in these loops, with the
fraction being somewhat greater at higher than at lower pH. Scale in the
5 mg/L P loop had a Cu concentration consistent with (^(PO^, but the scale
149
contained much less P than would be required for that mineral to represent a
significant fraction of the scale. Apparently, only a small fraction of the
scale need be 113 (PO^ or some other P-containing mineral to confer good
corrosion-protection properties on the scale.

Dissolved oxygen has a relatively minor effect on the short-term corrosion


rates of copper, but free residual chlorine can have a significant effect,
with corrosion rate increasing along with Cl residual. The effect of Cl is
much less pronounced if a P-containing scale is in place on the pipe surface.

Phosphate has been shown to protect copper when dosed at 1 or 5 mg/L as P at


a pH of 8. Short term tests showed that protection developed at pH 7.5 and
deteriorated at pH 5.9. The tests did not establish clear dosage and pH
boundaries for the development of a protective scale.

pH transients can have a significant effect om copper corrosion rates if


there is no existing protective scale. Aged coupons respond to an acidic pH
change by increasing corrosion rate, but the rate may return to a lower value
quite quickly, especially if o-P04 inhibitor is added continuously during the
pH transient.

150
CORROSION OF GALVANIZED PIPES

CUMULATIVE WEIGHT LOSS

Plots showing the weight loss of galvanized coupons as a function of exposure


time in the various loops are presented in Figures 48-50. By far, the most
dramatic result is the extremely high rate of weight loss in the Seattle tap
water at pH 6.0. The corrosion rate in this loop was relatively steady for
nearly two years, and it was three to eight times greater than the corrosion
rate in any of the other loops throughout the test period. Of loops
containing Seattle tap water, those adjusted to pH 8.0 with low alkalinity,
to pH 7.2 with high alkalinity, or to pH 8.0 with 5 mg/L o-P04 -P had
corrosion rates that were comparable to or slightly lower (depending on the
exposure period chosen for comparison) than those loops containing tap water
at pH 8.0 with no inhibitor or 1 mg/L PO^-P added. Overall, while alkalinity
or phosphate might have a marginal effect in slowing corrosion, it appears
that maintaining pH in the range of 7.4-8.0 is the most useful engineering
control to reduce corrosion of galvanized plumbing in soft waters.

While corrosion was faster in the low-pH loop than in the other loops from
the inception of the test, the main factor that distinguished this loop was
that its corrosion rate remained constant, while corrosion rate declined
significantly in all the other loops after a weight loss of about 250-300 mg.
Each coupon contained about 500 mg Zn in the galvanized layer, and therefore
it is not possible to relate the change in corrosion rate to complete loss of
this layer. However, the consistency of the weight loss at the time when the
corrosion rate changed suggests that some significant factor changed when
about half of the Zn had corroded. Perhaps at this point portions of the
underlying Fe surface are being exposed, and the lower corrosion rate
reflects the loss of Zn surface area in conjunction with a change in the
electrochemical characteristics of a Zn/Fe surface versus a pure Zn surface.
In any case, regardless of the mechanism involved, in all the loops other
than the one at pH 6.0 the average corrosion rate slowed substantially after
about 250 mg had been lost from each coupon. Because of this, the total
weight loss in the pH 6.0 loop was on the order of twice that in the other
loops after six months but about 7 to 10 times greater after 18 months.
151
5000

4000-

3000-
o pH= 6.0 (Water 1)
Weight Loss
(rag) pH= 8.0 (Water 2)

pH= 8.0 Low CO3 (Water 4)


2000-
o a PH= 7.2 High CO3 (Water 5)
80

1000-

200 400 600 800


Exposure (days)

Figure 48. Weight Loss for Galvanized Pipes Exposed to Various


Water Qualities in the Absence of Phosphate
1000

800-

600-
Weight
Loss o pH= 8.0 (Water 2)
(mg) pH= 8.0 Low CO3 (Water 4)
a o o
400- Op
o D n pH= 7.2 High CO3 (Water 5)
o

200-
& an 0 *

200 400 600 800


Exposure (days)

Figure 49. Weight Loss for Galvanized Pipe Exposed to Different


Water Qualities in the Absence of Phosphate
(Same data as Figure 48 on expanded scale)
1000

800-

600-
o pH- 8.0 (Water 2)
Weight
Loss pH- 8.0 PO4- 1 mg/L (Water 6)
(mg)
O D pH- 8.0 PO4- 5 mg/L (Water 7)
400-
o
o B
a

200- D B

cP
o
0 D

200 400 600 800


Exposure (days)

Figure 50. Weight Loss for Galvanized Pipe in the


Presence and Absence of Phosphate
SCALE CHARACTERISTICS

Scale Accumulation

The data characterizing the amount of scale that accumulated on the coupons
was more scattered than the weight-loss data, but in most cases trends within
data sets and distinctions between data sets were nevertheless discernible.
In the loop containing the most corrosive water (pH 6.0), scale accumulation
was unremarkable for 1 to 1 1/2 years and then increased tremendously (Figure
51). As shown previously, there is no corresponding change in the corrosion
rate. An obvious hypothesis is that the change in scale-accumulation rate
correlated with disappearance of the Zn layer, and that the accumulating
scale is primarily a hydrous iron oxide. The second part of this hypothesis
is consistent with the few analyses conducted during the course of this
research on the chemical composition of these scales (Table 18). None of the
scales contained less than 43 percent Zn except the scales from the pH 6.0
loop after 473 days exposure, which contained only 5 percent Zn.

In the two pH 8.0 loops with no added PO^, scale appeared to accumulate for
four to six months, after which a substantial portion of the scale was
released from the pipe surface (Figure 52). The change from a positive to a
negative accumulation rate corresponded reasonably well with the decrease in
corrosion rate (as determined by weight loss) described above. This change,
therefore, may be related to a switch from primarily Zn corrosion to
primarily Fe corrosion. However, there were substantial differences between
the long-term corrosion of the coupons from the pH 8.0 loops with no added
PO^ and the long-term corrosion of ungalvanized coupons. For instance, the
approximate steady-state scale weight for the last 200 days of the test in
the pH 8.0 loops is about an order of magnitude less than the scale weight
after 200 days of exposure of ungalvanized iron coupons to the same waters.
Furthermore, the Zn content of the scale in one coupon in the pH 8.0 low-
carbonate loop was 40 percent after 218 days, i.e., after the apparent
release of much of the previously accumulated scale. Finally, the rate of
weight loss from the galvanized coupons in the pH 8.0 loop was much lower
than in the pH 6.0 loop, while the weight-loss rates for nongalvanized iron
155
5000

o o

4000-

3000- o pH= 6.0 (Water 1)


Scale pH= 8.0 (Water 2)
Weight
(nig) pH- 8.0 PO4= 1 mg/L (Water 6)
2000- o pH= 8.0 PO4= 5 mg/L (Water 7)

1000-
D
a D D

Q fl
" oflo

200 400 600 800


Exposure (days)

Figure 51. Scale Weight Formed on Galvanized Iron Pipe Exposed


to Phosphate-Inhibited and Non-Inhibited Waters
400'

300-

Scale Weight o pH= 8.0 (Water 2)


Accumulation 200' o
(rag) o pH=8.0LowCO3(Water4)
GO
o o o pH= 7.2 High CO3 (Water 5)

o
100-
o o

0 o

200 400 600 800


Exposure (days)

Figure 52. Scale Weight Formed on Galvanized Pipe Exposed to Various Water
Qualities in the Absence of Phosphate
TABLE 18
Zinc Concentration in Zinc Galvanized Corrosion Scales
August, 1987

Zinc Concentration in
Dried Scale (Percent)

Loop 1 Zinc
473 days Zn 106 (zinc coating depleted) 5
140 days Zn 153 44
39 days previous analysis Zn 173 & 180 57

Loop 2 Zinc
140 days Zn 158 & 115 51
70 days previous analysis Zn 190 & 193 57

Loop 3 Zinc
141 days Zn 216-1 - 47
141 days Zn 216-2 47
175 days previous analysis Zn 172,175,162 49

Loop 4 Zinc
218 days Zn 165-168 48
135 days Zn 315 47
84 days previous analysis Zn 201 & 207 57

Loop 5 Zinc
316 days Zn 123-1 44
316 days Zn 123-2 44
178 days previous analysis Zn 160,161,162 43

Loop 2 Iron
Fe 301-2 1

158
coupons were about the same at the two pH values. Taken together, these
observations indicate that even if most of the Zn did oxidize early in the
test period, Zn continued to play an important role in controlling the
corrosion rate and restricting the accumulation of scale in these waters.

In the pH 7.2 high-carbonate loop, scale accumulation was always quite low,
and there was no evidence of significant scale release at any time. If
anything, it appeared that scale accumulation reached a steady-state value
rather quickly, and then underwent a small but consistent increase to a new
steady-state value around the time that the corrosion rate decreased.

In both phosphate-containing loops, scale weight attained a steady-state


within about six months that was maintained thereafter (Figure 53). In the
5 mg/L-P loop, the attainment of the steady-state correlated well with the
reduction in corrosion rate; in the 1 mg/L-P loop there may have been a
similar correlation, but the data are inconclusive. The absolute amounts of
scale in these loops were significantly greater than any of the other pH 8.0
loops, and the amount in the 5 mg/L-P loop was about twice that in the
1 mg/L-P loop.

Scale Mineraloqical Identification

The scales collected from the coupons generally had weak x-ray diffraction
patterns, if they had any at all. The high-alkalinity loop scales had x-ray
peaks suggestive of a mixed zinc hydroxide/carbonate solid known as
hydrozincite, but no positive identification could be made. The only other
scales that had any x-ray peaks at all were those from the pH 8.0, 5 mg/L o-
P04 -P loop. These scales were positively identified as a hydrated zinc
phosphate known as hopeite. Furthermore, this scale contained 14 percent P,
which matches the expected value for P in hopeite quite well. The coupons
exposed to water containing 1 mg/L o-P04 -P also contained a significant
amount of P (around 8 percent) but less than that expected for pure hopeite.

159
1000

D
D
D
800- n
ca
D

a
600- a
o pH= 8.0 (Water 2)
Scale Weight
Accumulation pH= 8.0 PO4= 1 mg/L (Water 6)
(mg)
a pH= 8.0 PO4= 5 mg/L (Water 7)
400-

200-
o o
o o
o o o o
oo

200 400 600 800


Exposure (days)

Figure 53. Scale Weight Formed on Galvanized Pipe


in the Presence of Phosphate
ELECTROCHEMICAL MEASUREMENTS ON GALVANIZED SURFACES

Validation

Electrochemical corrosion measurement techniques on galvanized surfaces were


nearly identical to those employed for copper. The main difference was that
the standard galvanized coupons were slightly larger (6 cm^) than the copper
coupons. There was significantly more variation of the surface electrical
potential measured on the zinc than on the copper surfaces. This is probably
due to a modest degree of nonuniformity in the corrosion process, which
results in uneven distribution of corrosion scale. Surface potential
measurements often vary by as much as 30 mV in scans on well-aged galvanized
coupons; comparable measurements on copper coupons vary less than 5 mV. To
standardize the measurement process, an exploratory surface scan was
conducted prior to each polarization sequence, and polarization readings were
then taken in an area where the surface potential represented an average of
the high and.low extremes.

Comparisons of electrochemical and gravimetric corrosion-rate measurements on


galvanized pipe indicate that the multipoint polarization test sequence gives
consistently low estimates of actual corrosion rates. Figure 54 shows
correlation of the electrochemical corrosion-rate measurements with the
gravimetric technique as a function of exposure duration. The discrepancy
between the two measurements is approximately 55 percent initially and
increases to about 65 percent over the first -30 days of exposure, after
which it remains approximately steady. The differential between the measures
remains regardless of the rate of metal loss. Figure 55 shows that a
comparable discrepancy occurs on a variety of galvanized coupons corroding at
substantially different rates. It was concluded that the multipoint
polarization technique may presently be inadequate as an indicator of
absolute corrosion rates of galvanized-pipe coupons, but that the technique
does have value as an indicator of relative corrosion rates among coupons
and, over time, for a given coupon.

161
200

ISO-

160-
Be
(mV/dec.)
140-

120-

100

200

180.

160-
Ba D
(mV/dec.) m a
140-
Ov
10
120-

100.
80
70.

Electrochemical 60-
Measure as a 50.
Percent of
Gravimetric 40^
30-
20
0 50 100 150 200
Exposure Period (days)
Figure 54. Electrochemical Measurement Variation as a Function of Exposure Time. Error Bars Represent
a Single Standard Deviation Around the Mean (pH 7.0, Conductivity 60 ^S, DO 10 mg/l)
Loop#1 Conditions

(10 day exp.)

Loop #2 Conditions

(8 day exp.)

Loop #5 Conditions

(14 day exp.)

Everett Dist. System


Electrochemical Measure
Weight Loss Values
(120 day exp.)

Corrosion Rate (expressed as mpy)

Figure 55. Comparison of Weight Loss and Electrochemical Corrosion Rate Measurements
Under Different Water Quality Conditions
Reduction Couples

Dissolved Oxygen

Metallic zinc is easily oxidized, which is why a galvanized coating provides


anodic protection to the underlying steel surfaces:
Zn = Zn2+ + 2e~ E -0.76 V
Its high oxidation potential offers the opportunity for zinc oxidation to be
coupled to several different reduction reactions, including reduction of
dissolved oxygen, chlorine, and hydrogen ion. However, electrochemical
measurements on galvanized surfaces indicate that neither dissolved oxygen
concentration nor that of residual chlorine strongly influences the rate of
corrosion, at least not to the degree observed on copper surfaces. Figure 56
shows the relationship of electrochemical corrosion-rate measures to
dissolved oxygen concentration for a relatively young zinc surface.
Corrosion currents are not substantially affected, even when the dissolved
oxygen is completely sparged from the system. Further evidence of noneffect
is manifested in the constant Tafel slopes over the full dissolved-oxygen
concentration range; on copper surfaces, any shift in the oxygen
concentration was accompanied by significant changes in Tafel slope values.

The only electrochemical measure that changed as a function of DO was the


surface potential of the test coupon, which underwent a negative shift of 200
mV as oxygen was stripped from the system. This change may reflect the
absence of adsorbed oxygen at the metal surface, but surprisingly does not
appear to affect the corrosion process. Figure 57 presents an Evans diagram
for the galvanized surface, using the measured surface potential (adjusted
relative to the standard hydrogen electrode) and the calculated Tafel slopes.
The indicated oxidation potential for zinc is based on the approximate
concentration of dissolved zinc in the recirculation test loop; the hydrogen-
and oxygen-reduction couples indicated are both adjusted to reflect the test
pH. Although the potential driving force for a zinc oxygen couple is
substantially higher than that for the zinc-hydrogen couple, the cathodic
Tafel slope does not intersect the line of the oxygen-reduction couple within
the range of a reasonable exchange current density. This, combined with the
164
-700

-800 -

-900 -
Surface
Polenlial -1000-
Ecorr (mV)
-1100-

-1200
180
Anodic TafeJ Slope
a Cathodic Tafel Slope
160-
a a o
Ba.Ik
(mV/dcc.)
140-

120
2.0.

1.5.
lo
(H/V/cm2)

0.5.

0.0
02468 10
Dissolved Oxygen (mg/L)
Figure 56. Electrochemical Measurement Variation on a Galvanized Surface as a Function of Dissolved
Oxygen (pH 7.0, Conductivity 100 ^S, Surface Exposure 20 Days in Water 2)
800 _
l/SDe HzD + 2e- = 20HT
Zn Polarization Diagram
600 pH 7.0
conductivity = 100 uS
dissolved oxygen = D.O ng/l
400 _

wo _

U
Surface
-400 Potential
Deaerated
Water
-00 _

-MO _

-1000

-1MO _

L
0.00001 0.0001
Current Density

Figure 57. Evans Diagram for Polarization of a Galvanized Coupon Surface


evidence of Figure 56, suggests that in contrast to the situation for ~oe>"
corrosion, dissolved oxygen plays no significant role in controlling the race
of corrosion of galvanized surfaces.

Chlorine

The effect of chlorine on corrosion of galvanized surfaces is similar to that


of dissolved oxygen. Although the zinc/chlorine couple has a higher
electrochemical potential than either the zinc/oxygen or zinc/hydrogen
couple, free chlorine has only a minor effect on the overall
electrochemically-measured corrosion rate. This is demonstrated in
Figure 58, where a galvanized test electrode was contacted with successively
higher concentrations of free chlorine, extending to values beyond those
normally employed in a distribution network. The solution pH was maintained
at 7.0 in all these experiments. There is no measurable effect of chlorine
on corrosion current until chlorine concentrations exceed 1.5 mg/L, and then
only to a minor degree. Surface potential is not substantially altered by
the presence of the chlorine. An increase in anodic Tafel slope was observed
at the highest chlorine concentration and was verified by repeated measures.
The shift is surprising, since a change in the concentration of the
constituent undergoing reduction is expected to affect the anodic Tafel slope
much more than the cathodic slope. In contrast to corrosion on copper
surfaces, the presence of a free chlorine residual is not rate controlling
for galvanized metal, nor does it substantially alter the electrochemical
surface properties.

pH

pH alteration between pH 5 and 7 does affect galvanized surface corrosion.


Figure 59 displays pH effects on several electrochemical parameters for a
moderately aged (60 days, pH 8.0) galvanized coupon. Altering the pH has no
appreciable effect on Tafel slopes, but does induce moderate increases in
corrosion current as the pH drops below 7.0. The surface potential is
sensitive to pH and decreases as the pH increases. This is probably
attributable to a decreasing hydrogen reduction potential as the

167
-700

-800-

Surface -900 ^
Potentia l
Ecorr(mV) 1000-
-1100-

-1200
180

160 J
Ba, Be
(mV/dec.)
140-
00 Anodic Tafel Slope
* Cathodic Tafel Slope

120
2.0

1.5.

lo 1.0-
(uA/cm2)
0.5-

0.0
0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
0.2 0.4
Free Chlorine Residual (mg/L)
Figure 58. Electrochemical Measurement Variation on a Galvanized Surface as a Function of Free Chlorine
Residual (pH 7.0, Conductivity 100 ^S, DO at Saturat ion, Surface Exposure 60 Days Water 2)
-700

-800

Surface
-900
Potential
Ecorr(mV)

-1100

-1200
180
Anodic Tafel Slope
Cathodic Tafel Slope
160-
Ba/Bc
(mV/dec.)
ON 140-

120
2.0

1.5 J
lo
(uA/cm2) i.o:
0.5-

0.0
10

Figure 59. Electrochemical Measurement Variation on a Galvanized Surface as a Function of pH.


Coupons Exposed to Water 2 for 60 Days Prior to the Test. Conductivity = 100 \iS,
DO at Saturation
concentration of hydrogen ion diminishes. Between pH 7.0 and 8.5, none of
the relevant electrochemical parameters changes significantly.

In evaluating these data, it must be remembered that during the 60-day


exposure to pH 8.0 water prior to the test, each coupon corroded
substantially (-200 mg weight loss, 150 mg scale accumulation), and that the
electrochemical tests are very short-term by comparison. The absence of a
large instantaneous effect of lowering pH to <6.0 undoubtedly reflects the
effects of the built-up scale, and a longer exposure to these conditions
would be expected to have a much more dramatic effect.

Overall, the pH sensitivity of the surface potential and corrosion current,


coupled with the lack of effect attributable to either dissolved oxygen or
chlorine residual, suggests that hydrogen ion reduction may be the principal
reduction reaction associated with zinc oxidation. Furthermore, corrosion
rates of galvanized pipe surfaces appear more stable and less susceptible to
environmental change in water quality than those of copper.

SUMMARY

The corrosion rate of zinc-galvanized pipe coupons was relatively insensitive


to all water quality variables other than pH. When exposed to pH 6.0, soft
water, the coupons corroded many times faster than they did in waters at pH
7.2 or 8.0, whether ortho-phosphate was present or absent. In the pH 6.0
loop, the corrosion rate was initially about twice as rapid as in the other
loops, and it remained very steady for the two-year duration of the test. By
contrast, the corrosion rate decreased substantially in all the other loops
after six to nine months, which corresponded to the time when the weight loss
equalled about half the weight of the original galvanized layer.

The pattern of scale accumulation differed significantly from loop to loop.


In the pH 6.0 loop, scale accumulated at a moderate rate for 1 1/2 years,
after which the rate increased very dramatically. The Zn content of the
scale from one coupon after more than 1 1/2 years exposure was much lower
than that from coupons exposed less than 1 1/2 years, suggesting that the

170
increase in scale accumulation rate after 1 1/2 years may have been due to
corrosion of the underlying iron.

By contrast, scale accumulation was slight in the pH 7.2 high-carbonate loop.


In two loops at pH 8.0, scale accumulated and was then released from the
surface. Scale accumulation in the loops containing ortho-phosphate was
moderate and attained a steady state within about six months.

The electrochemical studies using galvanized coupons largely supported the


findings based on weight loss. That is, the corrosion rate is sensitive to
changes in pH but not to changes in concentration of other water quality
parameters, including relatively strong oxidants such as dissolved oxygen and
free chlorine. Thus, in soft waters at least, pH adjustment appears to be
the most straightforward way to control corrosion rates of galvanized pipes.

171
SUMMARY. CONCLUSIONS. AND RECOMMENDATIONS

This research project investigated the corrosion rates of copper, mild steel,
and zinc-galvanized pipe coupons as a function of the chemical
characteristics of the water to which they were exposed. The coupons were
short sections of pipe typical of the kind that might be employed in a
household plumbing system. Several coupons of each material were placed in a
holder and exposed to water of controlled composition over a period ranging
from a few weeks to more than two years. After a predetermined exposure
period, the coupons were removed from the holding assembly and analyzed for
weight loss. Additionally, corrosion scales that had developed were
characterized by visual examination, by x-ray diffraction to identify their
mineral forms, and by chemical analysis. Overall, more than 1,000 pipe
coupons were subjected to corrosion testing.

MILD STEEL

The corrosion rates of mild-steel pipes and the composition and structure of
corrosion scales were related to water quality variables. The most
protective iron corrosion scale was formed in a high-alkalinity, moderate-pH
water (alkalinity 400 mg/L as CaC03 , pH 7.2). Pipe sections coated with this
scale had very low corrosion rates (about 2 mpy or less). The protective
scale was thin, hard, dark, and very adherent to the metal surface. Siderite
was found in this scale by x-ray diffraction analysis. This was the only
scale where siderite was detected. Wet chemical analysis of this corrosion
scale before and after air-drying showed that 5-10 percent of the iron was in
the ferrous state. Small amounts of carbonate were also found in both wet-
and dried-scale samples, and it appeared that most of the ferrous iron in
this scale was present as siderite.

Waters with alkalinities of 20 mg/L as CaC03 or less produced significantly


higher corrosion rates than the high-alkalinity (400 mg/L as CaC03 ) water.
Iron corrosion rates were generally between 10 and 20 mpy in all of the low-
alkalinity waters studied.

173
Scales formed in the low-alkalinity waters were thick and loose, consisting
of large tubercles covering shallow pits. After nine months to one year of
exposure to low-alkalinity water, small amounts of hard, tenacious scale were
formed under the soft, loose scale layers. Goethite (a-FeOOH) was the only
mineral detected in the x-ray diffraction analysis of dried low-alkalinity
scales. However, goethite was not the only compound present in these scales,
since ferrous iron and carbonate were detected in scale samples before they
were air dried.

Addition of 5 mg/L o-P04 -P at pH 8.0 reduced corrosion rates and scale


accumulation by about two-thirds compared to P04 -free controls. Solutions
containing 1 mg/L o-PO^-P were about as corrosive as the controls.

In general, increases in corrosion rates correlated with decreases in buffer


capacity. In two waters, the buffer capacity was decreased by bubbling the
water with CC^-free air. Both of these waters were more corrosive than
unaltered pH 8.0 Seattle tap water. The pH 7.2 high-carbonate water had by
far the highest buffer capacity and was much less corrosive than any of the
low-alkalinity waters. The pH 6.0 Seattle tap water was the only exception
to the correlation between buffer capacity and corrosivity; it had a higher
buffer capacity than Seattle tap water, but was also more corrosive. At
first, the pH 6.0 water produced similar corrosion rates to the pH 8.0
Seattle tap water, but after a year of exposure the pH 6.0 water was more
corrosive.

High chloride ion concentrations may have accelerated corrosion in waters


where HC1 was added to adjust the pH. A high chloride ion concentration may
explain why buffer capacity failed to predict the relative corrosivity of the
pH 6.0 water. Likewise, the high corrosion rates in the pH 8.0, low
carbonate water may have been caused by its higher chloride ion content,
since its buffer intensity was only slightly less than the buffer intensity
of pH 8.0 Seattle tap water.

174
COPPER

Copper corrosion rates in these tests were remarkably steady. The only loop
in which they changed substantially during the course of the experiments was
the pH 6.0 loop, in which corrosion rate decreased significantly after about
50 days exposure. Corrosion rates in the pH 7.2 high-carbonate water and the
pH 8.0 tap water were about equal to each other, and to the post-50-day rate
in the pH 6.0 water. The absence of a strong dependence of copper corrosion
rate on pH in the range 6.0 to 8.0 is a surprising result and is generally at
odds with prior literature results. The pH 8.0, low-carbonate water had a
significantly lower rate than any of the other three phosphate-free waters.

Both 1 and 5 mg/L o-P04 -P lowered copper corrosion rates dramatically. The
effect of the phosphate was apparent from the start of the test, and after a
few months of exposure, corrosion of copper in these loops nearly ceased
altogether.

The only mineral that could be positively identified in the scales in any of
the phosphate-free loops was cuprite, Cu20. A small fraction of the corroded
copper remained attached to the coupon walls in these loops, with the
fraction being somewhat greater at higher than at lower pH. Scale in the
5 mg/L P loop had a Cu concentration consistent with Cu3 (P04 ) 2 , but the scale
contained much less P than would be required for that mineral to represent a
significant fraction of the scale. Apparently, only a small fraction of the
scale need be ^(PC^^ or some other P-containing mineral to confer good
corrosion protection properties on the scale.

Dissolved oxygen has a relatively minor effect on the short-term corrosion


rates of copper, but free residual chlorine can have a significant effect,
with corrosion rate increasing along with Cl residual. The effect of Cl is
much less pronounced if a P-containing scale is in place on the pipe surface.

pH transients can have a significant effect on copper corrosion rates if


there is no existing protective scale. Aged coupons respond to an acidic pH
change by an increase in corrosion rate, but the rate may return to a lower
175
value quite quickly, especially if o-P04 inhibitor is added continuously
during the pH transient.

GALVANIZED STEEL

The corrosion rate of zinc-galvanized pipe coupons was relatively insensitive


to all water quality variables other than pH in these tests. The coupons
corroded many times faster in pH 6.0 soft water than they did in pH 7.2 or
8.0 water, and this rate was not affected by either the presence or absence
of ortho-phosphate. In the pH 6.0 loop, the corrosion rate was initially
about twice as rapid as in the other loops, and it remained very steady for
the two-year duration of the test. By contrast, in all of the other loops
the corrosion rate decreased substantially after six to nine months, which
corresponded to the time when the weight loss from the coupon equalled about
half the weight of the original galvanized layer.

The pattern of scale accumulation differed significantly from loop to loop.


In the pH 6.0 loop, scale accumulated at a moderate rate for 1 1/2 years,
after which the accumulation rate increased very dramatically. In one coupon
that had been exposed for more than 1 1/2 years, the Zn content of the scale
was much lower than that from coupons exposed less than 1 1/2 years,
suggesting that the increase in scale accumulation rate after 1 1/2 years may
be due to corrosion of the underlying iron.

By contrast, scale accumulation was slight in the pH 7.2 loop with high
carbonate. In two loops at pH 8.0, scale accumulated and was then released
from the surface. Scale accumulation in the loops containing ortho-phosphate
was moderate and attained a steady state within about six months.

The electrochemical studies using galvanized coupons largely supported the


findings based on weight loss. That is, the corrosion rate is sensitive to
changes in pH but not to changes in concentration of other water quality
parameters, including relatively strong oxidants such as dissolved oxygen and
free chlorine. Thus, in soft waters at least, pH adjustment appears to be
the most straightforward way to control corrosion rates of galvanized p.ipes.

176
RECOMMENDATIONS

The intent of this project was to develop a data base which might lead to a
more fundamental and complete understanding of corrosion of common plumbing
materials. This has been accomplished, and both the data and the methodology
will be of value to utilities and corrosion-control practitioners. However,
no single study can investigate the corrosion reaction over the full range of
water qualities typical of drinking water sources or potable water. For this
reason, it is inappropriate to make specific recommendations regarding
corrosion control practices in a given system; it is still the case that
control strategies must be developed on a case-by-case basis.

However, when a water supply has characteristics similar to those of the


water tested, i.e., very low alkalinity and dissolved mineral content, it
appears that raising the pH is the most effective corrosion-control approach
for galvanized pipes, adding carbonate alkalinity and increasing buffer
intensity are most effective for mild-steel pipe, and adding ortho-phosphate
is most effective for copper pipes. These control strategies are independent
but compatible, and any one or all three could be implemented in a given
system as needed. The addition of ortho-phosphate to water could lead to
serious problems with algal growth in open reservoirs or could significantly
increase the P load on local wastewater treatment plants, and these drawbacks
must be considered.

Similar studies can be, and to some extent are being, undertaken to
investigate the effects of corrosion-control strategies on corrosion rates in
systems with significantly different source water quality. These studies
should be supported so that, over time, an understanding of corrosion
mechanisms can be applied properly, thus reducing the effort and uncertainty
that are now required to develop strategies to minimize corrosion. In the
past, the Langelier Saturation Index and other corrosion indices have been
used inappropriately, and hence unsuccessfully, to predict corrosivities of
water supplies. This study has helped to confirm and extend the siderite
model for corrosion protection in steel pipes. It is reasonable to expect
that more studies of the type described will allow scientists and engineers
to develop protocols for assessing corrosivity for water with different
177
characteristics and plumbing materials, and that these protocols will be
considerably more reliable and useful than those previously available.

178
REFERENCES

American Society for Testing and Materials, D2688-83 Method B, Machined


Nipple Test (1983).
American Society for Testing and Materials, D2688-83 Method C, Coupon Test
(1983).
American Society for Testing and Materials, G5-82, Std. Reference Method of
Making Potentiostatic Polarization Measurements (1982).
Bandy, R. and Jones, D.A. "Analysis of errors in measuring corrosion rates by
linear polarization." Corrosion, 32, 126 (1976).
Baylis, J.R. "Prevention of corrosion and red water." Jour. AWWA, 15, 598-
633 (1926).
Bockris, J.O. and Reddy, A.K.N. Modern Electrochemistry, Vol. 2.
Plenum/Rosetta, New York (1973).
Budesinsky, Bret W. "Determination of metallic and oxide copper in ores."
Analyst, 102, 819-824 (1977).
Bundy, L. G. and Bremner, J. M. "A simple titrimetic method for
determination of inorganic carbon in soils." Soil Sci. Amer. Proc., 36,
273-275 (1972).
Caldwell, D.H. and Lawrence, W.B. "Water softening and conditioning
problems." Ind. Eng. Chem., 45, 533 (1953).
Cruse, H. et al. "Internal Corrosion of Water Distribution Systems."
Internal Corrosion of Water Distribution Systems, AWWARF, Denver, Co
(1985).
Daniel son, M.J. "An evaluation of the three-point method to measure
corrosion rates." Corrosion. 38, 580 (1982).
Dexter, S.C., Moettus, L.N. and Lucas, K.E. "On the mechanism of cathodic
protection." Corrosion, 42, 598 (1985).
Dye, J.F. "Calculations of the effects of temperature on pH, free carbon
dioxide and three forms of alkalinity." Jour. AWWA, 44, 356 (1952).
Eliassen, R., Pereda, C., Romeo, A.J. and Skrinde, R.T. "Effects of pH and
velocity on corrosion of steel water pipes." Jour. AWWA, 48, 1005-1018
(1956).
Evans, U.R. The Corrosion and Oxidation of Metals. Edward Arnold Ltd.,
London (1960).

179
Feigenbaum, C., Gaylor, L. and Yahalom, J. "Microstructure and chemical
composition of natural scale layers." Corrosion. 34(2) 65-69 (1978).
Freneir, W.W. and Growcock, F.B. "Mechanism of iron oxide dissolution - A
review of recent literature." Corrosion, 40, 663 (1984).
Gerchakov, S.M., Udey, L.R. and Mansfield, F. "An improved method for
analysis of polarization resistance data." Corrosion. 37, 696 (1981).
Ghosh, M.M., O'Conner, J.T. and Engelbrecht, R. "Bathophenanthroline method
for the determination of ferrous iron." Jour. AWWA. 59(7), 897-905
(1967).
Grauer, R., Moreland, P.J. and Pini, G. A Literature Review of Polarization
Resistance Constant Values. National Association of Corrosion
Engineers, Houston (1982).
Hatch, G.B. Corrosion Inhibitors (C. Nathan, editor). National Association
of Corrosion Engineers, Houston, TX, pp. 114-125 (1973).
Hatch, G.B. "Protective film formation with phosphate glasses." Ind. Enqr.
Chem.. 44, 1775-1786 (1952).
Hausler, R.H. "Practical experiences with linear polarization rate
measurements." Corrosion. 33, 117 (1977).
Hedberg T. and Johansson, E. Protection of Pipes against Corrosion. Special
Subject 20, IWSA Congress (1984).
Hepburne, M.L. Corrosion in Piping Systems. Piping Handbook. McGraw Hill
(5th ed., 1965).
Herrera, C.E. and Hoyt, B.P. Seattle Distribution System Corrosion Control
Study, Vol. 2, Tolt River Water Pilot Plant Study. U.S.E.P.A Report,
No. EPA-600/S2-84-065 (1984).
Heyer, C. Ursache und beseitigung des bleiangriffs durch leitungawasser.
Verlag Paul Baumann, Dessau (1888).
Hoyt, B.P., Herrera, C.E. and Kirmeyer, G.J. Seattle Distribution System
Corrosion Control Study, Vol. 1, Cedar River Water Pilot Plant Study.
U.S.E.P.A Report, No. EPA-600/S2-82-026 (1982).
Hudson, H.E. and Gilcreas, F.W. "Health and economic aspects of water
hardness and corrosiveness." Jour. AWWA. 68, 201 (1976).
Ijsseling, F.P. "Application of electrochemical methods of corrosion rate
determination to systems involving corrosion product layers." British
Corrosion Journal. 21, 95 (1986).
Jankowski, J. and Juchniewicz, R. "A four-point method for corrosion rate
determination." Corrosion Science. 20, 841 (1980).

180
Jobin, R. and Ghosh, M.M. "Effect of buffer intensity and organic matter on
the oxygenation of ferrous iron." Jour. AWWA. 64(9), 590-595 (197";).
Karalekas, P.C., Ryan, C.R. and Taylor, F.B. "Control of lead, copper and
iron pipe corrosion in Boston." Jour. AWWA. 75, 92-95 (1983).
Kennedy Engineers. Seattle Internal Corrosion Study, Phases I, II, and III.
Tacoma, WA (1978).
Kolle, W. and Rosch, H. "Untershugen an rohrnetzinkrusteirungen unter
minerologischen gesichtspunkten." Vom Wasser. 55, 159 (1980).
Langelier, W.F. "Chemical equilibria in water treatment." Jour. AWWA. 8,
169 (1946).
Langelier, W.F. "The analytical control of anti-corrosion water treatment."
Jour. AWWA, 28, 1500-1521 (1936).
Larson, I.E. "Corrosion by domestic waters." Bull. 59, ISWS, Urbana (1975).
Larson, I.E. and Buswell, A.M. "Calcium carbonate saturation index and
alkalinity interpretations." Jour. AWWA. 34, 1667 (1942).
Larson, I.E. and Skold, R.V. "Corrosion and tuberculation of cast iron."
Jour. AWWA. 49(10), 1294-1302 (1957).
Lehrman, L., Schuldener, H.L. "Action of sodium silicate as a corrosion
inhibitor in water piping." Ind. Enq. Chem.. 44, 1765-1769 (1952).
Leroy, R.L. "Evaluation of corrosion rates from polarization measurements."
Corrosion. 31, 173 (1975).
Lowenthal, R.E. and Marais, C.V.R. Carbonate Chemistry of Aquatic Systems:
Theory and Application. Ann Arbor Science Publishers, Ann Arbor, MI
(1976).
Lumsden, J.F. and Szklarska-Smialowska, Z. "The properties of films formed
on iron exposed to inhibitive solutions." Corrosion. 34, 169-176
(1978).
McCauley, R.F. and Abdullah, M.O. "Carbonate deposits for pipe protection."
Jour. AWWA. 50, 1419-1428 (1958).
Merrill, D.T. and Sanns, R.L. "Corrosion control by the deposition of CaCOo
films." Jour. AWWA. 69, 634 (1977).
Millete, J.R. "Aggressive water: Assessing the extent of the water." Jour.
AWWA. 72, 276 (1980).
J.M. Montgomery Consulting Engineers. Internal Corrosion Mitigation Study.
Prepared for the Bureau of Water Works, Portland, OR (1982).

181
Muller, E.D. and Ritter, J.A. "Monitoring and controlling corrosion by
potable water." Jour. AWWA. 72, 286 (1980).
Neff, C.H., Schock, M.R. and Marden, J.I. "Relationships Between Water
Quality and Corrosion of Plumbing Materials in Buildings--Vol. 1.
Galvanized Steel and Copper Plumbing Systems." EPA/600/2-87/036A (1987).
Pourbaix, M. Lectures on Electrochemical Corrosion. Plenum Press, New York
(1973).
Reiber, S.H., Ferguson, J.F. and Benjamin, M.M. "Corrosion Control in the
Pacific Northwest." Jour. AWWA. 79, 71-74 (1987).
Richards, W.N. and Moore M.R. "Lead hazard controlled in Scottish water
systems." Jour. AWWA. 76, 60 (1984).
Rossum, J.R. and Merrill, D.T. "An evaluation of the calcium carbonate
saturation indices." Jour. AWWA. 75, 95 (1983).
Ryder, R. and Wagner, I. Corrosion Inhibitors in Internal Corrosion of Water
Distribution Systems, AWWARF, Denver, CO, 617-655 (1985).
Ryznar, J.W. "A new index for determining amount of calcium carbonate scale
formed by a water." Jour. AWWA. 36, 472 (1944).
Schock, M.R.and Buelow, R.W. "The behavior of asbestos-cement pipe under
various water quality conditions: Part 2, Theoretical considerations."
Jour. AWWA. 73, 636-751 (1981).
Shull, K.E. "An experimental approach to corrosion control." Jour. AWWA.
72, 280-285 (1980).
Singley, J.E. and Lee, T. "Pipe loop system augments corrosion studies."
Jour. AWWA. 76, 77 (1984).
Singley, J.E. et al. Corrosion and calcium carbonate saturation index in
water distribution systems. Municipal Environmental Research
Laboratory, USEPA (1984).
Snoeyink, V.L. and Kuch, A. Principles of metallic corrosion in water
distribution systems in Internal Corrosion of Water Distribution
Systems. AWWARF, Denver, CO, pp. 1-32 (1985).
Sontheimer, H., Kolle, W. and Kuch, A. Uniform Corrosion and Scale Formation
in Internal Corrosion of Water Distribution Systems, AWWARF, Denver, CO,
pp. 62-88 (1985).
Sontheimer, H., Kolle, W. and Snoeyink, V.L. "The siderite model of the
formation of corrosion-resistant scales." Jour. AWWA. 73, 572-578
(1981).

182
Stern, M. and Geary, A.L. "Electrochemical polarization: Theoretical
analysis of the shape of polarization curves." J. Electrochem. Soc.,
104, 56 (1957).
Strickland, J.D.H. and Parsons, T.R. Fisheries Research Board of Canada,
Ottawa, 49 (1972).
Stumm, W. "Calcium carbonate deposition at iron surfaces." Jour. AWWA. 48,
300-310 (1956).
Stumm, W. "Investigation of the corrosive behavior of waters." ASCE-San.
Engr. Div., 86, 27 (1960).
Stumm, W. and Lee, G.F. "Oxygenation of ferrous iron." Indus. Eng. Chem..
53, 143 (1961).
Swayze, J. "Corrosion study at Carbondale, Illinois." Jour. AWMA. 75, 101-
102 (1983).
Trewick, G.P., Glicker, J., Chow, B. and Sprinker, M. "Pilot plant
simulation of corrosion in domestic pipe materials." Jour. AWWA.
77(10), 74-82 (1985).
Uhlig, H.H. Corrosion and Corrosion Control. John Wiley and Sons, New York
(1971).
Wagner, C. and Traud, Z. Electrochem., 44, 391 (1938).
Whitman, G.W., Russell, R.P. and Altier, V.J. "Effect of hydrogen ion
concentration on the submerged corrosion of steel." Industrial and
Engineering Chemistry. 16(7), 665-670 (1924).
Wood, J.W., Beecher, J.S. and Lawrence, F.S. "Some experience with sodium
silicate as a corrosion inhibitor in industrial cooling waters."
Corrosion, 13, 719t-724t (1957).

183
APPENDIX Al

VALIDATION OF WEIGHT-LOSS CORROSION-RATE MEASUREMENT TECHNIQUES

Table Al-1 presents the data from a series of surface-preparation tests


performed on the different coupon types. For each of the three types, a set
of 10 polished coupons was sequentially processed through three complete
surface cleaning and weighing procedures. The coupon set was handled as
described in Chapter 3, but was never exposed to actual water flow. Thus,
the observed weight loss and its variation between coupons is entirely due to
surface-preparation techniques.

TABLE Al-1
Weight Loss Variation due to Surface-Preparation Techniques

Complete Surface-Preparation Procedure


Performed in Sequence
Sequence 1 Sequence 2 Sequence 3
Avg. Std. Avg. Std. Avg. Std.
Wt. ^ Dev. T Wt. Dev. Wt. Dev.
Loss a Loss a Loss a
(mg) (mg) (mg) (mg) (mg) (mg)
Black-Iron Coupons 0.5 0.2 0.6 0.18 0.6 0.19
(set of 10)
Galvanized-Iron Coupons 1.1 0.28 1.3 0.32 1.0 0.27
(set of 10)
Copper Coupons 1.5 0.4 1.5 0.44 1.2 0.5
(set of 10)
* Average weight loss
f Standard deviation

On average, the metal losses for black-iron, galvanized-iron, and copper


coupons due to preparation procedures equalled 0.03, 0.07, and 0.12 mg/cm2 ,
respectively, demonstrating that careful application of the handling
procedures produces a minimal loss of metal. As important, the statistical
variation from coupon to coupon is both consistent and low for each cleaning
sequence.
185
The precision of this approach is substantially better than what can be
achieved using the techniques outlined in ASTM D 2688. A comparison of the
ASTM inhibited-acid cleaning procedure on similar coupon sets gave weight
losses more than an order of magnitude higher for black-iron and galvanized-
iron surfaces, and weight losses approximately equal on copper surfaces.
Similar results have been reported elsewhere (Montgomery, 1982). In the
experiments conducted for this report, exposure to the inhibited acid was
responsible for the bulk of the weight loss; even minimal exposure (5 min)
produced substantial metal loss.

In the mechanical cleaning procedure, the most serious metal loss occurs on
the copper coupons. Although it is only slightly higher than that observed
on black and galvanized iron, it is of greater significance because of the
substantially lower rate at which copper generally corrodes. The average
restoration loss of 1.5 mg represents the weight loss expected of an
equivalent pipe section corroding at a moderate rate over an exposure period
of from two to four days. For short-term weight-loss measurements, it is
critical to adjust for this restoration loss, and since the standard
deviation of the restoration-induced weight loss is substantial, a larger
number of coupons is necessary to provide a statistically meaningful result.
The number of coupons depends on the desired precision and the length of
exposure.

The reproducibility of an analytical measurement is an indication of its


precision; the less variation, the greater the precision. As discussed
earlier, standards of precision for weight-loss corrosion-rate measurements
are lacking, yet the weight-loss measure is often the standard by which other
forms of corrosion rate analyses (e.g., electrochemical or metal release) are
judged. One advantage of the multiple coupon sleeve assembly and associated
handling procedures is the relative ease with which measurement precision can
be determined.

Figure Al-1 displays a statistic of measurement precision for each of the


coupon types as a function of their exposure duration. The precision is
reported as the coefficient of dispersion, or normalized standard deviation
(a/mean). This number is usually shown as a percentage of the mean, and
186
50

40- 13 Black Iron

Galvanized Iron

30- Copper
Coefficient of
Dispersion (%)
00
-J 20-

10-

10 100 1000

Exposure Duration (days)

Figure Al - 1. Coefficient of Dispersion as a Function of Exposure Duration


represents the limits about the mean in which approximately 65 percent of a
normally distributed sample population will fall. These statistics derive
from exposure tests run in the Seattle distribution system. Each data point
represents a minimum sampling set of 10 coupons.

For all coupon types, there is a significant increase in measurement


precision as exposure duration increases and a concomitant decrease in
precision for shorter-term measurements, especially those measurements made
at less than two weeks. It seems intuitive that because of the lower overall
weight loss, short-term precision will be more closely related to measurement
variation induced by surface restoration procedures. Assuming that handling-
induced variance is independent of the variance produced by "natural"
corrosion processes, this theory can be tested by subtracting out the
variance (a*) due to surface restoration procedures (Table Al-2) from the
overall variance of the measure. The coefficients of dispersion for the
adjusted measure (Figure Al-2) demonstrate that after accounting for handling
procedures, the precision versus exposure duration is relatively constant for
both copper and galvanized-iron coupons, suggesting that the "natural"
variation in these corrosion processes is independent of time. Nonetheless,
this variation is substantial (6-10 percent of the mean) and requires that
any quantitative assessment of corrosion rates must apply the appropriate
statistical sampling strategy.

In contrast to the relatively uniform corrosion process of the copper and


galvanized-iron surfaces, black-iron exposure is subject to heavy pitting and
tuberculation. Visual observations of the corrosion pattern show that the
initial number, distribution, and depth of pits varies widely throughout the
coupon sample population. This initial variability likely accounts for the
lack of precision in the short-term measurements, irrespective of restoration
induced variation. As exposure duration increases and the involvement of the
black iron surface become more complete, weight-loss variations among these
coupons diminishes and the precision of the measurement approaches that
achieved for copper and galvanized-iron surfaces.

188
50

40- D Black Iron


Galvanized Iron
Copper
30-

Modified
Coefficient of
Dispersion (%)
20-
oo
sO

D
10-

0
10 100 1000

Exposure Duration (days)

Figure Al - 2. Modified Coefficient of Dispersion as a Function of


Exposure Duration
TABLE Al-2
Weight-Loss Variation due to Surface-Preparation Techniques

Complete Surface-Preparation Procedure


Performed in Sequence
Sequence 1 Sequence 2 Sequence 3
Average Weight Average Weight Average Weight
Weight Loss Weight Loss Weight Loss
Loss Variable Loss Variable Loss Variable
Black-Iron Coupons 0.5 0.04 0.6 0.032 0.6 0.036
(set of 10)
Galvanized-Iron 1.1 0.078 1.3 0.102 1.0 0.072
Coupons (set of 10)
Copper Coupons 1.5 0.16 1.5 0.20 1.2 0.25
(set of 10)

190
SUMMARY OF THE WEIGHT-LOSS TECHNIQUE

This report reviews new procedures for conducting corrosion-rate tests using
weight loss of metal coupons as the primary measure. A new assembly for
mounting the coupons and a new procedure for preparing and restoring the
metal surfaces are presented. The new procedure offers the advantages of
pipe flow hydrodynamics and multiple coupon mounting. The assembly is easier
to construct than the ASTM standard, and up to 10 coupons can be mounted in a
single sleeve, which makes it simpler to develop a statistically accurate
picture of the corrosion process.

Mechanical preparation and/or restoration of coupon surfaces as described in


this report offer a significant reduction in effort and time compared to the
combination of mechanical and chemical preparatory steps outlined in ASTM
standards. Weight loss associated with the mechanical preparation techniques
described here is both minimal and reproducible.

The reliability of the coupon preparation/restoration techniques and the


subsequent reproducibility of corrosion-rate measurements expand the
application range of the weight-loss approach to even short-term analyses.
In some cases, statistically meaningful corrosion rate data can be achieved
with as little as a two-week exposure period if a full 10-coupon sleeve is
used.

191
APPENDIX A2

ELECTROCHEMICAL CORROSION -RATE MEASUREMENTS: PRINCIPLES AND VALIDATION

Linear polarization has been widely used to measure corrosion rates since
1939. The basic relationship derives from the modified Butler-Volmer
equation and assumes that electrochemical kinetics are solely a function of
the difference between the imposed potential and the potential of the freely
corroding metal surface (Bockris and Reddy, 1973). The modified Butler-
Volmer equation is:

I - {E - Ecor)/Ba . -(E-Ecor)/BC
xmeas
Where:
Imeas = the measured polarization current that flows between the
metal and the counter electrode (equal to the difference
between the partial anodic and cathodic currents, i 0 m -
V,z)
Ecor = the open circuit corrosion potential (freely corroding
potential)
E = the potential imposed upon the metal
I cor = the corrosion current that exists under freely corroding
conditions; it is also the value needed to determine the
corrosion rate.

Since it cannot be directly measured (i 0 m = i r z at Ecor ), an estimate is


developed from the polarization data. Ba and BC are the anodic and cathodic
Tafel slopes, respectively. They represent the change in potential (E-Ecor )
per decade change in polarization current (I meas ) for oxidation and reduction
reactions on idealized electrode surfaces. In practical terms, if passivated
or coated surfaces are measured, they may be considered fitting parameters
for the modified Butler-Volmer equation.

Application of this equation to a corroding surface is valid given the


following assumptions: (1) no ohmic voltage drops and no concentration
polarization exist; and (2) the rest potential of the corroding system is

193
sufficiently displaced from the reversible potentials of the respective
anodic and cathodic reactions (usually differences of 200 mV or greater are
sufficient).

Classical polarization resistance techniques derive from the observation that


Equation A2-1 becomes nearly linear for small values of [E-Ecor]. The
derivative of current with respect to potential taken at [E-Ecor=0] results
in the expression:

[dImeas/dE]=2.3I cor [l/Ba+l/Bc ]=l/Rp (A2-2)

This can be rewritten in the familiar form:

where Rp , the polarization resistance, has units of ohms and is equal to the
slope of the potential versus current plot at [E-Ecor=0]. The factor K
becomes an explicit function of the Tafel slopes.

Equation A2-3 offers a simple and direct means of solving for I cor if an
accurate estimate of Tafel slopes is available. Commercial corrosion-rate
meters make use of this equation by employing an internal "hard wired" value
for K and measuring the polarization resistance at a single offset from Ecor ,
usually less than 20 mv. While these meters offer simplicity and speed,
absolute accuracy suffers because the electrochemical parameters cannot be
adjusted to fit different cell conditions, or the changing conditions of a
The given cell over time (Bandy and Jones, 1976; Hausler, 1977).

SURFACE- POLARIZATION PROCEDURES

Ecor is the freely corroding electrochemical potential that a surface assumes


(with respect to the reference electrode) in the absence of an impressed
current. Because the test electrode surfaces used in this study are
relatively large, significant spatial variations in Ecor may exist from point
to point on a coupon. As described by Uhlig (1971), the freely corroding
surface takes the form of multiple "local-action cells" where the respective
194
oxidation and reduction half-cell reactions occur. As will be demonstrated,
large variations in potential may exist between different local-action cells.
As corrosion reactions proceed, positive and negative electrode areas may
interchange and shift locations. The more negative cells form the anodic
sites where the metal is oxidized. These are often covered by a porous layer
of corrosion scale. The more positive cathodic cells form the sites where
the corresponding reduction of oxygen occurs. Such sites are usually clear
of corrosion products.

Of the three materials studied, the variation of surface potential was the
most pronounced on black-iron coupons. Variation was enhanced because of the
tendency of these surfaces to undergo pitting-type corrosion under the water
quality conditions tested. Conversely, with few exceptions, copper surface
corrosion is uniform under the distribution system conditions studied. The
zinc coating on galvanized pipe is also more uniform in its initial corrosion
patterns than black iron; however, once the galvanized layer has been
penetrated, it is subject to extreme pitting.

The distribution of anodic and cathodic sites on metal surfaces and resulting
variations in corrosion potential has been explored by several authors
(Dexter et al., 1985; Freneir and Growcock, 1984). Experience with this
technique has shown that the more nonuniform the corrosion pattern, the
larger the surface variations in Ecor . With the flexible reference electrode
and the rotating cap of the polarization cell, it is possible to scan the
surface of the test electrode and map the corrosion potential variation.
Figures A2-1, A2-2, and A2-3 present, respectively, a surface scan over a 1-
cm-wide circumferential band of an iron, galvanized-iron, and copper pipe
coupon. Each coupon had been exposed for approximately 96 hr to a 1 ft/sec
velocity of Seattle tap water (temperature = 15'C, conductivity = 55
umhos/cm, dissolved oxygen = 10 mg/L, pH = 7.8).

Figure A2-1 shows the dramatic variation in Ecor associated with the
nonuniform iron corrosion. Although this coupon had been exposed for only 96
hr, distinct pitting sites had been initiated, accompanied by an overlaying
iron oxide scale. The deep valleys of the scan correspond to these areas,
while the flatter higher-potential areas represent the bulk of the coupon
195
-1601
so Surfoca
Potential -200
faV)
-240 -i

Depth
(c)

40 160 200 240 380


Rotation (degrees)
Figure A2-1. Surface Potential Scan Across a One Centimeter Band of a Black
Electrode (96 hr. exposure)
\D
Surface
Potential
(nV)

Depth
(cm)

120 160 200 240 360


Rotation (degrees)
Figure A2-2. Surface Potential Scan Across a One Centimeter Band of a Copper
Electrode (96 hr. exposure)
-600

-640

-680-

-720-

-760
VO Surface
Of
Potential -800

-8401
-880

-9201
-960 7
....
7
/
7
/...../
/ DcPth
(rm)

-1000 |ii|ii|ii|rr[ T i | i i i i i | i r <0


40 80 120 160 200 240 280 320 360
Rotation (degrees)
Figure A2-3. Surface Potential Scan Across a One Centimeter Band of a Galvanized
Iron Electrode (96 hr. exposure)
surface that remains free of scale covering. The pits at this exposure
duration are generally about 2 mm in diameter with a penetration depth of
less than 0.1 mm. The scan, however, indicates that the area of influence
around the pit is substantially larger than the pit itself.

The uneven distribution of the scale distorts the measurements of


polarization resistance (dE/dl) used to characterize the polarization
behavior of the coupon. When the distribution of corrosion scale is uneven,
it is often necessary to make multiple readings at different surface
locations to determine a representative value of dE/dl. Comparisons of
electrochemical and weight-loss measurements indicate that extreme
polarization resistance values on the pitting surfaces, either high or low,
should be discarded. Measurements taken on the more cathodic areas of the
coupon tend to "cluster" around an average value that more accurately
reflects the overall corrosion rate as measured by coupon weight loss.

This averaging technique works satisfactorily on iron coupons where the


corrosion scale is relatively thin (less than 1 mm). However, experience has
shown that this is unsatisfactory on aged black-iron coupons with thick
scales or large patchy scales, as opposed to a distribution of smaller
discrete pits.

The potential scans on copper and galvanized iron (Figures A2-2 and A2-3)
show a more stable surface Ecor than iron. The uniform nature of their
corrosion process and scale deposition explains this difference.
Polarization resistance measurements on these surfaces show only slight
spatial variation, and since both copper and zinc scales are generally
thinner than iron scales, polarization measurements can be made anywhere on
the surface regardless of coupon exposure time or scale growth.

THE LEROY MODEL

Electrochemical parameters are calculated from the polarization data by


fitting each point to the modified Butler-Volmer equation and selecting Tafel
slopes that give the least overall error between observed and predicted
polarization currents. In essence, a brute force approach is used that tests
199
every feasible combination of anodic and cathodic Tafel slope and ultimately
selects the pair that gives the best fit. As such, it lends itself readily
to a computer-based solution. After establishing the appropriate search
boundaries and iteration increments, a corrosion current is calculated using
the Leroy Model and an initial selection of Tafel slopes (Leroy, 1975).
Leroy's model is a limited set of polarization data and a given set of
assumed Tafel slopes. It takes the form:

2. =1 [I.(lQ.(EJ-Ecor)/Ba . 10 -(Ej-Ecor)/Bc)
Icor = s~ ri 0 (Ej-Ecor)/Ba - iQ-Ctt-EcorJ/Bc^] (A2 " 4)
J *

where j=l...n represent the n data points. Using the entire data set (n
values of E and I), I cor is completed from Eq A2-4. This value of I cor is
then used with Eq 1 to compute the value of Imeas for each value of E. If
any computed value of Imeas differs from the experimental value of more than
some arbitrary cutoff (usually 5 percent to 10 percent of the experimental
value), the set of assumed Tafel slopes is rejected. If all n computed
values of Imeas fall within the criteria, the values of the Tafel slopes are
stored along with the sum of the squares error
^experimental) '* Tafel slopes are then incremented and the calculation
proceeds until all possible combinations established within the original
limits have been tested. Selection of the "best" Tafel slopes uses the
minimum sum of the square of the errors as its goodness-of-fit criteria.

The procedures described appear cumbersome and time consuming, even for a
computer-based solution. However, the task is simplified by the
insensitivity of Eq A2-1 to minor changes in the Tafel slope values, and the
well -documented range of feasible Tafel slope values for the metal of
interest. Because of insensitivity, the incremental adjustments to the Tafel
slopes need not be less than 10 mV. Choosing a smaller value may find a
marginally better fit, but such efforts are not justified considering the
precision of the actual polarization measurements. A large body of
experimentally determined Tafel slopes for each metal has been catalogued by
the National Association of Corrosion Engineers (Grauer et al., 1982). These
data will generally bracket a feasible solution. An iteration boundary of

200
plus-or-minus 100 mV around the anticipated value is usually sufficient to
locate a satisfactory fit.

Figures A2-4 and A2-5 are Evans diagrams constructed from polarization data
collected on black-iron coupons exposed to the Seattle distribution system
for 24 hr and 7 days, respectively. The lines of best fit (polarization
curves) resulting from the parameter estimation technique described above are
overlaid on the six measured polarization points. In this case, the maximum
single error between measured and predicted values is less than 5 percent.
It is generally possible to find a fit that falls within an error range of 5-
10 percent. The dashed lines intersecting at Ecor are extrapolations of the
predicted Tafel slopes from the linear region of the polarization curve. The
current density at the point of intersection represents the current density
of the freely corroding surface.

Figures A2-6 and A2-7 present Evans diagrams of measured and predicted
polarization values derived from copper and galvanized-iron surfaces exposed
for seven days to the Seattle distribution system. In both cases, the
analytical approach was the same as for the black-iron examples. Table A2-1
summarizes the parameters derived from the four polarization examples and
compares electrochemically measured corrosion rates to weight-loss estimates
of corrosion rates made on five coupon sets exposed to identical conditions
for similar durations. Corrosion current densities are converted directly to
a penetration rate by applying Faraday's law:

Corrosion Rate (mpy) = I cor [3.21(eq.wt.)/s.g.] (A2-5)

where s.g. is the specific gravity of the corroding metal and the equivalent
weight (eg. wt.) represents the gram weight of metal oxidized by the transfer
of one equivalent of electrons.

ACCURACY OF CORROSION MEASUREMENTS

Very little literature is available attesting to the accuracy of any


analytical technique for corrosion-rate measurements in water distribution
systems. Discussion of linear polarization measurements by the National
201
200

150

100

50
/ solid lines predicted values
/ - Masured values
Offset 0 F I,
(V)
10.2 uA/ca
NJ
o -50
K)

-100 Seattle Tap Voter


pH 7.8
Cond. 65u*ho8/c
-150 Bc - 250 V/decade Temp 15 deqrees-C
DO 10 M/l
lov Vel. Ift/eed
-200
-1 -0.5 0.5 1 1.5 2.5
Log Current Density (uA/cm2)

Figure A2-4. Evans Diagram of Measured and Predicted Polarization Currents on a


Black Iron Electrode (24 hr. exposure)
200

150
BQ - 220 iV/decode
100

SO
solid lines predicted values
* Measured values
Offset OF -I
(V)
K)
O
U)
-50

-100
Saottlo Top Water
pH- 7.8
Cond. * 65uMhos/o
-150 B c - 180 V/decade Te*p 15 deqrees-C
00 10 g/l
:low Vel. - 1 ft/ecc
-200
-1 -0.5 0 0.5 1 1.5 2.5
Log Currant Density CuA/ca2)

Figure A2-5. Evans Diagram of Measured and Predicted Polarization Currents on a


Black Iron Electrode (7 day exposure)
200

150
B0 - 80 V/dacoda
100

50
solid linos predicted values
* Measured values
Offset
(V) ICOT l.B uA/ai
\
N
-50

-100 Seattle Tap later


pH-7.8
Cond. 65mhos/CH
-150 Bc IBOiV/de Tenp 15 degrees-C
DO 10 mq/l
Flow Vel. - Ift/sed
-200
-1 -0.5 0 0.5 1 1.5 2.5
Log Current Density (uA/cn2)

Figure A2-6. Evans Diagram of Measured and Predicted Polarization Currents on a


Copper Electrode (7 day exposure)
200

150
8Q * 150 V/dacoda
100

50
olid Una* predicted values
* Matured values
Offset n r
<V) t

-50

-100
Seattle Tap Water
pH-7.8
Cond. 85uho/c
-150 Bc 120 V/decode Tep "15 degrees-C
uO - 10 TO/1
:lot Vol. - Ift/eed
-200
-1 -0.5 0 0.5 1 1.5 2.5
Log Currant Density (uA/c2)

Figure A2-7. Evans Diagram of Measured and Predicted Polarization Currents on a


Copper Electrode (7 day exposure)
TABLE A2-1
Comparison of Electrochemical Corrosion Rates
With Weight-Loss Measurements

Galva
Black Black nized
Iron Iron Copper Iron
Exposure Period 24 hr 7 days 7 days 7 days
Exposure Medium Seattle Seattle Seattle Seattle
tap tap tap tap
Fitted Anodic Tafel Slope (mV) 250 220 80 150
Fitted Cathodic Tafel Slope (mV) 250 180 180 120
Corrosion Current Density 10.2 7.2 1.8 1.37
(mA/cmz )
Corrosion Rate - Electrochemical 114 81 41 18
(lM/yr.)
Weight-Loss Measurement 5.5 2.8 1.02 0.92
(5 coupon set) (mg/day)
Corrosion Rate - Weight Loss 134 68 33 24
(/im/yr)
Difference in Weight Loss vs. -15 +19 +24 -25
Electrochemical Corrosion Rate (%)

206
Association of Corrosion Engineers speaks of predicted values being within
one-to-two fold of actual corrosion rates (Grauer et al., 1982). Even ASTM
standards are reluctant to state an acceptable level of precision because of
the large variability encountered between systems. By contrast, the examples
of Table A2-1 come within 25 percent of weight-loss measurements and yield
additional information about the electrochemical processes. The hardware and
analytical techniques described here offer improved accuracy for corrosion
measurements on plumbing materials exposed to tap water. The accuracy
approaches that necessary to distinguish between the more subtle effects of
various chemical corrosion inhibitors or to access the effects of minor water
quality changes.

Good correlation of the electrochemical techniques with weight-loss


measurements can be readily achieved on copper's uniformly corroding
surfaces. However, surfaces that exhibit thick or extremely patchy scale,
such as well-aged black-iron or galvanized-iron coupons, severely complicate
the measurement of surface potential and polarization resistance. The
spatial averaging technique described earlier is only a partial resolution
that is appropriate at the incipient stage of pit and scale formation.
Exposure periods of as little as 30 days in some aggressive waters may
produce extreme pitting with abundantly scaled surfaces that cannot be
evaluated in this fashion.

SURFACE POTENTIAL AND PITTING

As demonstrated, the polarization cell is useful in mapping the surface


potential of the corroding metal coupons. For corroding iron, the anode and
cathode reactions can be traced to specific areas of the metal surface. The
surface scans define these areas by their relative potentials. Often, the
extreme anodic sites represent well-defined or incipient pits. Pitting
corrosion is generally a more serious problem for black-iron pipe than is
uniform corrosion, and it is often failure due to pitting that determines the
useful service life of the pipe. Hence, penetration measurements based on an
assumption of uniform corrosion do not yield useful information about the
critical processes that may be occurring. Furthermore, pit formation and

207
scale growth appear to impair the accuracy of polarization measurements and
make estimation of even a uniform penetration rate imprecise.

PIPE SURFACE ELECTROCHEMISTRY

In order to accurately assess corrosion rates, a reasonably accurate estimate


of the basic electrochemical kinetic parameters (Tafel slopes) is necessary.
The examples in Table B.I demonstrate how these parameters differ from metal
to metal, as well as between coupons of the same metal under slightly
different conditions. Clearly, the assumption that one set of parameters, or
a single "hard-wired" constant, can be applied to a variety of electrodes,
may lead to serious errors in the corrosion-rate estimate.

It is not surprising that substantial differences in electrochemical kinetics


exist between different metals, or the same metal for different exposure
periods. That a substantial difference can develop on different parts of the
same metal surface exposed to a constant water quality is less intuitive, and
it is important to remember this when interpreting distribution system
corrosion-rate measurements. The variation demonstrated for black-iron
surfaces in Table A2-1 is typical of the difference observed between fresh
and relatively aged surfaces.

208
NOMENCLATURE

a = Standard Deviation
ASTM = American Society for Testing and Materials
Ba = Anodic Tafel Slope (mV/decade of polarization current)
Bc = Cathodic Tafel Slope (mV/decade of polarization current)
C = Degrees Centigrade
COg = Carbonate
Cj = Total carbonate-concentration
Cu = Copper
DO = Dissolved oxygen, mg/L
E = Measured Surface Potential relative to the Ag/AgCl Reference
Electrode (mV)
Ecor = Freely Corroding Surface Potential Relative to the Ag/AgCl
Reference Electrode (mV)
Fe = iron
fps = Feet per second
H2C03 = Carbonic Acid
ty
I cor = Corrosion Current Density (/jAmps/cnr)
Imeas = Measured Current Density Impressed upon the Test Electrode
Surface (/jAmps/cnr)
I 0 = Equilibrium exchange current density, /tA/cnr
nU * IIIm = Partial
ara 4- Anodic Current Generated from the Oxidation of Metal
m to nr
i r z = Partial Cathodic Current Generated from by the Reduction of
' Substance z
ISWS = Illinois State Water Survey
K = Constant, a Function of Tafel Slopes [BaBc/2.3(Ba+Bc )]
Ks = Solubility constant for siderite
L = Liter
209
mA = Milliamperes
mg = Milligrams
mpy = Mils per year
mv/dec. = Millivolts per decade
mV SHE = Electrical Potential Relative to a Standard Hydrogen Electrode
P = phosphorus
P04 = phosphate
Rp = Polarization Resistance, (dV/dl at E-Ecor=0)
rpm = Revolutions per Minute
s.g. = Specific Gravity of the Corroding Metal
S.I. = Langlier Saturation Index
uS = MicroSiemens
Zn = zinc

210
American Water Works Association

TM

6666 WQuincy Avenue, Denver, CO 80235


(303) 794-7711

1 P-7.5C-90561 -2/90-CM ISBN 0-89867-506-5

You might also like