United States
        Environmental Protection    Office of Water   EPA 815-B-97-004
        Agency        ,    4607       December 1997



&EPA  The Effect of Temperature


        on Corrosion Control

-------

-------
              "The Effect of Temperature on Corrosion Control"
                                 EPA 815-B-97-004
                                   December 1997
This Publication Contains the Following Documents:

Colling, J.H., Croll, B.T., Whincup, P.A.E., and Harward, C. June 1992. Plumbosolvency
Effects and Control in Hard Waters. JIWEM, 6(3):259-268.

Dodrill, D.M., and Edwards, M.  July 1995.  Corrosion Control on the Basis of Utility
Experience. JournalAWWA, 74-85.

Edwards, M., Schock, M.R., and Meyer, T.E. March 1996. Alkalinity, pH, and Copper
Corrosion By-Products Release. Journal A WWA, 81-94.

Rezania, L.W. and Anderl, W.H.  1996. Copper Corrosion and Iron Removal Plants.
Conference Paper. Section of Drinking Water Protection, Minnesota Department of Health.

U.S. Environmental Protection Agency. An Evaluation of the Secondardy Effects of Enhanced
Coagulation, With Emphasis on Corrosion Control.  Conference Paper prepared by D.A. Lytle,
M.R. Schock, and RJ. Miltner, Treatment and Technology Evaluation Branch, Water Supply and
Water Resources Division, National Risk Manageent Research Laboratory.

U.S. Environmental Protection Agency. December 19, 1996. Memo from Michael R. Schock,
Treatment and Technology Branch, Water Supply and Water Resources Division, National Risk
Management Research Laboratory, to Jeffrey B. Kempic of the  Office of Ground Water and
Drinking Water, entitled: Seasonal Monitoring Revision. December 19, 1996.  (Note:
References 5, 6, and 7 cited in the memo are not provided for public review and comment. The'
Agency is not factoring the data contained hi these studies into its decision making.)

Wagner,  I. June  18-22, 1988. Effects of Inhibitors on Corrosion Rate and Metal Uptake.
Proceedings of American Water Works Conference.  Memo from Michael R. Schock to Judith
M. Lebowich, February 13, 1997, regarding Requested References.

-------
            FE?-06-1997   12:37
                      USEPA
                                                                           5135597172    P.02
 •-•nt of new Hltcr
 •/. IVX7.86. (II.

 v from sludno.
 !. /. E. Ilallj.
        Plumbosolvency .Effects  and  Control in Hard Waters

By J. H. COLLING. BSc'.  B. T. CROLL. BSc. PhD (Fellow)**. P. A. E. WHINCUP.
                      PhD",  and C. HARWARD. BSc (Member)**
                                        ABSTRACT

                      A laboratory lead-pipe rig has been used to support
                      Anglian  Water's successful orthophosphate dosing
                      programme to reduce plumhosolvency. The hard
                      waters in the region generally fall into low or high
                      plumfaosolvency categories according to the types of
                      crystalline deposit formed.  To  improve the cost-
                      effectiveness of plumbosolvency control, the effects
                      of temperature, phosphate  doses,  blending and
                      alternation of these waters were investigated.
                        Initial  phosphate concentrations must be above
                      0.6 mg P/I  (as  phosphorus)  to establish plumbo-
                      solvency control.  Subsequently, phosphate  doses
                      may be reduced, provided that dosing is continuous
                      and sufficient phosphate  reaches the extremities of
                      the distribution system. When high and low piumbo-
                      solvency  waters are  blended before distribution.
                      both (or the mixture) must  be phosphate dosed.
                      However, where waters  alternate in distribution.
                      laboratory studies have  shown  that  low  plumbo-
                      solvency  deposits are more stable, resulting in low
                      lead  concentrations.  The high  plumbsolvency  of
                      some hard waters may be  due to the presence of low
                      concentrations of humic substances.
                      AVy words: Hard water: loath phosphate: plumhosolvency:
                      treatment: vvurcr.
                                     INTRODUCTION

                      Up to the mid-1970s hard waters, such as exist in.the
                      Anglian Water  region, were not  normally  con-
                      sidered to be plumbosolvent. the problem normally
                      being seen as one  of soft waters.  The increased
                      water supply  monitoring  in the Anglian region
                      following the reorganization of the water industry in
                      1974 showed that some hard  waters were  plumbo-
                      solvent. although not to the >ame  degree as  soft
                      waters.  This information  wu>  continued  by  the
                      survey organized by the Department of the Environ-
                      ment'(DoE) from'1979 to IVSI1" which identified
                      priority  areas  for action to  reduce lead  concen-
                      trations at the  tap.

                      This paper is an updated version 01 .1  paper presented 1" the
                      Institution's Scientific Suction Symp>v-mm tin l.twl in Vt'uier held in
                      London on 12 April lixy.
                      •Research Fellow and Senior Lecturer iHumberside I'nlytcchnici.
                      "Proecv.  Science  Manager  and r.-tncipiil  IVocc^  Scientist.
                      Anglian Water Services. Ltd.
0. 6. June.
                     J.IWKM. l«M2. 6. June.
                                             .  3;
                                                 Concurrent with  the  DoE survey, experimental
                                               work at the Water Research Centre (WRc) identi-
                                               fied that the addition of orthophosphate to waters at
                                               a concentration of approximately 1.0 mg P/l  (as
                                               phosphorus) was able to  reduce lead  solvency in
                                               hard waters1-'.  This treatment  was  successfully
                                               installed in two triai areas in Anglian Water (Boston
                                               and  Sleaford)1-" and was  subsequently  recom-
                                               mended for plumbosolvency control in  hard waters
                                               by the \VRc'a' and DoE'-"."
                                                 The 1979-S1 survey identified  those areas where
                                               the incidence of lead solvency was highest, based on
                                               the proportion of random samples givina elevated
                                               lead concentrations. Therefore, if an area had very
                                               few lead pipes but  a  plumbosolvent  water, it was
                                               possible that  it  would  not be  designated  high
                                               priority. The  former  Anglian  Water Authority
                                               decided that the small number of customers in such
                                               areas should be protected against high lead concen-
                                               trations, and therefore adopted a policy of treating
                                               all high plumbosolvency waters,  unless it .--mid be
                                               demonstrated  -h.it lead pipes  were abseii: in the
                                               distribution s\stem.
                                                 In order  to classify its waters,  a  contract was
                                               placed with Humberside College of High Education
                                               (now Humberside Polytechnic) to develop a test to
                                               classify   waters  us  high-plumbsolvency  or  low-
                                               plumbosolvency and subsequently to test all Anglian
                                               Water treated waters1'".  On the basis of this work.
                                               phosphate dosing to control plumbosolvency  has
                                               been successfully installed at  S5 water-treatment
                                               works in Anglian Water.
                                                 Following the survey performed  by  Humberside
                                               College, further information on the topics  listed
                                               below was required in order to ensure cost-effective
                                               control of plumbosolvency:
                                                    Effects nf temperature on Icatj solubility:
                                                    The  minimization «t' phosphate do>e to eMuhlish
                                                    and maintain plumliOsoHency control:
                                                    The effects of breaks it! phosphate dosing:
                                                    The mixing of waters in distribution: and
                                                    The effect of ustnu exhumed old lead pipes in the
                                                    lest apparatus.
                                                During  the  experimental  programme  on  the
                                              above topics, it was discovered  that low  concen-
                                              trations  of organic compounds, probably naturally
                                              occurring, can influence plumbosolvency. This work
                                              is also described later in this paper.

                                                                                        •  259

-------
   FEB-06-1997
    12:38           USEPft

     WHINCUP AND HARWARD ON

LEAD CONCENTRATION
                                        5135697172   P.03'
                                                         HIGH
                                                 PLUMBOSOLVENCY
                                                        LOW
                                                PLUMBOSOLVENCY
                                              PHOSPHATE DOSED (1.0mg P/i)
                              5          10         15
                                      TEMPERATURE (°C)
                         Fig. 1.  Effects of temperature on lead solvency
 PLUMBOSOLVENCY TESTING EXPERIMENTS

 Full details of the standard test have been pub-
 lished' '. The test water is continuously pumped at
 30 ml/h through ISO mm sections of new 12 mm lead
 pipe at 25°C for up to 25 days. The measured lead
 concentrations in the exit water from each pipe
 normally begin to be stable after 7-10 days. The
 stable lead concentration is quoted as the test result.
 Most waters give  a 'high piumbosolvency' result
 (100 ± 20 ug/l), or a Mow plumbosolvency' result {30
 ± 10 ug/l).
  These differences are thought to be due to the
 formation of either the  more soluble  basic lead
 carbonate or the less soluble normal lead carbonate.
 as the corrosion product on the lead surface'-'. This
 is despite the prediction that normal lead carbonate
 should be the thermodynamicaHy stable product
 over the  pH  and  alkalinity  range involved1--71.
 Scanning electron microscopy (SEM) reveals char-
 acteristic differences in the crystal deposits formed
 in pipes during the test. Thin hexagonal plates are
 formed  from high  piumbosolvency waters,  and
260
 relatively smooth solid surfaces from low piumbo-
 solvency waters.  Typical micrographs  have  been
 published earlier"".
   It should be noted that distinct crystals cannot be
 identified in  old  pipes.  Pipes that  have  been in
 service are generally coated with relatively thick.
 stratified and contaminated deposits. However, the
 presence of  basic lead  carbonate in some  such
 deposits has been confirmed by x-ray diffraction and
 infra-red absorption spectroscopy1-'.
   PHOSPHATE-DOSED RESULTS AND SEM

Waters dosed with other phosphates  (1.0 me  P/l)
give  low lead levels, frequently less than I0"ug/l.
With continued running, lead values as low as 3 ug/I
can be obtained.  The  lead phosphate compound
formed is clearly of extremely low solubility, and its
exact  composition  is  not known.  It has a very
characteristic  appearance  by  SEM  (rounded
granules and  surface  features, as evident  in  the

                           J.IWEM. 1992. 6, June.
                                                       published photograp
                                                       phorus and some calc
                                                       dispersive x-ray spec

                                                           EFFECT OF TI
                                                                       S(

                                                       The monitoring rei
                                                       solvency Boston dist
                                                       lead concentrations
                                                       both untreated and j
                                                       effect was thought t
                                                       water temperature.
                                                       accurately quantify i
                                                       piumbosolvency wa'
                                                       vency  water dosed
                                                       phate. were examine
                                                       and 25°C (in randor
                                                       are plotted against
                                                       the  different  lead
                                                       waters follow a simi
                                                       be expected for simp
                                                       (Pb) versus tTemp
                                                          Deposits on a  :
                                                       examined by  SEM.
                                                       crystals are generall
                                                                                                   LEAD

                                                                                                    110-

                                                                                                    100-

                                                                                                     90-

                                                                                                     80-

                                                                                                     70-

                                                                                                     60-

                                                                                                     50-

                                                                                                     40-

                                                                                                     30-

                                                                                                     20-

                                                                                                     10-

                                                                                                       o-



                                                                                                    R9-

                                                                                         J.IWEM. ISN2. 6. June.

-------
USEPA
           FEB-06-1997   12:38
                 published photograph(6), and the presence of phos-
                 phorus and some calcium can be detected by energy-
                 dispersive x-ray spectroscopy (EDS).

                     EFFECT OF TEMPERATURE ON LEAD
                                  SOLUBILITY

                 The monitoring results from the high piumbo-
                 solvency Boston distribution system showed higher
                 lead concentrations in summer than  in  winter in
                 both untreated and phosphate-dosed waters'6'. This
                 effect was thought to be due to the differences in
                 water temperature.  In order to confirm and more
                 accurately quantify the effect, typical high and low
                 plumbosolvency waters, and the high plumbosol-
                 veney water dosed to 1.0  mg P/l  with orthophos-
                 phate. were examined at 0°C. 5°C. 10°C. 15°C, 20°C
                 and 25°C (in random order); the mean lead results
                 are plotted against temperature in Fig.  1. Despite
                 the  different  lead  concentrations, the undosed
                 waters follow a similar relationship, and, as would
                 be expected for simple solubility effects, plots of log
                 (Pb) versus. l/Temp (°K) are effectively linear.
                   Deposits on  a  selection of pipes  were also
                 examined by SEM.  At the lower temperatures the
                 crystals are generally similar to those described for
                             LEAD CONCENTRATION
                                                                                                 5135697172    P.04
 un bo-
 been

 lot be
 en in
 thick.
 r,  the
 such
 nand
M
ound
id its
very
nded
  the
June.
                    25°C(6). although smaller  in size. These examin-
                    ations therefore support the water lead concen-
                    tration results in indicating  that the: temperature
                    effects in  undosed waters are largely due to crystal
                    solubility.  Since no change  in  mechanism  is
                    involved,  it is clear that  the  accelerated results
                    obtained  at 25°C  in the  standard  test are  also
                    representative of effects at lower temperatures.
                      The phosphate-dosed high plumbosolvency water
                    is different, showing little temperature dependence,
                    and apparently a small decrease in lead level with
                    increasing temperature.  However, all these lead
                    values are low, and the results may simply reflect a
                    more rapid approach to ultimate coverage of phos-
                    phated deposit  at  the   higher   temperatures.
                    SEM/EDS examinations show the expected round
                    nodules containing phosphorus and calcium.

                       INITIAL PHOSPHATE CONCENTRATION

                    The WRc investigations suggested  that for high
                    alkalinity   waters  phosphate  concentrations  of
                    0.7-1.0 mg P/l(4) would be effective in establishing
                    plumbosolvency  control  within  a  distribution
                    system. In order to optimize the phosphate concen-
                    tration required  for a  particular high  plumbo-
                                                                                0.2 mg P/l
                                                                           PHOSPHATE DOSED
                                  20
                                      T
      0          5         10         15
                            TIME (DAYS)
Fig.  2.  Effects  of  initial  orthophosphate  concentration  on   high
                    plumbosolvency water at 25°C


-------
       FEB-06-1997  12:39           USEPfl

     COLLING. CROLL. WH/NCUP AND HARWARD ON
                    LEAD CONCENTRATION
                            (pg Pb/i)
                100
                                        PHOSPHATE DOSED
                                               (mg P/I)
                                          10         15
                                       TEMPERATURE (=
  solvency water, experiments were  performed in
  concentrations of 0.2-1.6 mg P/I. The results are
  shown m fig. 2. They demonstrate that the phos-
  phate dose affects  not only the finai lead concen-
  trauon. but also the rate at which the phosphate is
  effective. It IS clear that 0.2 mg/! has little effect. For
  the water tested. 0.4 mg/1 has some effect, but at a
 slow rate, and this dose is less effective with some
 other waters. It is  only at doses of 0.6 ma/1 and
 above that the major effect of phosphate dosing is
 apparent. However, the faster rate of establishment
 of control at  1.0 mg P/I and above is sienificam.
 particularly when distribution system conditions are
 considered.
   Theoretically,  as  noted  by  staff of the WRC'7)
 phosphate dose  requirements should be tower at
 lower calcium concentrations. This was confirmed
 by similar experiments using  water  containing 50
 mg/l of calcium compared with about 150 me/I for
 the water used  in  Fig. 2.  The  effect, however
appeared to be relatively small, the final lead value
262
   The effect of a range of initial phosphate concen-
 rranons was also examined at a series of different
 temperatures (0-25°C). The results (Fie. 3) indicate
 that the 0.2 and 0.4 mg p/| waters beh^ ^^
 undo*d waters, wh.lst the concentrations of 0.6me
 P/I and above behave as phosphate-dosed waters "
   The above results indicate that an initial phos-

 Sr?se   (!'6 mrP/I or above  is ne™'»
 «tabhsh control of  phimbosolvency in these waters
 The higher the phosphate concentration above 0 6
 mg P/l. the more rapidly is control established.


      BREAKS IN PHOSPHATE DOSING

The effect of breaks in phosphate dosing is import-
ant tor two  reasons, (i) to assess  the~ impact of
equipment failure at the treatment works, and (ii) to
investigate intermittent dosing as a means of achiev-

                            J.IWEM. IW2. 6. Juno.
                                             5135697172    P.05
                                                         ing cost-effec
                                                         dence expert
                                                         trations in ui
                                                         winter than si
                                                         a given lead <
                                                         policy of sum
                                                           In  order u
                                                         phosphate do
                                                         were establisi
                                                         weeks with d.
                                                         low  plumboi
                                                         stopped and i
                                                         period. Mean
                                                         ing the chans
                                                           It is clear tl
                                                         differently.  1
                                                         there  was a
                                                         concentrator
                                                         undosed  lea-
                                                         cessation  of •
                                                         water howevi
                                                         and lead vak
                                                         occurred witf
                                                         a slow deer
                                                         values.
                                                           The pipes l
                                                         lead values w
                                                         from the reU
                                                         of basic  and
        2-\A



  100  -

   90  -

   80  -

   70  -

   60  -  [j

   50  -

   40  -

   30  -

   20  -

   10  -

    0  -*=

         D>

      Fig.-

J.JWEM. 1W2.

-------
     FEJ3-06-1997   12:40
                             USEPfl                                              5135697172  '  P. 06

                           PLUMBOSOLVENCY EFFECTS AND CONTROL IN HARD WATERS
   ing cost-effective control. The temperature depen-
   dence experiments indicated that  lead concen-
   trations in untreated waters were much lower in
   winter than summer, and therefore control to below
   a given lead concentration might be  achieved by a
   policy of summer-only phosphate dosing.
    In order to investigate  the effect of breaks in
   phosphate dosing, appreciable phosphated deposits
   were established in pipes run continuously for 33
   weeks with doses of 1.0 mg P/l on typical high and
   low  plumbosolvency  waters.  Dosing  was then
  stopped and undosed water run for a further lengthy
  period. Mean results from two-week periods follow-
  ing the change are presented in Fig. 4.
    It is clear that the two types of water behave very
  differently.  For  the  high  plumbosolvency  water
  there  was  a  slow but  steady  increase 'in lead
  concentration,  eventually  reaching  the  normal
  undosed  lead  level  about 30  weeks  after the
  cessation  of dosing. With the low plumbosoivency
  water however, there  was a relatively rapid effect.
  and  lead values higher than normal undosed water.
  occurred within four weeks. Subsequently there was
  a slow decrease to typical low  plumbosolvency
  values.
   The pipes were then examined by SEM. The final
  lead values were similar to those normally obtained
  from the relevant waters indicative of the presence
  of basic and normal lead carbonates respectively.
         2-WEEK MEAN LEAD
           CONCENTRATION
              • fog Pfa/o
                                         However,  both the surfaces consisted  of  small
                                         smooth. round  'phosphate-type"  granules, and bv
                                         EDS analysis,  phosphorus and calcium were still
                                         present. The well-phosphated surface, which has
                                         very  low solubility, had apparently not been des-
                                         troyed on cessation of phosphate dosing but must
                                         have  been encapsulated within a carbonate coating.
                                           The results in Fig. 4 illustrate that short breaks in
                                         phosphate dosing of. say. a few days are unlikely to
                                         cause large increases in plumbosoivency. It appears
                                         that the phosphated pipe deposits are  reasonably
                                         stable over short periods, resulting in  only  small
                                         rises  in lead  concentrations.  Over~Ionger periods
                                         (several weeks)  the pipe deposits gradually change.
                                         The rate of increase in lead concentrations can vary
                                         between different  pipes,  and the  prospects for
                                         plumbosolvency control by summer-only phosphate
                                         dosing appear unpromising.


                                         LONGER-TERM REDUCTION IN PHOSPHATE
                                                            DOSE

                                         Although phosphate concentrations of at least 0.6
                                         mg P/l are  necessary to establish  plumbosolvency
                                         control it is possible that once established it will be
                                         maintained by low or continuous phosphate doses.
                                         In order to test this hypothesis,  pipes which had
                                         been exposed  to high plumbosoivency water dosed
                                                                       HIGH
                                                               PLUMBOSOLVENCY
           PHOSPHATE
              DOSING
              CEASED
                                                                             LOW
                                                                     PLUMBOSOLVENCY
                                                   ~1	1	1	1	1	1	1	T
                                   8  10   12  14  16  18  20  22  24  26  28  30  32
                                          UNDOSED TIME (WEEKS)
      fig. 4.  Effect of cessation of orthophosphate dosing (after 33 weeks at 1.0 mg P/l)
  (32)  ,
DOSED-
J.1WEM. IW2. 6. June.

-------
          FeB-06-1997   12:40
USEPft
5135697172    P.07
 -COLO-lNG. CROU-. WH1NCUP AND HARWARD ON

 - at 0.8 and 1.0 mg P/l for eight weeks, and which had
 established steady lead concentrations in the water
 from the test  rig, were exposed to reduced phos-
 phate concentrations of either 0.2 or 0.4 mg P/l. The
 reduced phosphate concentrations were run for a
 total period of 44 weeks at various temperatures
 finishing at 25°C. During this  period lead concen-
 trations remained low and stable.
   The phosphate concentration in two of the pipes
 was then reduced to 0.1 mg P/l. After several weeks"
 operation the lead concentrations from these pipes
 became more variable and slightly higher than those
 operating at 0.2 and 0.4 mg P/l. However, the lead
 concentrations remained  low and similar to those
 from the other pipes.
   Subsequently,  the phosphate dosing to all  the
 pipes was ceased. Unlike  the pattern of the pipe in
 high plumbosolvency  water dose'd at  1.0 mg P/I
 shown in Fig. 4. the lead concentrations from all the
 pipes increased rapidly, reaching typical high plum-
 bosolvency values within  two weeks.
   These  results  illustrate that  plumbosolvency
 control can be maintained by phosphate doses of 0.4
 mg P/l  or less, but that breaks in dosing are then
 more important,  presumably due to the less robust
 nature of the pipe deposits.


 MIXING AND ALTERNATION OF WATERS IN
                 DISTRIBUTION

 BLENDED UNDOSED WATERS
 The effects of mixing waters, equivalent to blending
 them before distribution, 'have been  previously
 published'61. It was shown  that only 5-10% of a high
 plumbosolvency water mixed with a low plumbo-
 solvency water was  sufficient   to impart  a. high
 plumbosolvency to the mixture.  Further mixtures
 have  since  been  tested,  confirming  the  original
 conclusion.

 ALTERNATING UNDOSED WATERS
  A more usual situation in distribution is that pan
 of a distribution zone will be exposed to different
 waters for  periods  determined  by  demand  and
 pumping regimes, i.e. alternating exposure to differ-
 ent waters.  In  order to investigate  the effect of
 alternating  high and low  plumbosolvencv waters.
 various periods of alternation were investigated on
 the pipe rig. Several pairs of waters were alternated
 daily for a number of weeks in new lead pipes with
 some commencing the experiment on high plumbo-
 solvency waters and some  on low. In alf cases only
 low plumbosolvency results were recorded.
    After weekly alternation in new lead pipes, low
 plumbosolvency results were obtained  after one
 week  on the pipes  beginning  with low plumbo-
solvency water (as expected), and after two weeks
on the pipes beginning with high plumbosolvency

264
      water. Subsequently,  stable,  low plumbosolvency
      results were obtained from both sets of pipes.
         In practice, waters may be  alternated after long
      periods of pipe exposure to one type of water where
      pipe  deposits are well-established..  In  order to
      investigate this situation, pipes were exposed to high
      or low plumbosolvency waters for a period of eight
      weeks in  order  to  establish  longer-term  pipe
      deposits. At  the end  of  this  period~ lead concen-
      trations were typical of the waters used. The waters
      in some of the pipes were then alternated weekly.
      Within two weeks of alternation only low plumbo-
      solvency  results were obtained. In other pipes the
      waters were alternated on a 6 day/1 day basis, with
      the same result;
          In  order to  investigate even  longer-term
      changes,  pipe deposits from low and high plumbo-
      solvency waters were established over 33 weeks and
      then the waters interchanged in some of the pipes.
      In the pipes originally  exposed to high plumbosol-
      vency. low plumbosolvency  results were obtained
      within two weeks. Contrastingly, the pipes originally
      exposed to low plumbosolvency water showed only
      slight increases in lead concentrations over the 27-
      week period of the experiment.
        It can therefore be concluded that  low plumbo-
      soivency pipe deposits  are much more stable than
      high piumbosolvency pipe deposits,  and  that  in
      areas where high and  low plumbosolvency  waters
      alternate  only low plumbosolvency  lead concen-
      trations can be expected.


      MIXING AND ALTERNATION OF PHOSPHATE DOSED A.ND
      UNDOSED WATERS

        When a high plumbosoivency water dosed  with
      phosphate is blended with  an undosed  low plumbo-
      solvency water before distribution, the  phosphate is
      diluted and the concentration may drop below that
      necessary to establish phosphated pipe deposits1"'.
      Due to the dominance of  high plumbosolvency in
      mixtures,  control may not  be established in such a
      situation and it will be necessary also to dose the low
      plumbosolvency water to raise  the  phosphate con-
     centration in the mixture  to above 0.6 me P/l in
     order to effect control.
       When such waters were alternated weekly, stable
     lead values were not obtained quickly,  only steady-
     ing  after  about 16  weeks.  Lead concentrations
     remained low and. on average, decreased throuch-
     out the experiment. However,  each time that the
     water was changed to an undosed low plumbosol-
     vency water, the lead concentrations became erratic
     and  early in  the experiment showed  some values
     higher than those normally obtained for low plum-
     bosolveney waters. This effect reduced with con-
     tinued running. It should therefore be possible to
     alternate  phosphate-dosed high  plumbosolvency
     and undosed low plumbosolvencv waters in distribu-
     tion  and still  obtain  low plumbosolvency results.

                                  J.IWKM. CM2. 6. June.
       Howevei
       be careft.
       is mainu
       The exp*
       were pei
       that the
       different
       of  their
       extraneo
       experimt
       exhumec
       minimun
       relevant
       ments  ii-
       concentr
       lead pip«
       using ne
       reach st<
       twenty  t
       phospha
       several  \
        The resi
        using un
        potable
        high plui
        phospha
        essential
        plumbos
        time tak


           INFI
        As part
        process.
        exchangi
        tested fc
        Isleham
        dosed at"

        J.IWEM. I

-------
FEB-06-1997   12:41           USEPfi

                             PLUMBOSOLVENCY EFFECTS AND CONTROL IN HARD WATERS

           ^TABLE I.  EFFECT OF ION-EXCHANGE NITRATE REMOVAL ON ISLEHAM WATER
                                                                                               5135697172    P.08
Water
Combined raw
Raw - added chloride
IEX treated water (mean mix)
IEX treated water with added nitrate
IEX treated water with added sulphate
IEX treated water with hoth additions
IEX treated water (end of run)
Nitrate
(mg/l NO,)
97
97
2U
97
20
97
35
Sulphate
(me/I)
X7
X7
,X
X
X7
X7
58
Chloride
(me/I)
ftX
177
177
177
177
177
121
Plumbosolvcncy test results
(UR/I Pb)
KW. KG : HIGH
KXJ. 97 : HIGH
:s. 31 : LOW
2X. 29 : LOW
31. 31 : LOW
27. 29 : LOW
3X. 33 : LOW
  However, the experiments indicate that this should
  be carefully investigated to ensure that good control
  is maintained.
                 OLD LEAD PIPES

  The experiments described in the preceding sections
  were performed using new lead pipe. It is possible
  that the old lead pipes in distribution will behave
  differently, due to the more heterogeneous  nature
  of their old surface deposits  which may include
  extraneous and organic  matter.  Several of the
  experiments were therefore repeated using carefully
  exhumed old lead pipes which were subjected to
  minimum  disturbance and kept  filled  with the
  relevant water.  From  the  results of these experi-
  ments it can be concluded  that  the  final lead
  concentrations obtained on various waters using old
  lead pipes are not significantly different from those
  using new  lead pipes. However, the time  taken to
  reach stable values is generally much longer, up to
  twenty times as  long. With  ol'd pipes the  effect of
  phosphate-dosing  can  continue  to increase  for
  several years.


               SURFACE WATERS

 The results illustrated in this paper  were obtained
 using underground waters.  AH the surface-derived
 potable  waters in the Anglian Water region have
 high plumbosolvency. The final test results from the
 phosphate  dosing of surface-derived waters are
 essentially the same as those derived from  hieh
 plumbosolvency underground water. However, the
 time taken to reach similar lead levels is longer.


    INFLUENCE OF ORGANIC MATTER ON
              PLUMBOSOLVENCY

 As part of the commissioning procedure for a new
 process,   water  from  the  "Anglian  Water   ion-
 exchange nitrate-removal  planf at   Isleham  was
 tested for plumbosolvency. The untreated water at
 Isleham has high plumbosolvency and is phosphate-
 dosed after nitrate removal. The water from  the ion-

J.tWEM. 1

-------
       FEB-06-1997   12:43
USEPft
                                                                                                5135S97172    P. 10
                                     PLUMBOSOLVtNCY trr-ECTS AM) CONTROL IN HARD WATtKS
                          TABLE III.  EFFECT OF REMOVAL OF ORGANICS FROM HIGH
                      Pl.UMBOSOLVENCY WATERS WITH FRESH GRANULAR ACTIVATED C\RBON
Highplumbo$oivcn<:y water
islchum (raw)
Typical groundwaier
Typical surface water
TOC (mg C/l)
Umrcutcd
1 .5
1.2
4.2 '
GAC filtered
'!.?
(1.2
C.3
Test result (ug/1 Pbl
Untreated
I»5
107
IIS
GAC tillered
5X
30
:s
  a real inverse relationship. Results at Boston over a
  much longer period indicated higher summer lead
  concentrations {20 ug/I compared to 10 ug/1 in the
  winter).
    Short breaks in phosphate dosing do not lead to
  significantly increased lead concentrations in waters
  dosed at 1.0 mg P/l; however, long breaks lead to a
  re-establishment of plumbosolvent  pipe deposits.
  Waters dosed at.concentrations lower than 1.0 mg
  P/l show a  more  rapid  return  to  plumbosolvent
  conditions.
    Once phosphated pipe deposits have been estab-
  lished, they  can  be maintained in the pipe rig by
  phosphate concentrations as low as  0.1  mg P/l.
  However, breaks in phosphate dosing at these lower
  maintenance concentrations result in a rapid rise in
  lead solvency, returning to the original high plumbo-
  solvency results within two weeks. Thus if mainten-
  ance  dosing  is adopted,  it becomes important to
  ensure that interruptions to dosing are not longer
  than  a day or two. Longer periods will lead to  a
  return, to plumbosolvent conditions necessitating at
  least a one-year period of dosing to 1.0 mg P/l to re-
  establish phosphated deposits before maintenance
  dosing can be reintroduced. Phosphate doses as low
  as 0.1 to 0.2 mg P/l may not penetrate to the ends of
  a distribution system, leading to loss of control at
  the system extremities.  In practice, therefore, ir is
  expected that in  the longer term phosphate doses
  may be  reduced  by up to 50%. thereby lowering
  costs  whilst maintaining effective plumbosolvency
  control.  Lower  doses may  be possible  in some
  systems. However,  care must be taken to maintain
.  an  adequate phosphate  concentration at  the ex-
  tremities of the distribution  system and  to avoid
  breaks in phosphate dosing.
  • Where high and  low plumbosolvencv waters are
  blended before distribution, even a small proportion
  of high plumbosolvencv water will give a plumbo-
  solvent  mixture.  It  will  therefore  normally  be
  necessary to  phosphate dose  both waters (or the
  mixture) in these circumstances. In the more normal
  situation of  waters mixing in distribution in  an
  undefined manner  and  some  parts of the  system
  receiving alternating waters, it has been shown that
  when high and low plumbosolvencv waters alternate
  the low plumbosolvencv pipe deposits are the more
  stable.  In such a  situation  low  plumbosolvencv
  results ure obtained whichever water  is present.
 These  results explain why some parts of certain

 J.JWEM. IW2. 6. June.
         distribution  systems  give   low  plumbosolvencv
         results, even though they more often contain high
         plumbosolvency water than low. In practice, how-
         ever,  there  will  always  be part of  the  system
         exclusively supplied by high plumbosolvency water
         and therefore this water will be-phosphate dosed.
         Alternation of this water with the low plumbosol-
         vency water may give erratic initial lead concen-
         trations but should enable control to be maintained
         in the long term.
           The general introduction of phosphate into plum-
         bosolvent zones in Anglian Water has given success-
         ful control of lead solvency. The mean reductions
         (thirty-minute stagnation lead levels) at fixed points
         are entirely consistent with the trials and experimen-
         tal data. 45% for 3-6 months dosing, increasing to
        • 63% for 9-12 months. The lead concentrations now
         achieved easily comply with the requirements of the
         EC  drinking  water  Directive1'"  and  the  more
         stringent requirements  of the Water Supply (Water
         Quality) Regulations 1989'"".
           Experimental evidence from  high  plumbosol-
         vency waters, where passage through ion exchange
         or GAC columns ,gave low plumbosolvency. and
         from the addition of peat extract to low plumbosoi-
         vency waters,  making them  highly plumbosolvent.
         indicates that low concentrations of humic acids may
         be the cause of high plumbosolvency in hard waters.
         In practice this rinding is unlikely to alter piumbo-
         soivency control measures in Anglian Water in the
         foreseeable future  as phosphate dosing is much less
         expensive, at present, than the removal of organic
         compounds.

                        CONCLUSIONS

         1. In order  to  establish effective piumbosoivency
           control, initial doses of phosphate of at least 0.6
           mg P/l {as phosphorus) are required.
         2. Lower continuous phosphate doses can maintain
           Control.
         3. Short breaks in phosphate dosing of  1 or 2 days
           should not affect plumbosolvency control.
         4. The effects of longer breaks in phosphate dosing
           depend on  the  phosphate dose. High plumbo~-
           solvency conditions can return within two weeks
           at low phosphate doses.
         5. Where  high and low piumbosoivency water are
           blended before  distribution both waters,  or the
           mixture, must be phosphate dosed.

                                                   267

-------
FEP-06-1997   12:42
                                             USEPfl
                                                                                                  5135697172    P.09
   COtCING. CROLL. WHINCUP AND HARWARD ON

   water.  Increasing  quantities  of  this extract  were
   added to 25 1 of low plumbosolvency water and the
   resulting solutions tested for their plumbosolvency.
   It will be seen that addition of sufficient quantities of
   the peat  extract  gave  higher   plumbosolvency.
   During these experiments the TOC of the water
   tested ranged from 0.6 mg C/l in the control water to
   1.7 mg C/l in the water with 1000  ml of peat-extract
   added.  Addition of the peat extract to the ion-
   exchange treated  water from Isleham gave  high
   plumbosolvency results (Table II).

    TABLE II.  EFFECT OF ORGANIC ADDITIONS TO Low
     PLUMBOSOLVENCY WATERS USING A STANDARD PEAT
                      EXTRACT
Addition to 25 1 low
plumbosolvency waier
no addition
+• 10 ml peat water
+• 50 ml peat water
* 75 ml peat water
+ 1(X) ml peat water
+ 250 ml peat water
* ItXKI ml peat water
Isleham IEX treated water
•>• HXI ml peat water
Plumbosolvencv test result
f[jg Ph/n
30 \ LOW
26 f Plumbosolvcncy
5!
S2 %
V7 1 HIGH
1 14 [ Plumhosolvencv
tIC }
29 LOW
X2 HIGH
 REMOVAL OF ORGANIC COMPOUNDS FROM HIGH
 PLUMBOSOLVENCY WATER

   The organic content of three typical high plumbo-
 solvency waters was reduced by passage at a slow-
 rate (1 1/h) through a column (500 mm x 400 mm
 dia.)  of fresh granular activated carbon (GAC).
 TOC analysis  confirmed that a substantial propor-
 tion of the organics present had been  removed. The
 results of testing these waters for plumbosolvencv
 are shown in Table III. It will be seen that in ail
 cases  the  high plumbosolvency waters had  been
 reduced to low plumbosolvency. Thus two different
 methods  of removing  organic  compounds  from
 water have been shown to reduce plumbosolvency.
 If humic matter contains the factor promising high
 plumbosolvency. it would be expected that~GAC
 columns would have a short  lifetime for plumbo-
 solvency reduction,  as  they  have   only  a  short
 lifetime for the removal of "humic  material. Con-
 tinued use of the GAC  column confirmed that  its
 capacity for plumbosolvency reduction was rupidlv
 exhausted.
  Although the above initial  experiments are not
 conclusive, and further experiments to  investigate
 the  influence  of organic compounds on  plumbo-
 solvency  are  being  carried out. they  give a strong
 indication that the main factor causing  high pfumbo~-
 solvency  in hard waters  is their content of humic
 substances.
266
                                           SURVEY OF ANGLIAN WATER POTABLE WATERS

                                             The  TOC content of Anglian Water potable
                                           waters was compared with their plumbosolvency. It
                                           was  found that  waters with  the highest  TOC
                                           concentrations (>3 mg C/l) were all pluinbosolvent.
                                           This  category includes all the Anglian  Water sur-
                                           face-derived water supplies.  Waters with  very low
                                           TOC («J.6 mg C/l) were mostly of low plumbo-
                                           solvency.  However, for the  majority of Anglian
                                           Water supplies TOC  content  is  not  a  reliable ,
                                           indicator of plumbosolvency. This is not surprising if
                                           humic substances or some fraction or component"of
                                           them  are responsible for the  effect, as  the contri-
                                           bution of humic matter to the TOC will vary with
                                           the water source.
                                                           DISCUSSION

                                          The plumbosolvency of hard waters is determined
                                          by the type of crystalline  deposit formed  on the
                                          inside of the  lead pipe. These  deposits can  be
                                          identified using electron microscopy. The solubility
                                          of the deposits  is temperature  dependent.  The
                                          addition of orthophosphate  to high plumbosolvencv
                                          waters causes  a  phosphated pipe deposit of low
                                          solubility to be formed. In order to form an effective
                                          deposit, it has been shown  that phosphate concen-
                                          trations of at least 0.6 mg/l are required. The higher
                                          the phosphate concentration (up to 1.6 mg P/lf the
                                          more rapidly the deposits are formed.
                                            Old lead  pipes contain  thick  surface deposits
                                          which have been shown to change more slowly than
                                          the younger deposit in the new lead pipes normally
                                          used in the  pipe rig tests; however, the long-term
                                          effects of treatment  are the same.  Thus" in  a
                                          distribution system where the phosphate takes some
                                          time to penetrate to the ends of the system and only
                                         old lead pipes are present, it is expected that initia!
                                         phosphate concentrations of at least 1.0 mg P/l will
                                         need to be maintained in order to establisrTplumbo-
                                         splvency control within a reasonably short period of
                                         time.  Indeed in most distribution "systems, where
                                         phosphate may be reduced by adsorption or absorp-
                                         tion on deposits in non-lead pipes, it may not be
                                         possible to  achieve  the 0.6 mz  P/l.  the  lowest
                                         concentration required to establish plumbosolvency
                                         control, at the ends of the system unless phosphate
                                         concentrations above this value are closed.
                                         _ The  minimum period  to establish control in  the
                                         held will depend upon the situation, but experience
                                         at Boston and Sleaford"" indicates that a year  is not
                                         unreasonable. !n subsequent years further  small
                                         reductions in  lead solvency will be achieved. In test-
                                         rig  results. lead concentrations from  phosphated
                                         lead pipes  appeared to decrease with  increasing
                                         temperature.  This apparent effect was thought to be
                                         due to the  more rapid establishment  of highly
                                         phosphated deposits at  the higher temperatures in
                                         the  limited time-scale of the experiments rather than

                                                                     J.IWF.M. 1W2. 6. Juno.
 a real
 much
 conce
 winte
   She
 signiti
 dosed
 re-est
 Watei
 P/l s|-
 condi'
   On<
 lished
 phosp
 Howe
 maint
 leads
 solver
 ance
 ensur-
 than
 returr
 least ;
 establ
asO.l'
a dist
the s\
ex pec
may 1
costs
contrt
syster
an  nt.
tremii
break
  Wh
hlend
of his
solver
neces:
mixtu
situat
undet
receK
when
the lo
stable
result
These

-------
      FEg-06-1997   12:44
USEPfl
 - PLUMBOSOLVENCY EFFECTS AND CONTROL IN HARD WATERS

  6. Where  high  and low plumbosolvency  waters
     alternate  in distribution, low plumbosolvency
     results are obtained.
  7. Where  phosphate-dosed high plumhosolvency
     and  low  plumbosolvency  waters  alternate in
     distribution, low  plumbosolvency  results  are
     obtained after a short period of erratic  behav-
     iour.
  8. Initial results indicate that humic substances (at
     low concentrations)  may be the cause of high
     plumbosolvency in hard waters.
                                                            5135697172
                                                                    t
      P. 11
                              REFERENCES
              ACKNOWLEDGEMENTS

  The authors wish to thank  the  Director  of Quality of
  Anglian Water for permission  to  publish this paper.
  Grateful thanks are also expressed to Mr C. R. Hayes. Mr
  D. N. Harris and all the-other Anglian Water and former
  Anglian Water staff who provided assistance, to Mr R.
  Gregory of the WRc and  to Mr J.  G. Watson  of SCM
  Chemicals Limited for scanning electron microscopy.
                    DISCLAIMER

 The views expressed in this paper are those of the authors
 and do not necessarily represent  the views of Anglian
 Water Services Ltd.
             (I) DEPAXiMKNr OF tut  ENVIRONMENT.  Report of the Expert
                  Advisory Group on identification and Monitoring. Lead in
                  Poiable Water. Technical Note No.  I. 19X0 (official use
                  only).
             (2) SHHIHAM.  I.. AND JACKSON.  P. ). The scientific basis for
                  control of lead  in drinking water. J. last. Wat. Eng. Sci
                  mi. 35. (6). 491-515.
             (3( COLUNO, J. H.. AND  HARMS. D. N.  The use of orthophos-
                  phatc for plumbosolvency control  of hard graundwatcrs.
                  Paper presented to  Scicmiric Section  of the Institution of
                  Water Engineers and Scientists. Huntingdon. 19K5.
             (4) GREGORY. R.. AND JACKSON. P. J. Reducing Lead in Drinking
                  Water. WRc Regional Seminars.  May-June  19X"»  WRc
                  Report 2JV-S. 19X3.
             (5) DEPARTMENT or run ENVIRONMENT. Report of the  Expert
                  Advisory Group on Remedial Water Treatment for Reduc-
                  ing Lead Concentrations in Tap Water. Lead in Potable
                  Water. Ttrhaical \ate No. 5. 19K4 (official use onlv).
             (ft) COLLING. J. H.. WHINCLT. P. A. £.. AND HAVES. C.  R. The
                  measurement of piumbosolvency propensity to guide the
                  control of lead in lapwaters. J. Irani Wai. &' Envir  Manet
                  19X7. |. (3). 263-;69.
             (7) JACKSON.  P. J.. AND SHEIHAM. I.  Calculation of Lead
                  Solubility in Water.  WRc Technical Report TR 152. [980.
             <*W7»EEC). Official Journal L229. August I9KO.
           (10) DEPARTMENT OF THE ENVIIIONMENT. The Water Supply (Water
                 Quality). Regulations 19S9. Siauaory Instruments tVf». No.
                 11-17. Her Majesty's  Stationery- Office. 1989.
IV

T
T
si

al
«

tc
tf-
N
K
rr
268
                                                                                       J.IWEM. 1992. 6. June.
                                                                               *i'-c  ••'•• •  ^'"^^-'---^^^

-------
     DISTRIBUTION  SYSTEMS
    Although selection of corrosion control optimization strategies is more art than
      science, utility experiences can provide a basis for rational decision-making.
Donna IVI. Dodrill
and Marc Edwards
  ";Utility experience'under the Lead and Copper Rule was
 'examined to provide improved insight into corrosion control.
 , Average 90th percentile lead concentrations were highest in very-
 ' low-alkalinity waters (<30 mg/L as CaCO3) at utilities that did not
  use inhibitors; lead release was significantly reduced at higher
  . alkalinities. Average lead releases were 20-90 percent lower for
 ' utilities using phosphate inhibitors (orthophosphates,
  polyphosphates, and blended phosphates) in very-low-alkalinity
 ["waters than for utilities not using inhibitors. At alkalinities of
 • 30-74 mg/L as CaCO3 and at pH values > 7.40, it appeared that
 . P^YPljosphate inhibitors had adverse effects on average lead
 ^release.ptffities with pH < 7.40 and high-alkalinity waters had
 {/the highest copper concentrations. Phosphate inhibitors were
 ' 'usually beneficial in mitigating copper release; however, most
  benefits were at utilities with pH < 7.80 and alkalinity > 90 mg/L
 il.a? 
-------
             Average 90th percentlle lead release and percentage
             of utilities exceeding the lead action level for utilities
             not adding phosphate Inhibitors

                                          <7.40 -.--»;-;?«;F*^
                                              -^jSsfa^fe^i
considered by assigning the utilities to specific pH
and alkalinity categories. pH was divided into four
ranges: (1) pH < 7.40, (2)  pH 7.40-7.80, (3) pH
7.81-8.40, and (4) pH > 8.40. Utilities in each pH
category were further subdivided according to alka-
linity: (1) < 30, (2) 30-74, (3) 75-174, and (4) > 174
mg/L as CaCO3. These divisions maintained an
approximately even distribution of the data, ensured
an ample number of points within most pH-alkalin-
ity categories, and allowed for qualitative classifications
of potable water according to pH and alkalinity (i.e.,
very low, low, moderate, and high).
   After sorting the data according to these cate-
gories, average 90th percentile lead and copper con-
centrations were calculated using data from all util-
          ities in each category. In addition, the
          percentage of utilities in each pH and
          alkalinity category that exceeded the lead
           or copper action level (i.e., 0.015 mg/L Pb
           or 1.3 mg/L Cu) was also examined. Para-
           metric statistics were used to rigorously
           evaluate the significance of observed
           trends for selected data at the 95, 90, and
           85 percent confidence levels (i.e., the 5,
           10, and 15 percent significance levels,
           respectively) using a standard comparison
           of means test.2 Given the lognormal dis-
           tribution of the data, it was first necessary
           to normalize the data using the log-trans-
           formation of Benjamin and Cornell.3
           Details of the statistical analysis are pro-
           vided elsewhere.4
             It is important to qualify the work by
           noting several points.
              •  Given that  this analysis compares
           aggregate data of utilities with differing
           water qualities,  results will not apply
           quantitatively to a specific utility. Indeed,
           even the qualitative trends must be
           viewed with caution when viewed from
           the perspective of a single utility.
              •  Terms such  as "increase" and
           "decrease" refer to comparative changes
           in the pH and alkalinity categories for the
           aggregate data. For example, the phrase
           "increasing the pH from 7.0-7.4 to pH >
           8.4"  does not mean that a given utility
           or utilities made  such changes and
           observed the cited effect. Rather, it is a
           comparison of averaged corrosion by-
           product release for utilities having pH val-
           ues of 7.0-7.4 with those utilities having
           pH values > 8.4.  The cited trends might
           apply if a  given utility did make such
           changes, but such conclusions cannot be
           made from this work unambiguously.
              •  pH and alkalinity values are six-
           month averages  calculated from water
           released to the distribution system. Some
           utilities reported pH and alkalinity data
           averaged from a  number of distribution
sites or point-of-entry (POE) values. In some
instances, pH and alkalinity values were derived from
samples collected over several months or days.
Regardless of how the data were taken, variations in
pH and alkalinity are  expected in any distribution
system water, and this analysis cannot account for
these effects.

Lead corrosion
   Of the 397 total survey respondents, 365 utili-
ties reported their first-round  90th percentile lead
concentrations. The  effects of  pH, alkalinity,
inhibitors, calcium, and temperature on 90th per-
centile lead values are based  on their responses. Use
of phosphate inhibitors was  given special consider-
                                                                                       JULY 1995  75

-------
              Percent reduction In lead release as a result of the indicated
              increase In alkalinity for utilities not adding phosphate inhibitors
                                 Alkalinity Chango—tng/L 03 CaCO,
                        <30 to 30-74
                                       30-74 to 75-174
                                                        75-174 to >174
                           ••21*
                           '•51*.
                            74*
                           ..8 r
                                Reduction In Lead Release—percent
•'18':'.vr-
          '—'--' atittie 95 percent corifidenceTnte'n4l"

 ation in this analysis—the term "inhibitors" refers
 only to phosphate-based inhibitors (e.g., orthophos-
 phate,  zinc orthophosphate, hexametaphosphate,
 polyphosphates, and various blends) in subsequent
 discussion.
    Utilities not using phosphate-based inhibitors.
 For utilities not adding inhibitors, average 90th per-
 centile lead levels were highest in very-low-alkalin-
 ity (< 30 rng/L as CaCO3) waters (Figure 1). Raising
 pH to < 8.40 did not appear to strongly affect lead
 release in very-low-alkalinity waters, although
 increasing pH above 8.40 did appear to reduce lead
 release.
    Lower lead levels (at a given pH) were associated
 with higher alkalinities in each pH category with only
 one exception (Table 1). Differences in lead levels in
 each category were represented by calculating percent
 reductions. In this study,
 percent reduction  was
 defined as [(Initial Pb -
 Final Pb)/Initial Pb x 100
 percent]. Thus, in Table 1,
 the initial Pb was the lower
 of the two alkalinity cate-
 gories being compared in a
 given alkalinity change. As
 shown in Figure 1, increased alkalinity of 30-74 mg/L
 as CaCO3 (as compared with alkalinity < 30 mg/L as
 CaCO3) yielded statistically significant lower lead
 release a{ the 95 percent confidence level when pH
 was < 8.4 (Table 1).
   For utilities that did not add inhibitors, the per-
 centage exceeding the lead action level seemed to be
 correlated with differences in both alkalinity and pH
 (Figure 1). In the lowest pH and alkalinity category,
 utilities had an 80 percent likelihood of exceeding
 the lead action level. This likelihood was less in higher-
 alkalinity or higher-pH categories almost without
 exception. As observed for lead release, a large reduc-
 tion in the percentage of utilties exceeding the lead
 action level was realized at the transition from < 30
 to 30-74 mg/L as CaCO3.
   Utilities adding phosphate-based inhibitors.
 The use of  phosphate inhibitors may reduce lead
 concentrations in drinking waters.5-9 The initial analy-
 sis of the data did not distinguish between the vari-
                                      ous  types  of  phosphate
                                      inhibitors or doses used. That
                                      is,  all utilities that added
                                      orthophosphate,  blended
                                      orthophosphate, or polyphos-
                                      phate  were  classified  as
                                      "adding inhibitors." In the
                                      1992 survey,  116 utilities
                                      reported using  phosphate
                                      inhibitors;  261  utilities
                                      reported they did not use
                                      phosphate inhibitors. With-
                                      out considering the effects of
                                      either pH or alkalinity, differ-
                                      ences between average  90th
                  percentile lead were insignificant between utilities
                  regardless of whether they added inhibitors.
                     Any comparison of utilities that add inhibitors
                  with those that do not add inhibitors may be subject
                  to systematic biases that are impossible to quantify. At
                  one extreme,  most utilities might be adding phos-
                  phate-based corrosion inhibitors to control problems
                  with iron corrosion. Consequently, the data might
                  be overweighted in pH-alkalinity categories that prop-
                  agate red water problems. Nonetheless, there is no
                  reason to suspect that this bias would invalidate use
                  of the database for the examination of lead and cop-
                  per corrosion problems.
                     At another extreme, utilities that add inhibitors
                  might be doing so to mitigate serious problems  with
                  lead or copper corrosion. In. this case, the utilities that
                  add inhibitors might represent only the most prob-
ncreased  lead release appeared to occur
at higher alkalinity when inhibitors
were used.
                 lematic waters with respect to lead and copper corro-
                 sion. Given the relative lack of concern regarding lead
                 and copper corrosion by-product release at the tap
                 before enactment of the Lead and Copper Rule, how-
                 ever,  there is no strong reason to suspect this bias
                 exists in the data set used from the first round of sam-
                 pling. Also, the database cannot quantify the effect
                 of time on the development of passivating films or
                 decreasing lead leaching (by physical depletion).
                    Keeping these potential problems in mind, it is
                 still illustrative to make comparisons between utilities
                 that currently add inhibitors and those that do not.
                 Particular attention was paid to identifying pH-alka-
                 linity regimes in which inhibitors appeared to have
                 beneficial, detrimental, or no effect on lead and cop-
                 per corrosion by-product release. The comparison is
                 used to highlight water qualities for subsequent lab-
                 oratory confirmation studies that unambiguously test
                 these identified trends in inhibitor performance.
                 Trends in the data were also examined for consis-
76 JOURNAL AWWA

-------
tency with theoretical predictions that
guide inhibitor use.5-7-10
   For utilities that added inhibitors
between pH 7.40 and 8.40,  lead release
was adversely affected by increasing alka-
linity from < 30 to 30-74 mg/L as CaCO3
(Figure 2; Table 2, negative numbers).
These trends were statistically significant
at the 85 percent confidence interval. In
contrast,  at pH < 7.40 increasing alka-
linity from < 30 to 30-74 mg/L as CaCO3
had the opposite effect, leading to a 53
percent reduction in lead release. This
observation was not supported at the 85
percent confidence level, however, given
the high standard deviation of the data in
this category.
   Trends in the percentage of utilities
exceeding the lead action level were com-
plicated by inhibitor addition. For
instance, the highest percentage of util-
ities (47 percent)  exceeded the lead
action level at 30-74 mg/L as CaCO3 and
at pH 7.40-7.80 (Figure 2). Lead con-
centrations  did  not exceed the action
level in systems using inhibitors at very
 low alkalinities (<30 mg/L as CaCO3)
 above pH 7.80, and the use of inhibitors
 generally reduced lead leaching through-
 out the low-alkalinity range.
    Inhibitor effectiveness. An exami-
 nation of the difference between systems
 with and without inhibitors is instruc-
 tive, as shown by the following equation:

     Reduction in lead release  associated
     with inhibitors = Average lead release
     without inhibitors - Average  lead
     release with inhibitors
     In this equation a positive number
  means  an improvement (reduced lead  :•
  release), whereas a negative number  ;
  means  ah adverse effect (increased lead
  release) associated with inhibitors. This
  general formula was used to evaluate the effect of
  phosphate-based inhibitors on average 90th perccntile
  lead release as well as on the percentage of utilities
  exceeding the lead action level.
     Beneficial effects of inhibitors were observed only
  in the lowest alkalinity category (<30 mg/L as CaCO3)
  at all pH values (Figure 3). Conversely, in  several
  instances in the 30-74-mg/L as CaCO3 alkalinity cat-
  egory  and at pH > 7.40, total lead release and the
  percentage of utilities exceeding the lead action level
  were apparently higher. Differences between systems
  that add inhibitors and those that do not add inhibitors
  were inconsequential at alkalinities > 74 mg/L as
  CaCO3.
      In  the lowest alkalinity category (<30 mg/L as
  CaCO3) average lead release was 20-90 percent lower
Average 90th percentite tead release and percentage
of utilities exceeding the lead action level for utilities
adding phosphate inhibitors
                           r.46-^8b;%M^p^i?«*a:;
                           &&&^&£@g%&S3Sg%*
for utilities using inhibitors compared with those not
using inhibitors (Table 3). These results were all sig-
nificant at the 95 percent confidence interval, indi-
cating that addition of inhibitors may be very effective
in reducing lead release in very-low-alkalinity waters
regardless of pH (Table 3). Inhibitors did not produce
a statistically significant reduction in lead release (at
the 95 percent confidence level) in any other pH-alka-
linity category tested. In fact, in the alkalinity range of
 30-74 mg/L as CaCO3 and at pH > 7.40, inhibitors
 had higher average lead release (negative numbers
 in Table 3). These  increases were all significant at
 either the 95 or 85 percent confidence interval. Thus,
 the indication is that inhibitors exacerbate 90th per-
 centile lead release when used in waters with pH >
 7.40 and alkalinity 30-74 mg/L as CaCO3.
                                                                                           JULY 1995  77

-------
              Percent reduction in lead release as a result of the indicated
              Increase In alkalinity for utilities that add phosphate inhibitors

                                 Alkalinity Change—mg/L tat CaCO3 ' /,-'
                                                         75-174 to >174
                                                         "~ic*.-.rv	
              Percent reduction in lead release for utilities adding phosphate
              inhibitors compared with those not adding phosphate inhibitors
                                                               >174
                                  Reduction In Lead Release—percent
   Explanations for increasing lead levels at moder-
ate alkalinities for utilities adding phosphate-based
inhibitors include the possibility of inadequate phos-
phate dosing to form protective films.  That is,
orthophosphate is known to inhibit the growth of
basic carbonate and oxide films on copper pipe.11 It
is not known whether it may also inhibit the growth
of cerrusite (PbCO3) or hydrocerrusite [Pb3(CO3)2OH].
For example, utilities may not have the optimal
inhibitor dosage to form a protective layer of hydroxy-
pyromorphite [Pbs(PO4)3OH] or lead phosphate
[Pb3(PO4)2]. Phosphate may also be interfering with
other film formation.
   Detailed examination of inhibitor use. Of the
116 utilities using phosphate inhibitors, there was an
approximately even split
between  utilities  using
orthophosphate and those
                               Phosphate-based inhibitors
                            had significant effects only in
                            the two lowest alkalinity cat-
                            egories (<30 and 30-74 mg/L
                            as CaCO3). In the lowest pH
                            region (<7.40), average 90th
                            percentile lead release for util-
                            ities with alkalinities <30
                            mg/L as CaCO3 and using
                            polyphosphates was more
                            than double that for utilities
                            applying  orthophosphate
                            alone  (Figure 4). For  poly-
                            phosphates, average lead re-
                            lease was even higher than
                            was observed for utilities not
                            adding inhibitors. Thus, in this
                            pH-alkalinity category at least,
                            it seems reasonable to at-
                            tribute beneficial effects of
                            inhibitors to orthophosphates
                            and either no or adverse ef-
                            fects to polyphosphates.
                               Interestingly, in the  same
                            alkalinity category but at pH
                            7.4-7.8,  the opposite effect
                            was observed (Figure 4); util-
                            ities using polyphosphates had
                            an average corrosion by-prod-
                            uct release about three to four
                            times  lower than  utilities
       using orthophosphates. However, the average lead
       release was lower at utilities using either type of
       inhibitor than at utilities not using inhibitors. For
       utilities with alkalinities of 30-74 mg/L as CaCO3 at
       pH 7.4-7.8, there did not appear to be a significant dif-
       ference between the two types of inhibitors in per-
       formance (but inhibitors did have a statistically sig-
       nificant adverse effect in this pH-alkalinity category).
          In the two highest pH categories (7.81-8.40 and
       >8.40) there was either insufficient data or a highly
       uneven division between inhibitor  types, so an
       inhibitor-to-inhibitor comparison is not possible. Nev-
       ertheless, this overweighting is still noteworthy. For
       instance, at pH >8.40 and alkalinity 30-74 mg/L as '.
phates or did not specify the
type  of inhibitor added
(Table 4). For the inhibitor
effects cited as statistically significant for lead in the pre-
ceding section, the types of inhibitors used in each
identified pH and alkalinity category were then care-
fully scrutinized. That is, the information on inhibitor
types was used to clarify whether beneficial or adverse
inhibitor effects could be attributed to orthophosphate
alone, polyphosphates alone, or both in combination.
release is reduced by higher pH
presence and absence of inhibitors.
       CaCO3 when inhibitors increased average lead release
       by 91 percent (Table 3), all 13 utilities used polyphos-
       phate inhibitors. Thus, it seems reasonable to attribute
       the significant adverse effects observed for inhibitors
       in this category to polyphosphate exclusively. The
       finding that utilities are using polyphosphate chem-
       icals above pH 8 apparently contradicts  the notion
78 JOURNAL AWWA

-------
that phosphate inhibitors are used only
near neutral pH.12
   In the lowest alkalinity range (<30
mg/L as CaCOj) and pH values > 7.81,
where significant beneficial effects of
inhibitor were observed, all of the utili-
ties used polyphosphate inhibitors. Lead
release was significantly lower for utilities
in these two pH categories; however, only
three utilities used phosphate chemicals
(two in the pH 7.81-8.40 region and one
in the pH > 8.40 region). In the alkalin-
ity category 30-74 mg/L as CaCO3 at pH
7.81-8.40 in which lead release was
increased by inhibitors, five utilities used
orthophosphate and only one utility used
polyphosphate.
   Potentially adverse effects of poly-
phosphate addition on lead release have
been noted by others. Holm and Schock,
using a discrete two-ligand model, pre-
dicted the effects of polyphosphate addi-
tion at pH 8.0 and 40 mg/L Ca as a func-
tion of alkalinity.13 The authors modeled
solutions in equilibrium with either
hydrocerrusite or hydroxypyromorphite.
In both situations, waters treated with a
proprietary  liquid   formulation  of
polyphosphate (consisting of 40 percent
orthophosphate) dosed at 1 mg P/L were
predicted to have significantly higher
plumbosolvency than untreated waters
(or  waters  dosed  with  0.4  mg/L
orthophosphate in the case of hydrox-
ypyromorphite precipitation). Similarly,
Sheiham  and Jackson demonstrated that
lead  release was higher in new and old
lead pipes (at pH 6.5 and 7.5, respec-
tively) for two polyphosphate chemicals
than for orthophosphate.8  All three
inhibitors were dosed at 1 mg/L in a low-
alkalinity (< 10 mg/L as CaCOj) water.
    Although the preceding discussion   f^.,':V    ,
considerably clarifies matters with respect   £i •      g
to the practical effects of inhibitors,   {       >
inherent shortcomings of using the re-   :
suits to guide phosphate inhibitor use   ?::• - :; -•• = '•
remain. Nevertheless, in the absence of   j~-  : •:"-::.';
systematic laboratory studies to confirm
trends, the  preceding discussion pro-   '.  •
vides a good starting point for bench or
pipe-loop investigations at individual
 utilities.  Future work and experimentation should
 examine the inhibitor question in greater detail.
    Effects of temperature. Average 90th percentile
 lead as a function of alkalinity and cold (<15°C) or
 warm (> 15°C) water was calculated in the presence
 and  absence of inhibitors (Figure 5).  Temperature
 was  taken by utilities at several distribution system
 water quality sampling locations (data were collected
 between January and June  1992). In general, lead
  Improvement In percentage of utilities exceeding
  the lead action level as a result of adding phosphate
  inhibitors
  I Comparison of average 90th percentile lead release
  ' for utilities adding orthophosphate and polyphosphate
  ; and utilities not adding inhibitors
    t§£ Orthophosphate  HI Polyphosphate .:..P| No phosphate^: jS'i_
0.050
                  y&,-.-f& PH  . 'V-r?"£v'.'*:Vv-:??!'.-:"*?•-':
                  aCp3;mimber- -•--— ^—•*--»---••.	•-
 many utilities were using that Inhibitor
  levels decreased with increasing alkalinity for both
  cold and warm distribution waters regardless of
  inhibitor use. In contrast to the conventional wis-
  dom, warm waters did not appear to induce higher
  lead release than did cold waters at a given alkalin-
  ity.14-16 If anything, the opposite effect was observed
  in the  lowest alkalinity category. In retrospect this
  trend probably should have been expected, given
  that at-the-tap monitoring precludes wide variations
                                       JULY 1995  79

-------
              Utilities using specified type
              of phosphate inhibitor
                                          ->"S--cv ft"?*-:; -,:
      •LargeTs defined as utilities sening 50.000 peocterj^SSSS- •£>•..
in temperature (either seasonal or geographical) after
an overnight stagnation period.
   Effects of calcium. Similar to the effect of tem-
perature on lead release, the effect of calcium on
90th percentile lead was considered as a function of
alkalinity. As with the pH and alkalinity data, cal-
cium concentrations were determined either at point-
of-entry locations or in the distribution systems. Two
divisions for calcium content, low (<50 mg/L as Ca)
and high (>50 mg/L as Ca) were made (data not
shown). Overall, there was not a strong trend between
lead release and calcium concentration in any sys-
tem studied.
                      Unity < 74 mg/L as CaCO3, raising pH from < 7.40 to
                      7.40-7.80 resulted in a 43-68 percent reduction in
                      average 90th percentile copper release—changes that
                      are significant at the 95 percent confidence level.
                      Raising pH from 7.81-8.40 to >8.40 always led to
                      significant reductions in copper release without regard
                      to alkalinity, whereas increasing pH from 7.40-7.80
                      to 7.81-8.40 significantly decreased copper release
                      only if alkalinity was >75 mg/L as CaCO3.
                         The described trends are magnified when viewed
                      in the context of the percentage of utilities exceed-
                      ing the copper action level in each pH and alkalinity
                      category (Figure 6). No utilities with pH > 7.80
                      exceeded the copper action  level (Figure 7). The
                      highest percentage of utilities exceeding the action
                      level was at pH < 7.40 and alkalinity <30  mg/L as
                      CaCO3. However, all utilities exceeding the cop-
                      per action  level in the very-low-alkalinity cate-
                      gory  also had pH <  7.0. Excluding those utilities
                      with pH < 7.0, all other utilities exceeding the cop-
                      per action level had alkalinity > 75 mg/L as CaCO3
                      and pH < 7.80. A closer examination of that data
                      (not shown) demonstrated that all utilities exceed-
                      ing the action level had alkalinity > 90 mg/L as
                      CaCO3 and many had pH > 7.0. Copper corrosion
                      by-product release is likely worse at higher alka-
                      linity as a result of the formation of soluble cupric-
                      bicarbonate complexes.18-19
Copper corrosion
   In-the 1992 AWWA sur-
vey, 361 utilities reported their
90th percentile copper levels
from the first sampling round.
The same strategies previously
described for lead release were
used for copper.
   Utilities not using phos-
phate-based inhibitors. For
utilities not adding inhibitors,
average 90th percentile cop-
per levels were highest in
waters with pH < 7.40 (Fig-
ure 6). A cpmbination of low
pH (<7.80) and high alkalin-
ity (>74 mg/L as CaCO3) pro-
duced the worst-case 90th
percentile copper levels. This
agrees with recent research
hypotheses that Cu(OH)2
solids control copper solubility
under some circumstances in
distribution systems.10-17-19
   In the absence of inhib-
itors, average 90th percentile
copper release  clearly tended
to decrease at higher pH (Fig-
ure 6). Percent reductions
ranged from I to 100 percent
with  only  one exception
(Table 5). In waters with alka-
      .--T'-   ':•' ~
     ?iS 3 Alkalinity '.':•££
     ':\.\'mg/L as CaCO3
     ^.T.- •;-~"!C'-.^--r-.r •--r'--^- :i
     •.. i-:.-,--i<30 • •-;•'., •
     .;.; '',:-/3O-74 .>•.'-•••-
       •"  '''•
            .                •
      •Significant at the 95 percent confidence interval
      tSignificant at the 85 percent confidence Interval
     , {Significant at the 90 percent confidence interval
              Percent reduction in copper release as a result of the indicated
              increase in pH for utilities not adding phosphate inhibitors
                      .  '., Reduction .In CppporReleaso^p«e^_,>;^;^^;^;
                       43*     • : "'•'•- "30   '  •.}^^--'^SJ^Sf'i^^i-rV
                       39       :"      si*      "..*"'^ff^:'^..'.
                                             51*
                                             53f
              Percent reduction in copper release for utilities adding phosphate
              inhibitors compared with those not adding phosphate inhibitors
                              • ...

                       <7.4      : 7.4O-7.80
                                             pH Cliango
                                                                  s-8.40
            Alkalinity
          mg/L as CaCOt
                              Reduction In Copper Release — percent
m®,
sC?fr--'«;-
^S#r:
Sw-ia>5'
s?-??-XW
•S?H?"S.';
                             56*:'   .-.-.-•: 13
                            . 11  ..:,;•:..•: -,-2
                             51. ;.,;:^V^.,.34t
                            V23 - •" =••' ''^^V" 4 =

V:.SNo data or IrisufBctent data for the category "."'"'    ~'
f^'^::-'-~i^tJ^'i.'^^^^^J^"'^-^'-- -;. --:
80  JOURNAL AWWA

-------
             Average 90th percentile lead in cold (<15°C) or warm
             (al5°C) waters for utilities not adding inhibitors
   Inhibitor effectiveness. As was observed for
lead corrosion, when the effects of pH and alkalinity
were ignored,  overall differences in copper release
attributable to inhibitor addition were not signifi-
cant. Thus, performance of inhibitors is clearly depen-
dent (at a minimum) on the pH and alkalinity of the
specific system.
   Although few systematic studies have evaluated
the effects of phosphate inhibitors on copper release,
most work completed to date demonstrates improve-
ments to copper corrosion at higher pH values (7.0-8.0)
and no effects at lower pH (< 7.0). For instance, Reiber
found that  corrosion rates were reduced by a factor
of 3-5 for utilities having waters at pH 7.5 dosed with
1-5 mg/L orthophosphate compared with utilities hav-
ing waters at the same pH that did not add inhibitors.20
In another  study, weight loss of copper pipes in the
presence of 1 or 5 mg/L orthophosphate was dramat-
ically reduced by orthophosphate addition at pH 8.O.21
However, at pH < 6.0, orthophosphate inhibitors did not
affect corrosion rates in the Reiber study. Similarly,
when dosed with 3.2-4.5 mg/L of zinc orthophos-
phate) Boston water at pH 6.7 exhibited no change in
copper release  when compared with the same water
without inhibitors.15
   As was the case for utili-   T—;    ——
ties that did not add in-   s :•=;';.
hibitors, copper release de-   '"••• •*$
creased  with increasing pH   !.';  '•;
for utilities that added in-   =»
hibitors (data  not shown).
Percent reductions as a result
of increasing pH category
ranged from 0 to 80 percent
with two exceptions. Interestingly, the apparent bene-
fits of inhibitor addition are mostly confined to the
lower pH ranges for the copper release data (Figure 7),
although the percentage of utilities exceeding the cop-
per action level was reduced by inhibitors in all pH and
alkalinity categories (Figure 7).
   At pH < 7.80 utilities using inhibitors generally
had lower  average 90th percentile copper release
                         when compared with utilities that did
                         not add inhibitors, with percentage
                         improvements as high as 56 percent
                         (Table 6). However, only two of these
                         trends were significant at greater than
                         90 percent confidence. Above pH 7.80,
                         addition of inhibitors had disparate
                         effects on copper release, depending on
                         the specific alkalinity category, with per-
                         centage increases attributable  to
                         inhibitors (negative numbers in table)
                         as high as 45 percent. The result at pH >
                         8.40 and alkalinity < 30 mg/L as CaCO3
                         was significant at the 95 percent confi-
                         dence interval, suggesting that inhibitors
                         can have adverse effects for copper cor-
                         rosion, at  least under, some circum-
                         stances. No other adverse effects were
                         significant at or above 85 percent confi-
              dence. On the basis of this analysis, it seems safe to
              state that the effects of inhibitor use above pH 7.80 are
              highly variable for copper corrosion.
                 Effects of temperature. Average 90th percentile
              copper as a function of pH and cold or warm water
              was quantified in both  the presence and absence of
              inhibitors. In general, copper levels decrease with
              increasing pH for both cold and warm waters whether
              or not the utilities added inhibitors (data not shown).
              There does not appear to  be a significant trend in
              terms of temperature (i.e., copper concentration is
              dependent on pH and other water quality parameters
              and not on whether the water is warm or cold). In
              addition, inhibitors are most effective in reducing
              average 90th percentile copper at pH below 7.80 and
              did not seem affected by temperature.
                 Effects of calcium. Similar to the trend with
              temperature, calcium content did not have much of
              an effect on average 90th percentile copper release.

              Comparison to solubility modeling theory
                 It is instructive to compare the trends described
              earlier to current frameworks of corrosion  control
              based on solubility modeling.
utilities with pH <7.40 and high-alkalinity
 waters had the highest copper
 concentrations.
                 Lead. The chemical equilibrium modeling pro-
               gram MINEQL+ (version 3.01)22 and associated
               database were used to make equilibrium predic-
               tions of lead solubility using various models. The
               basis for the prediction was to model the approxi-
               mate midpoint of the pH-alkalinity categories used
               previously in  segregating the practical utility data.
               That is, pH 7.0, 7.6, 8.1, and 8.8 were modeled at
                                                                                      JULY 1995  81

-------
              Average 90th percentile copper release for utilities not
              adding phosphate inhibitors
         --
alkalinities of 15, 50, 75, and 175 mg/L as CaCO3.
The different modeling approaches are based on
distinct assumptions as to solids that control lead sol-
ubility—in this work solids considered included
cerrusite (PbCO3), hydrocerrusite [Pb3(CO3)2(OH)2],
and both in combination.
   The best agreement between the utility data and
solubility models was observed when both cerrusite
and hydrocerrusite formation were considered. In
this model lead levels are predicted to decrease with
higher alkalinity at pH < 7.6 (Figure 8). This predic-
tion is consistent with the significant trends in the
utility data (Figure 1 and Table 1). The solubility mod-
els also predict nearly insignificant reductions in lead
solubility at alkalinity > 50 mg/L as CaCO3 at pH < 7.6,
results that are again consistent with significant trends
in the practical data (Table 1).
82  JOURNAL AWWA
              At pH £: 8.1 there was general lack
           of agreement between the solubility mod-
           els and the practical data. That is, the
           model invariably predicted higher lead
           solubility at higher alkalinity, whereas
           the practical data very clearly demon-
           strate the opposite trend (Figure 8 versus
           Figure  1 and Table 1).  Some of these
           trends in the practical data (at the low-
           .est alkalinity category) are significant at
           the 95 percent confidence level.
              In addition to the solids PbCO3 and
           Pb3(CO3)2(OH)2, hydroxypyromorphite
           [Pb5(PO4)3 OH] was included in the sol-
           ubility model for lead in the presence of
           phosphate-based inhibitors. A dose of
           0.95 mg PO42'-/L orthophosphate was
           used in the model. In general, lead re-
           lease  showed some qualitative agree-
           ment between solubility modeling and
           the practical data. For instance, in the
           lowest alkalinity category, formation of
           Pb5(PO4)3OH is predicted to greatly re-
           duce lead solubility regardless of pH, a
           trend observed in the practical data as
           discussed previously (Figure 2 and Table
           3). Predicted lead release approximately
           doubled when alkalinity increased from
           15 to 50 mg/L as CaCO3 in the  pres-
           ence of orthophosphate at all pH values
           tested (Figure 8).  Once again, this trend
           is qualitatively consistent with some of
           the practical observations defined  in
           Figure 2.
              Despite  the  noted agreements
           between the model and practical data,
           in general the solubility models tended to
           overpredict beneficial effects of inhibitors
           at higher alkalinities. That is, in the pres-
           ence of orthophosphate inhibitors, lead
           levels are predicted to be  reduced  in
           nearly all pH-alkalinity categories (Figure
           8). This was not observed in the utility
           data. Much of this disagreement is
because utilities were probably dosing polyphosphate
inhibitors or  dosing orthophosphate at concentra-
tions too low to observe significant effects. Never-
theless, some observable beneficial effects would be
expected at higher alkalinity if the model was perfectly
accurate. There are also many reasons why predictions
based on solubility models are not expected to agree
with the observed trends In the practical first-draw
lead data, as  discussed  by Schock.5 This work does
not provide any additional basis for speculation as to
which  of those effects is dominant in the practical
data collected for this work.
   Copper. There is good correlation between
Cu(OH)2 solubility models10-17-19 and the previously
described utility database (utilities without inhibitors).
As pH increases, copper solubility decreases for both
theory and practical data. In  addition, high copper

-------
 release is observed at high alkalinities
 when pH is below about 7.8-8.
    Comparisons between theory and prac-
 tice for copper release with inhibitor addi-
 tion are problematic because it has been
 difficult to determine which copper-phos-
 phate solids are controlling copper solu-
 bility. Recently, assuming the formation
,of Cu3(PO4)2 -2H2O solid, Schock18 pre-
 dicted the effects of orthophosphate on
 Cu(OH)2 solubility.  Below pH 8.0,
 orthophosphate addition reduces copper
 release according to these predictions,
 whereas above pH 8.0 copper release does
 not decrease when orthophosphate is
 added. Although solubility modeling can-
 not currently predict adverse effects attrib-
 utable to inhibitor addition (because there
 are no  complexation constants  for
 polyphosphates and because interference
 with the formation of protective oxide,
 hydroxide, or hydroxycarbonate scale
 cannot be predicted), the general trend of
 the solubility model provides insight into
 the disparate effects of phosphate
 inhibitors at pH above 7.8.
 Recommendations to utilities
   Although other water quality parameters (e.g.,
 dissolved oxygen, temperature, calcium, sulfate, and
 chloride) can influence lead and copper corrosion,
 this treatment indicates these factors are of secondary
 importance when compared with pH and alkalinity.
 That is, if any other water quality factor was "of pri-
 mary importance in lead or copper corrosion con-
  Improvement in percentage of utilities exceeding the
  copper action level as a result of adding phosphate
  inhibitors
                   hate inhibitors were usually
            beneficial in mitigating copper release;
            however, most benefits were at utilities
            with pH < 7.80.
 trol, no significant trends would have been observed
 when this data set was sorted on the basis of pH and
 alkalinity. Trends in by-product release, when viewed
 in the context of pH and alkalinity categories, were
 highly significant in some cases.
    On the basis of the graphical and statistical findings,
 a variety of simple measures have been identified that
 utilities may follow to mitigate high lead or copper
 release. Given the variation in corrosion phenomena
 from system to system, however, these predictions may
 not be universally applicable. Further research is needed
 to better define the types of inhibitors and chemical
 dosages that are effective in specific water quality types,
 and more information is needed regarding desirable
target pH values or alkalinities. Laboratory experiments
performed thus far are encouraging in that they gen-
erally support the trends identified in this work regard-
ing lead and copper release as a function of pH, alka-
linity, and phosphate inhibitor addition.23
   Lead. Most problems with lead release were in
low-alkalinity (< 30 mg/L as CaCO3) waters. In such
waters,  if pH is < 8.40, increasing the alkalinity to
30-74 mg/L as CaCO3 is likely to significantly reduce
lead release. Likewise, adding inhibitors to low-alka-
                      linity waters may also be
                      effective in reducing lead
                      release  regardless of pH.
                      Utilities with very-low-alka-
                      linity waters (<30 mg/L as
                      CaCO3)  that already  add
                      inhibitors are not advised to
                      increase alkalinity because
                      of potential aggravation of
                      lead by-product release.
                          Whether considering
                      alkalinity-pH adjustment or
addition of orthophosphate inhibitors, maintaining
pH is absolutely critical for corrosion control in poorly
buffered  waters.  For  example,   adding acidic
orthophosphate formulations (most commercial zinc
orthophosphate liquids are pH 1  or 2) can inadver-
tently lower the pH below the optimal level. Utili-
ties must ascertain and maintain the proper pH.
   For utilities in all other pH-alkalinity categories not
adding inhibitors, either increasing alkalinity, pH, or
both is predicted to reduce lead release and the likeli-
hood of exceeding the lead action level. Inhibitors were
generally not useful in reducing lead release in waters
with alkalinity > 30 mg/L as CaCO3, although, once
again, it would be useful to test  such conditions if
                                                                                       JULY 1995  83

-------
               Model-predicted lead solubilities in tha absence
               of orthophosphate (A) and in the presence of
               orthophosphate (B)
             Model assumes no orthophosphate (A) and 0.95 mg/L'
            .orthophosphate as PO4-3(B) '    ..  .•
inhibitor use was desired at a given utility. At pH >
7.40 and alkalinity of 30-74 mg/L as CaCO3, adding
inhibitors can apparently increase lead release. Many of
these adverse effects of inhibitors appeared attribut-
able to polyphosphates, so stopping polyphosphate dos-
ing if it is used for sequestering might be a viable opti-
mization strategy to be tested further at affected utilities.
   In high-alkalinity waters near or exceeding the
action level, pH adjustment often is not feasible
because of possible calcite precipitation. Orthophos-
phate or blended phosphate chemicals are probably
            the only viable alternatives in these
            waters, and they may have to be added
            at high dosages. Overall, the use of phos-
            phate inhibitors alone did not signifi-
            cantly correlate with low lead release.
            The qualitative effect observed depends,
            at a minimum, on the specific pH and
            alkalinity of the water of interest.
               Copper. The following strategies are
            suggested as effective for control of cop-
            per corrosion. At alkalinities < 90 mg/L
            as CaCO3, no utilities exceeded the cop-
            per action level unless pH was < 7.0.
            Thus, increasing pH to > 7.0 would seem
            to be  an effective and simple mitigation
            strategy for utilities in the low-pH and
            very-low-alkalinity regime. At alkalini-
            ties > 90 mg/L as CaCO3,  two strategies
            seem  viable. Because utilities with pH >
            7.80 in this category did not exceed the
            copper action level, increasing pH seems
            to be  promising.
               For utilities with high-alkalinity and
            high-calcium waters that cannot in-
            crease pH above 7.80 because of con-
            cerns regarding calcite precipitation,,
            any  increase in pH would likely
            improve matters substantially. Many
            small systems using groundwater sup-
            plies  are likely to have  such waters.
            Some trends  in the  data suggest that
            inhibitors were successful in mitigat-
            ing copper release in such systems;
            however, these trends were variable
            and not of high significance. Above pH
            7.80 there is some evidence indicating
            that inhibitors can adversely affect cop-
            per corrosion by-product release. In
            general, inhibitors caused highly vari-
            able effects above pH 7.80.
            Conclusions
               The  following conclusions were
            reached regarding lead.
'•!=.:"-': ' -^      • For utilities not dosing inhibitors
•i":" ":>v?.i,;-.;   at pH < 8.4> lead release was significantly
   :".:._:L   lower if alkalinity was 30-74 mg/L com-
            pared with alkalinity < 30 mg/L CaCO3.
               • For utilities dosing inhibitors at
            alkalinity < 30 mg/L as CaCO3, inhibitors
appeared to improve lead corrosion by-product release
compared with utilities not dosing inhibitors. At pH
< 7.4 these benefits were attributable to orthophos-
phate  and  not polyphosphates. No significant
improvement in lead release attributable to inhibitors
was  observed at other pH-alkalinity categories,
although this might be the result of improper inhibitor
dosing with respect to corrosion control.
   • At alkalinities 30-74 rng/L as CaCO3 and at pH
> 7.40, inhibitors seemed to adversely affect lead
release. Many of these effects were directly attrib-
84 JOURNAL AWWA

-------
uted to polyphosphates and not orthophosphates.
Increased lead release seemed to be significantly cor-
related with higher alkalinity (<30 to 30-74 mg/L as
CaCO3) when inhibitors were used.
   The following conclusions were reached regarding
copper.
   • Copper release is reduced by higher pH in the
presence and absence of inhibitors.
   • Problems with meeting the copper action level
are confined to two water quality regimes: (1) pH <
7.0 at alkalinity < 30 mg/L as CaCO3 and (2) pH < 7.8
at alkalinity > 90 mg/L as CaCO3.
   • The optimal inhibitor dosages  are probably
dependent on pH. Inhibitors appear to reduce aver-
age 90th percentile copper release at pH < 7.80. Above
pH 7.80, inhibitors had highly  variable and some
adverse effects on copper corrosion by-product release.

Acknowledgment
   This work was supported by a grant from the
AWWA Research Foundation (AWWARF). The
authors appreciate the constructive criticism provided
by their project  advisory committee members: Joel
Catlin, Leonard Graham,  Preston Luitweiler, and
Tiffany Iran. The insightful comments provided by
Michael Schock and Steve Reiber were  of immea-
surable value to  this work. Special thanks to all util-
ity personnel and project  participants in the 1992
AWWA Lead and Copper Survey—without their effort
this work would not have been possible.

References
  1. Final Report: Initial Monitoring Experiences of
     Large Water Utilities Under USEPA's Lead and
     Copper Rule. Prepared by Montgomery Watson,
     Economic & Engineering Services, Inc., and Peter
     Karalekas, Consulting Engineer.  AWWA, Den-
     ver, Colo. (1993).
  2. GIBRA, I.N. Probability and Statistical Inference for Sci-
     entists and Engineers. Prentice Hall, Englewood
     Cliffs, N.J. (1973).
  3. BENJAMIN, J.R. &• CORNELL, C.A. Probability, Sta-
     tistics, and Decision for Civil Engineers. McGraw Hill,
     New York (1970).
  4. DODRILL, D.M. Lead and Copper Corrosion Con-
     trol Based on Utility Experience. Master's thesis,
     Univ. of Colorado, Boulder (1995).
  5. SCHOCK, M.R. Understanding Corrosion Control
     Strategies for Lead. Jour. AWWA, 81:7:88 (1989).
  6. SCHOCK, M.R. & WAGNER, I. The  Corrosion and
     Solubility of Lead in  Drinking Water. Internal
     Corrosion of Water Distribution Systems,
     AWWARF/DVGW-Forschungsstelle Cooperative
     Res. Rept. (1985).
  7. Lead and Copper Rule Guidance Manual. Cor-
     rosion Control Treatment, vol.2. AWWA, Denver,
     Colo. (1992).
  8. SHEIHAM, I. & JACKSON, P.J. The Scientific Basis
     for Control of Lead in Drinking Water by Water
     Treatment. Jour. Inst. Water Engrs.  & Scientists,
     35:6:491 (1981).
 9. LEE, R.G.; BECKER, W.C.; & COLLINS, D.W. Lead
    at the Tap: Sources and Control. Jour. AWWA,
    81:7:52 (Aug. 1989).
10. SCHOCK, M.R. & LYTLE, D.R. The Importance of
    Stringent Control of DIG and pH in Laboratory
    Corrosion Studies: Theory and Practice. Proc.
    1994 WQTC, San Francisco, Calif. Nov. 6-10.
11. HOLM, T.R. ET AL. Polyphopshate Water Treat-
    ment Products: Their Effects on the Chemistry
    and Solubility of Lead in Potable Water Systems.
    Proc. 1989 WQTC, Philadelphia, Pa., Nov. 12-16.
12. BOFFARDI, B.P. Polyphosphate Debate. Jour.
    AWWA, 83:12:10 (Dec. 1991).
13. HOLM, T.R & SCHOCK, M.R. Potential Effects of
    Polyphosphate Products on Lead Solubility in
    Plumbing Systems. Jour. AWWA, 83:7:76 (July
    1991).
14. COLLING,  J.H. ET AL. Plumbosolvency Effects and
    Control in Hard Waters. Jour. Inst. Water &Envir.
    Mngmnt., 6:3:259 (1992).
15. KARALEKAS, P.C. JR.; RYAN, C.R.; & TAYLOR, F.B.
    Control of Lead, Copper, and Iron Pipe Corrosion
    in Boston. Jour. AWWA, 75:2:92 (Feb. 1983).
16. GREGORY, R. & JACKSON, P.J. Central Water Treat-
    ment to Reduce Lead Solubility.  Proc. 1984
    AWWA Annual Conf., Dallas, Texas.
17. SCHOCK, M.R.; LYTLE, D.A.; & CLEMENT, J.A. Mod-
    eling Issues of Copper Solubility in Drinking
    Water. Proc. 1994 Nad. Conf. on Environ. Engrg.,
    Boulder, Colo., July 11-13, 1994.
18. SCHOCK, M.R.; LYTLE, D.A.; & CLEMENT, J.A. Effect
    of pH, DIG, Orthophosphate, and Sulfate on
    Drinking Water Cuprosolvency. (Unpubl.)
19. MEYER, T.E. & EDWARDS, M. Effect of Alkalinity on
    Copper  Corrosion. Proc. 1994 Natl. Conf. on
    Environ. Engrg., Boulder, Colo., July 11-13,
    1994.
20. REIBER, S.H. Copper Plumbing Surfaces: An Elec-
    trochemical Study. Jour. AWWA, 81:7:114 (July
    1989).
21. BENJAMIN, M.M. ET AL. Corrosion of Copper Pipes
    in Chemistry of  Corrosion Inhibitors in Potable
    Water. AWWA, Denver, Colo. (1990).
22. MINEQL+, Version 3.01: A Chemical Equilib-
    rium Modeling System for Personal Computers.
    Environmental Research Software  (1994).
23. HIDMI, L.; GLADWELL, D.; & EDWARDS, M. Effect
    of Phosphates on Lead and Copper Corrosion.
    Poster session, 1994 WQTC, San Francisco, Calif.,
    Nov. 6-10.

                  About the authors: Donna M.
                  Dodrill is an engineer with Black &
                  Veatch, 8400 Ward Pkwy., Kansas
                  City, MO 64114. At the time this work
                  was conducted, Dodrill was a research
                  assistant at the University of Colorado
                  at Boulder. Marc Edwards is an assis-
                  tant professor in the Department of .
 Civil Engineering, Box 428, University of Colorado at Boul-
 der, Boulder, CO 80309-0428.
                                                                                       JULY 1995  85

-------
      REGULATORY GRIDLOCK
 The conventional Langelier index or
 Larson's ratio approach to controlling
 copper corrosion by-product release is
 demonstrably inaccurate.


 Marc Edwards,
 Michael R. Schock,
and Travis E. Meyer   '                                    Ithough bicarbonate was previously
                      "• '	 -     ^^^^b^r^i^B^is^^tiaj to pj-eyen^g copper c6rr8sibn
                                                               work has conclusively demonstrated
                                                                  bicarbonate under certain condi-
                                              ..   ....-,  _-;__-	—ities exceeding the US Environ-
                                              •--=;. mental Protection Agency (USEPA) 'action limits" for
                                              	Corrosion by-products must consider pH and bicar-
                                                                       bonate (alkalinity) adjust-
                                                                       .ments to optimize corro-
                                                                       sion,3 understanding the
                                                                       role of bicarbonate in cop-
                                                                       per corrosion is of critical
                                                                       importance. That is, if the
                                                                       basic effects of pH and bi-
                                                                       carbonate are not under-
                                                                       stood, utilities that follow

 JuuCl^X aoDlYiarfWa-ri ooforr-oxT-'-^-f-^'^^i-r^W^;-^ Wic-i^ii^s,;.-—-—£:-;T-i -\:L-: * -:- • •  "-.'   -'•      *^     -*        "•<
                                                                       aggravate copper corrosion
                                                                       problems.
                                                                          This work was de-
                                                          ter pH       signed to provide practical
                                                                       insight into the effects of
                                                                       bicarbonate on copper
*
      55®?^
    .-jixqsic	




--—tyaS-contrdl
•^^mts&zg&tt-
                                ^m
                                                   SVSKSS
                                                SS»
                                                         !^si


                              oi^m
                              ftS^Snfc
:COt
             attractive sti
 *.    	1 —™-r*>-«-»'»-^".di.»_gj ut\-&uac UCJULCULS jLjrujtn mgne
. are realized without adverse effects from higher alkalinity.
                                                                                 MARCH 1996 81

-------
                    °°PPer corros'on by-product release from pipe rigs



                                                                    ^fl^^ro^^S^c^Vi^^^l^'^bS'':'-''-'/:.!-"
82  JOURNAL AV^/WA

-------

     excessive concentrate
    corrosion by-prodiii
   corrosion and
   unambiguous!
   mechanism by
   acts. An in-depth analysis of mon-
   itoring data from large utilities,
   basic solubility experiments, and
   pipe-rig testing allowed rigorous
   evaluation of accepted corrosion
   mitigation strategies. A special
   focus of this investigation was the
   range of pH 7.0-8.5, over which a
   critical transition in bicarbonate
   effects was observed in previous
   research.2
  Materials and methods
     Tap water test solutions. Because.combined
  lirne-CO2 processes are the least expensive meth-
  ods of pH and alkalinity adjustment for many utili-
  ties, experiments were conducted on tap water .from
  Boulder, Colo., modified with lime and CO2.  Dur-
  ing these experiments, unaltered Boulder tap water
  had an average pH of 7.2 and alkalinity of 15 nWL
  as CaCOj (Table 1). CO2 was always bubbled to the
  solution before the addition of lime to prevent pre-
  cipitation of CaCOj.
              ihoug'h bicarbonate was previously
                nsidered essential to preventing
             copper corrosion problems, recent work
             has conclusively demonstrated adverse
             effects from bicarbonate under certain
             conditions.
                                                 stituents in addition to pH and alkalinity. Thus, these
                                                 tests were conducted with "synthetic waters" consti-
                                                 tuted in the laboratory. Copper solubility was exam-
                                                 ined in 1-L solutions at a predetermined alkalinity
                                                 and pH in the presence of a 0.0 1 -M NaClO4 swamp-'
                                                 mg electrolyte. A 4 x 4 experimental matrix was eval-
                                                 uated, including alkalinities of 10, 50, 100, and 250
                                                 mg/L as CaCO3 at levels of pH including 7.0, 7.5, 8.0
                                                 and 8.5. Experiments examining bicarbonate and
                                                 chloride interactions also utilized synthetic solutions
                                                 In these experiments alkalinity and pH were adjusted
                                                                     to the target value using an
                                                                     appropriate combination of
                                                                     C02 and NaHC03. There-
                                                                     after, if desired, 1 mM of
                                                                     NaCl was added. Finally,
                                                                     solution ionic strength was
                                                                     adjusted to 0.01 M by addi-
                                                                     tion of NaClO4 (an inert
                                                                     electrolyte).
                                                                       Soluble and total cop-
                                                                     per. Soluble copper was
                                                                     measured at 20°C after
                                                                     membrane filtration (0.2-
                                                                     um-pore-size cellulose ace-
 acid solutions const mtom    ldt
 bubbled with CO, to pH 4 O^re dosed 3?hnw
 controller. This
                                                 T                         - Approximately 25
                                                  °r S3mple Were filtered for ^-analysis, and no
essary) without ah      kmv buhhhn  rn     1   acidification to pH 2.0. for inductively coupled
   Synthetic test solutions. A few experiments
required precise control over other inorganic con-
                                                 TOPCON Co.. Paramus. N.J.
                                                 fHach Co., Lovcland. Colo.
                                                                              MARCH 1996  83

-------
                  _  .     •      ••3:->&.tw&msS!££*s&%-. ^•X'
                  Typical base water quality durir-" ~;--" - "" *" at(=
                  period
                                                      short-
                                                        by-
                             .                            *
                 Electrochemical erosion rate measurement
                                 g Procedure. Specifics of stan-
                                 tests and the apparatus have
                                   Y^ la^ the key parameter
                                 eiiaf3l testing, is the total cor-
                                  by the exposed macroscopic
                              corroding metal; i.e., the instan-
                              ate of the sample. 7^ was deter-
                   ^S^an -electroanalysis system.f This article
                          Sfe-results from duplicate analysis
                        ses/determinations were quantitatively
                         :    uA/on2).6
                                   tests, the general goal was
                             on-copper pipe representative of
                                   of conventional exposure.
                                 • T^asp^^^Ji^er years o  conventional exposure.
                                        ~~" '"' ~'"~ " "•'  "CoPPe..r was exp'osed to the solution of interest for
   using the colorimetric test. For all waters in which the   TP*-P ***- f^n a c°frosion-a^elerating potential
   colorimetric test was used-possible maS irSriS    £~< co?  ~    ™Y  W3S appIiecL At the end of
   samples.
      Scanning Electron Microscope (SEM) analy-
   sis. Samples were prepared for SEM analysis by cut-
   ting the pipes lengthwise. The pipe surfaces were not
   coated because the conductivity of the copper itself
   was sufficient, thereby guar-
   anteeing  that  surfaces
   would be unaffected by a
   coating procedure. The sam-
   ples were examined in a
   vacuum at 30 kV  on an
   electron microscope.*
     Pipe-rig tests. Copper
  corrosion by-product release
  was examined in standard-
             copper surfaces, as reported in the previous s°ec-
             tion. The flow rate (0.5 gpm [0.031 L/s] ± 0.02 °pm
             [0-0013 L/s]), pH (± 0.03 pH units), and alkalinity
             (± 3 mg/L CaCO3) were rigidly controlled during
             experiments.
    concentrations are a surrogate
     copper carbonate complexation
capacity in many waters.
  ized "pipe rigs." Type M copper pipe measuring 24
  m (609 mm) x Win: (19 mm) was purchased from
  a local hardware store and was washed in 0.1 M
  NaOH for 2 min to remove organic deposits" There-
  after, pipes were rinsed five times with deionized
  water before immediate exposure to target solutions
  Pipes were filled with the target solutions three times
  a week (Monday, Wednesday, Friday) and maintained
  in a horizontal position at all other times A rubber
  stopper tightly sealed each end.
    Filtered and unfiltered samples were taken weekly
 to determine soluble and total copper concentrations
 after 72 h of exposure (Friday to Monday stagna-
 tion). After three months of exposure,  representa-
 tive  coupons were cut from  each pipe  for electro-
 chemical and SEM examination. Prior to the
 electrochemical analysis, pipe coupons were exposed
 to the target solutions at a flow rate of 0.2 gpm (0.013
 L/s) for,30 min. Each water quality was tested using
 replicate pipe rigs. The only difference between repli-
 cates was that exposure to  test solutions was initi-
              Monitoring experience of large utilities. Data
           collected during a national survey of 435 large utili-
           ties were used in this investigation.7 The analysis of
           that database proceeded as follows.
              • Because inhibitor additions might hopelessly
           complicate the role of pH and alkalinity in copper
           corrosion, any utilities using corrosion control other
           than pH, alkalinity adjustment, or both were deleted
           from the database.
              • Any utility not reporting essential information
           including pH, alkalinity, or 90th percentile copper
           release was also deleted.
              The 151 remaining utilities were sorted into pH
           and alkalinity categories for analysis. pH and alka-
           linity values in the database, typically six-month aver-
           ages for water released to the distribution syst<
                           possible variations within the i
                         L or between different sources.
84 JOURNAL AWWA

-------
       Bicarbonate, alkalinity,
    and dissolved inorganic
    carbon. Over the pH range
    of most natural waters (pH
    5.3-8.7) and at alkalinities
    >10 mg/L as CaCO3, bicar-
    bonate concentrations are
    directly proportional to alka-
   linity within ± 3 percent error
    (i.e., 50 mg/L alkalinity as
   CaCO3 = 1 mAf HCO3-). Ac-
   cordingly, this  work  will
   apply the terms alkalinity and
   bicarbonate interchangeably
   according to utility conven-
   tion. Approximate units of
   dissolved inorganic carbon
   (DIG) are provided for reader
   convenience using die equa-
   tion DIG (mg C/L) == 0.24
   [alkalinity (mg/L as CaCO3)].
  At pH values between 7.5
  and 8.7 this approximation is
  valid to within ± 5 percent,
  but the approximation un-
  derpredicts actual DIC by 17
  percent at pH 7.0.

  Experimental results
    The experimental results
  are divided into three sec-
  tions, with  the first two
  addressing the effect of bi-
  carbonate on copper corro-
 sion by-product release and
 solubility. Thereafter, the
 combined effects,.of bicar-
 bonate and  chloride are
 examined in detail.
    Effect of bicarbonate
 on by-product release and
 solubility. Bicarbonate and
 by-product release from pipe
 rigs. Copper corrosion by-
 product release was. more
 dependent on water .quality
than on time of exposure
 (Figure 1). That is, although
the exposure of replicate
                                (•s^ssssr-""-"1-----*.---'-
 endar date rather than versus time of exposure Th s
 was particularly evident on July 5 1 993 when each
 pipe exhibited a dramatic reduction in coppIrSeas?
 If differences in exposure time caused th? observed
 decrease only the pipe exposed for the extra 12-day
 time penod would have exhibited this drop  IntS
 estmgly, the decrease in by-product release  corre
 sponded to a period of highsLrmwater runoS *
   Copper release was fairly constant at each condition
over the last four weeks of'exposure^reTS
                                                                                '
                                                          W3S Very g°°d' avera§e c°PPer corr°-
                                                                     * "*** funai°n (Figure
                                             sion                  C°™ati°n)- The «»«.
                                             data SDH 7 ?nri? ^     ^ ^ f°r the averaged
                                             nerrenr 2 rh     resPecuvely. Because more than 95
                                                         C°P?f r m the samples Passed
                                                              /' ^ "^ W3
                                                            3nd not Particulate-
                                                                       *****
                                                                             MARCH 1996 85

-------
                  Measured soluble copper after solubility
                  solubility
                                                   experiments and predicted soluble copper based on Cu(OH)
                                                                           ('•JL*r 

-------
    conducted. In these experi-
    ments soluble cupric ion was
    added to solutions at fixed
    pH, ionic strength, and alka-
    linity. After allowing 20 min
    for precipitation to occur, sol-
    uble copper was then mea-
    sured  in filtered samples.
    Consistent with previous
   trends, soluble copper was
   approximately a linear func-
   tion of alkalinity  (Figure 4).
   The experimentally observed
   slope at pH 7.0 and 7.5 was
   about 0.0116 mg soluble cop-
   per/mg alkalinity, whereas
   the slope at pH 8.0 and 8.5
   was about 0.004 mg soluble           . .
   copper/mg alkalinity. Thus, at pH 7.0-7.5 -soluble
   copper was about 2.9 times more sensitive to the pres-
   ence of alkalinity than at pH 8.0-8.5. This was fairly
   consistent with the 2.0-2.7-fold slope differential in
   the by-product release data (Figure 2) and utility mon-
   itoring data (Figure 3).
     Given the previous experimental findings, various
                            Model constants from MINEQL+ database12'1*
                                   complexes at higher pH, because the Cu(OH)2
                                   model predicts a weak -dependency on alkalinity
                                   whereas the solubility data (Figure 4), by-product
                                   release experiments (Figure 2), and utility data
                                   (Figure 3) indicate a much stronger dependence
                                   on alkalinity.
                                      By comparison, predictions of soluble copper
                                                                        pH and alkalinity ranges,
                                                                        whereas that based  on
                                                                        bronchantite  overpre-
                                                                        dicted copper solubility by
                                                                        a factor of three to four.
                                                                        Moreover, malachite sol-
                                                                        ubility predicts less solu-
                                                                        ble copper at higher alka-
                                                                        linities if pH  was >7.4, a
                                                                       tendency that was obvi-
                                                                       ously inconsistent with
     the pH range of afaout 7.0-8.0,
adverse effects from bicarbonate are
significantly reduced by even small
increases in pH.
 evaluation, though an internally consistent critique
 has been presented elsewhere.10 Four cupric solids
 known to form on copper pipe were examined for
 consistency with experimental data including tenorite
 [CuO], cupric hydroxide [Cu(OH)2],  malachite
 [Cu2(OH)2C03], and bfonchantite [Cu4OH6(SO4)]."
 The model prediction'was based on the assumption
 that soluble copper was controlled by equilibrium with
 one of these solid phases (Table 2).
    Predictions of soluble copper based on cupric
 hydroxide solubility were both qualitatively and
 quantitatively consistent with the previous data
 (Figure 4). Quantitative predictions always agreed
 to within  1.5 mg/L copper, a very reasonable result
 given uncertainties in the required equilibrium
 constants. The agreement between the Cu(OH)2
 model and observed solubility was excellent at pH
 7.0, with weaker predictive capability at the higher
pH values. The discrepancy between the model
prediction and the experimental data appears  to
be due to  an underestimate in copper carbonate
                                  tance of cupric hydroxide has recentfy been put
                                  forward by Schock et al1" using data from USEPA
                                  pipe rigs.
                                     Effect of alkalinity on corrosion rates and
                                  scale morphology.  Corrosion rates of copper in
                                  pipe-rig tests. At the end of the by-product release
                                  experiments, coupons were cut from the pipes, and
                                  instantaneous corrosion rates were determined elec-
                                  trochemically under flow conditions. Although the
                                  trends in corrosion rate are not a linear function of
                                  alkalinity, the copper corrosion rates did increase
                                  consistently with higher alkalinity at both pH 7.0
                                  and 8.0 (Figure  5).
                                    The corrosion rate, /„„., represents total copper
                                  oxidized at the pipe surface per unit of time. This
                                  oxidized copper may either be incorporated into solids
                                  on the pipe surface, forming a scale, or be released to
                                 solution as a corrosion by-product. Assuming that
                                 corrosion proceeds through a one-e- transfer, the fol-
                                 lowing conversion was derived relating corrosion rate
                                 to copper mass corroded:
                                                                                 MARCH 1996  87

-------
                  Average copper release and measured corrosion current in naturally
                  aged samples showing the same trend
                 Corrosion currents in the presence and absence of 1 mM chloride
                                -«3§$t5;=£-• ^yprSSl.-1 r.~-Z-'^^.:3.-"!*£-'S->-'=p?xy:v? •--*• -w-w ••,•*. » " •


                                [^•SlSrr^::"?'^^^^"^'^^^"*"-^"'*^-^"^
    IpA     1A  .
    cm2  .  106 jzA X coupon
20 cm2   1 C/s
       x	
le-
              _  n
              ~ °'018
                           (3)
                                        mmolCii
                                          day
    Two major factors might contribute to differ-
 ences m corrosion-rate estimates based on copper
 mass lost to solution (i.e., the by-product release
 data) versus the direct electrochemical measure-
 ment. The first factor was that by-product release
 measurements were made under stagnant con^-
 tions, whereas corrosion rates were determined
 under flow conditions. This difference would tend
 to increase the corrosion rate' during flow versus
 stagnation because of improved mass transport The
 second factor was that much of the oxidized copper
 could be incorporated into a growing scale layer
 and not released to solution. Any copper incorpo-
 rated into scale will not be quantified by measure-
 ment of released corrosion by-products, thereby
                         reducing the estimate of cor-
                         rosion rate based on by-prod-
                         uct measurement.
                            Despite these difficulties,
                         it was instructive to examine
                         the actual measured corrosion
                         rate and compare it with that
                        .estimated on the basis of by-
                         product release to stagnant
                      |   waters. Assuming that corro-
                      :   sion rate was not influenced
                        by differences in flow, calcu-
                        lations indicate that about 3
                        percent of copper that was
                        corroded was actually released
                        to solution; in this case, about
                        97 percent of the corroded
                        copper must have been incor-
                        porated into scale formed on
                        the pipe surface. At another
                        extreme, if it was assumed
                        that all the corroded copper
                        was actually released to solu-
                        tion as a corrosion by-prod-
                        uct and none was incorpo-
                        rated into scale, the corrosion
                        rates under  conditions of
                        stagnation must have  been
                       more than 33 times  slower
                       during  stagnation than under
                       conditions of. flow. Reality
                       probably reflects a compro-
                       mise between these two ex-
                       treme assumptions.
                          Dual ion interactions:
                       chloride and bicarbonate.
                       Although the previous exper-
                       iments  clearly isolated the
                       effects of pH and bicarbonate
  on copper corrosion rates, other water quality con-
  stituents such as natural organic matter, chloride
  and sulfate can also strongly influence copper  cor-
  rosion.s.ii Because both chloride and bicarbonate
  appear to be critical to controlling copper corrosion
  rates, their combined effect was examined.
    The accelerated aging technique was used to
  examine copper corrosion in well-defined synthetic
  waters at pH 7.0 and 7.3 and constant ionic strength
  At a given level of alkalinity, solutions were consti-
  tuted with or without the addition of 1 mM chloride
  at all alkalinity values (Figure 6). Chloride addition
  decreased corrosion rates between 84 and 95 per-
  cent in aH cases. Thus, from this perspective.at least
 the presence of chloride can counter adverse effects
 of bicarbonate.
    A comparison of scales formed in the presence
 and absence of chloride verified the dominant effect
. of chloride on scale appearance. Chloride supported
 formation of a reddish scale that  was nearly identical
 to that observed at pH 7.0 in the presence of chloride
 only. 6 in contrast, scales formed in the absence of
** JOURNAL

-------
 chloride were quite smooth in appearance under the
 SEM, giving the surface-a nonporous appearance.
 Thus, chloride induced very significant visual and
 electrochemical changes to copper corrosion.

 Synthesis: role of bicarbonate
 in copper corrosion
    This section critically evaluates key results of this
 investigation in light of previous research. The effects
 of bicarbonate on corrosion by-products and corrosion
 rates are discussed separately. The section concludes
 with a short discussion of copper carbonate com-
 plexation versus dissolved CO2 as a source of copper
 corrosion problems.
    Effect of bicarbonate on copper corrosion by-
 product release. Evidence from the pipe-rig exper-
 iments and the monitoring experience of large utili-
 ties is clear and unambiguous: soluble copper
 corrosion by-product release increases as a linear
function of alkalinity. Because these trends were con-
                                                      using conventional units with copper in mg/L W.as- -~
                                                      pH, and alkalinity (= HCO3-) in mg/L as CaCO,^" *"*T
                                                      constant pH of 7.0 and 8.0, this equation simplifies to  *L -
                                                      a linear form:           ' .           • •
                                                     Soluble Cu (mg/L) = 0.83 + 0.015 [alk] at pH 7.0
                                                                                                 (7)-
                     ut considering inhibitor dosing
                                                       Soluble Cu (mg/L)Wo.58;*p;boi3 {aft] atpHK6$?(8)'

                                                        Qualitatively, this predicted linear relationship
                                                     between soluble copper and alkalinity (at constant
                                                     pH) is in excellent agreement with the experimental
                                                     results. .The quantitative agreement is ab'd quif<|pod,:
                                                     especially at lower pH values. Indeed, a comparison
                                                     of Eqs 1 and 7 demonstrates remarkable agreement
                                                     between the model at pH 7.0'and-theactaal by-prod--
                                                     uct release data, consistent with the excellent pre-
                                                     dictive capability of the model at the lower pH for
                                                     the solubility data (Figure 4). A comparison of Eqs 2
                                                                           and 8 reiterates the fact
                                                                         •  that'the model underpre-
                                                                           dicts sensitivity of copper
                    :emperature, the best advice
                is to raise pH.
                                                                           The slope of the equa-
                                                                        tion, equal to the increase
                                                                        in.soluble copper (mg/L)
                                                                        per incremental increase
                                                                        in alkalinity  (mg/L as





Cu(OH)2 (solid) were quantitatively and qualitatively
consistent with the available data.
   In this model, the predominant soluble copper
  species over the pH range 7.0-8.5 include
  cu(OH)2(aq), CuC03,aq), and CuHCO3+. Thus, the
  concentration of soluble Cu can be calculated as-
                                              (4)
                                             (5)
                                     -]2 K  -
 in which Kj = [Cu(OE)2
                                  ], jr  = [cu+2]
          w = [OH-] pff+], and Cu is in mol/L
-    Because Klf K^ K3, 1^, and ^ are constant, cal-
 culating results at 25°C without activity corrections
 yields:

 Soluble Cu (mg/L) = io(i3.4-2PHj + 0.58 + io(5.i-pH)[alkj
                  + 10 (U.4-2pH)
                                                                               .
                                                pH by 0.5 units (i.e., from 8.5. to 8.0 or from 8;0..to .
                                                7.5) increases the slope, by. about, a factor of three, •
                                                indicating a threefold increase in copper carbonate
                                                complexes at the lower pH.: •.'•; ....' •-;.;.- ..•:!-.\-:; :-..-;.-•.-,
                                                   If enthalpy values for the Cu(OH)2 solid in'the ••--'•
                                                MINEQL+ database are considered (Figure.7), cop-,
                                                per-carbonate cpmplexatiori is predicted to be a strong
                                                function of temperature, Tyith each  10°C increase
                                                halving the slope. In other-words, because Cu(OH)2- - -
                                                dissolution is exothermic,,a given concentration of"
                                                bicarbonate (alkalinity) is predicted to complex about
                                                twice as much copper at  5?C as it would aM5°C;-
                                                rnterestingly, this predicted temperature-dependency
                                                can explain recent findings that copper corrosion by--
                                               product release is lower in household hot-water taps
                                               than it is in cold-water taps.? This temperature depen^
                                               dency also.'calls into question recent proposals:to-
                                               require utilities to monitor: only in the summer
                                               because the presumption of higher copper release at-   .
                                               higher temperature is at odds, with predictions based
                                               on solubility and is not supported by  utility moni-
                                               toring data.17                         .    .  . :
                                                 The Cu (OH)2 model is also consistent with empir-
                                               ical linear relationships developed at KIWA to predict
                                               the maximum concentration of copper released to
                                               water during stagnation:18-19
                                                                                    MARCH 1996  89

-------
                                    rfubiltty to alkalinity and temperature
Old pipe:
 New pipe:
                     (mg/L) = 0.01 10 [Alkalinity]
                            - 1 -37 pH + 0.02 1 [S04-2]
                       • '    + 10.2
                                                                            onstrated that bicarbonate has
                                                                            a dual nature dependent on
                                                                            the solution pH. Above about
                                                                            pH 8.1, the presence of bicar-
                                                                            bonate tends to passivate cop-
                                                                            per surfaces and decrease cor-
                                                                            rosion rates.6 However, below
                                                                            about pH 8.1, bicarbonate is
                                                                           increasingly aggressive at
                                                                           higher alkalinities (>100
                                                                           mg/L as  CaCO3). A closer
                                                                           look at the  literature reveals
                                                                           an interesting trend; i.e., the
                                                                           solution pH is greater than 7.7
                                                                           in nearly all case's in'which
                                                                           bicarbonate is reported to
                                                                           have beneficial effects.20-22
                                                                             In cases in which the cur-
                                                                           rent findings seem to be con-
                                                      tradictory to the conclusions of previous research a
                                                      closer look reveals just the opposite. For instance
                                                      bicarbonate was studied at pH 6 and pH 8 by Matts-
                                                      son and Fredriksson.23 Though the authors conclude
                                                      that bicarbonate forms passivating scales, the data
                                                      showed that at pH 6 anodic corrosion currents did
                                                      increase with bicarbonate concentration and a very
                                                      large increase was observed at the highest alkalinity
                                                      tested (Figure 8). At pH 8, however, the currents
                                                     and  adverse effects  of bicarbonate were *reatlv
                                                     reduced.23                                    J .
                                                        Cohen and Myers24 examined treatment alterna-
                                                     tives for an outbreak of copper pitting in Fort
                                                     Shawnee, Ohio. Prior to treatment, the water had an
                                                     alkalinity of 296 mg/L as CaCO3 and an average pH




  concentrations of particulate-    .-v - -   .
  copper not removed by a



  tion and actual experience.
  Nevertheless, it is encourag-
•  ing that trends in the experi-
  mental data and the Cu(OH)2
  predictive model are consis-
  tent with those observed in
  aggregated monitoring data
  of large utilities.
     Effect of bicarbonate on
  copper corrosion rates and
  scale morphology. At first
  glance, findings that bicar-
 bonate has an adverse effect
 on corrosion rates appear con-
 trary to results of previous
 researchers. The authors' pre-
 vious work, however,6 dem-
                       (mg/L) = 0.0104 [Alkalinity]
                              -2.26 pH + 18.1

   in which alkalinity is in mg/L as CaCO3 and SO4-2 is
   in mg/L (units have been changed for consistency
   with US utility convention). The empirically deter-
   mined slope of 0.0104-0.0110 mg Cu/mg alkalinity
   is within the range of those determined in this work
   corresponding approximately to the slope predicted
   at pH 7.3 at 15°C by the Cu(OH)2 model (Figure 7)
     Predictions based on the Cu(OH)2 model, even
     ill l(~\T-f\  oy»l»/4*-> *<) A ««._*._ _1 __ *_ 1  -•*• .  .
90 JOURNAL AWWA

-------
   disappeared. Results from the current work indicate   canacitv That fc thV £m£i>£«X " * ^SKS^S^y-C^f 'S
   ^SiSSSHS"   pi&s^^?P^^iilf
     f°f ^d.^°£er?^^fat.ed 18 ^te1? I11 New_, -whichjna^beiewritteii as'^i -" ;
     dand and: althoue-h thevakn mr,,*,,*,* th** x;~,.' :•.-. . . .. ....^ *  •. „. ," -.n~:;.'S-:  :.-.: .:••
  most aggressive waters had 411-"and 178-mg/L bicar-
  bonate concenttatipns'krpH 7.3rarid 7.1, respec-
  tively.2^ Both of these waters are within the aggres-
  sive region for bicarbonate' denned in recent work
  done by the auttb'fs of tffis'article. In addition, the
  beneficial effects attributed to bicarbonate in Moss •
  and Potter's work was partly based on anodic polar-
  ization experiments/and the.cathodic oxygen reduc-
  tion reaction was ignored; Thus, these investigators
  would not have observed an increase in the cathodic
  oxygen reduction reaction rate as was observed in'
  previous work.2'5
                                           v^^^yj^i^^^-i^gl^^^^r^Xp^lii
                                            CCCu (mg/L) = [C0j (moW.)] .x {3j09 x 10? + IO^H [5.79^ upfi "

                                            usini
                                                                                    K&i.i
                                                                                    •' Ce,-:. 'A?/-

         e described benefits of aeration are
        ixpected to apply to many high-alkalinity,
       low-pH waters.
 copper-carbonate-'cbmplexes become increasingly
 important at higher alkalinity;'tending to reduce the
 free copper (Cu-*-2) actiyity'in solution. At a fixed pH
 and dissolved oxygen' concentration, the driving force
 for the anodic corrosion reaction increases as free
 copper activity decreases:

         Cu 4- &02 + H2O =* 2OH- + Cu+2

   Thus, the presence of coinplexihg-carbonate
speaes could reduce iihe free copper concentration
thereby increasing the corrosion rate. Alternatively,
PTi S)f^trf*rcf* ii-»fIii£»*-»^^ nf V;«._i	^.__    *•*    • _
                                                             the maximum concentra-
                                                             tion of carbonate complexes
                                                             that would form or, in other
                                                             words, the' tendency of a
                                                             given solution to form car-
                                                             bonate complexes. Future
                                                             research:, should revisit
                                                             hypothesized direct roles for
                                                             CO2 in copper corrosion,
                                                             because, correlations be-
                                           tween copper corrosion problems and CO2 may be
                                           an artifact of copper-carbonate complex formation.

                                           ImplicationsfoVutHities^O;';!:  ','.^',    "
                                             Important implications of this research fall mto
                                           two broad categories: corrosion control strategies to
                                           avoid and recommended strategies. Given the regu-
                                           latory framework of the Lead and Copper Rule, these
                                          .strategies apply .spedficaUy t6:mitigation of copper
                                           (i.e., not lead) corrosion by-product release.
                                            Corrosion control strategies to avoid. Utili-
                                          ^^^copperco^^^n^teses




rates would be forniitrti:iP:^°    -           friendly corrosion optimization software sold
      ^^^    J-V/A.I.UJ.LWU..S.         •      ,       triFono'Ti ATATXA/A 27 *T*J-»*    1 J

soIvTcarb^dToxSe6°(CO^SZ^T^n ^   ^?*tbeaefa"feMS^outbSmSa^unS
been correlated to ma^y SS'o?^^^   SSTl C°nditir th^ Provide the ^vice most
P^^^u^gp^g^^^°^0^0^   ^lyo exacerbate copper corrosion by-product

                                          ^example.the^gelierindex^approachis
                                                                     MARCH 3996 91

-------
                At a given pH, calcite saturation indexes tend to increase at higher
                alkalinity (data shown for pH 7.2-7.4)
  calcite supersaturation are correlated with reduced
  corrosivity:28
      Calcite saturation index (SI) = [Ca
  in which [Ca+2] = calcium concentration, [CO3-2] =
  carbonate concentration, and K = calcite solubility
  product.
    To connect the SI expression with the traditional
  LI approaches, a positive II corresponds to SI values
  above 1.0, whereas a negative II corresponds to SI
  values below 1.0. The SI and II both increase as Ca+2
  and CO3~2 concentrations increase. Thus, in the sim-
  plest test of this model's validity, at constant pH cop-
  per corrosion problems should decrease as Ca"1"2 and
  CO3~2 increase.
    The work described here directly repudiates this
 hypothesis. In experiments in which both Ca+2 and
 .CO3-2 were increased at pH 7.0 or 8.0 (Figure 2), lead-
 ing to a higher SI and presumably a better condition for
 copper according to Langelier theory, problems with '
 copper corrosion by-product release worsened. Simi-
 larly, as evidenced by the costly experience of large
 utilities under the Lead and Copper Rule (Figure 3), in
 a given pH category copper corrosion by-product release
 increases at higher alkalinity. Because higher-alkalin-
 ity waters have much higher calcite SI values at a given
 pH (Figure 9), the II predicts reduced problems with
 copper by-product release at higher alkalinity. In sum,
 both findings are completely inconsistent with the Lan-
 gelier approach but are consistent with the Cu(OH)2
 model developed by Schock et al and in this work.10-29
   With respect to Larson's ratio (LR) :30

                LR = [Cl-]/[HCO3-]

 the higher the LR ratio, the higher the corrosivity of
 the water supply. With respect to  copper corrosion,
 this work and others have demonstrated that chloride
 has long-term beneficial effects (on copper corrosion
 rates), whereas bicarbonate has adverse  effects on
 both by-product release and corrosion rates.6 Thus, if
                        anything, the exact opposite
                        of Larson's prediction is likely
                        to be valid for copper.
                          The preceding discussion
                        should not be interpreted as a
                        criticism of the index origi-
                        nators. They, more than those
                        who followed, understood
                        the limitations.of their work
                       For instance, the Larson's
                       index was not derived for
                       copper, it was derived for steel
                       and cast iron. Likewise, with
                       respect to the. Langelier the-
                       ory, it has long been known
                       that calcite does not actually
                       precipitate on pipes in most
                       distribution systems; thus, the
                       theory might'actually work
  quite well for the few cases in which precipitative
  calcite scales do form.
    The question arises, however, as to why the Lan-
  gelier model has gained such widespread and appar-
  ently misguided acceptance. The  answer, in the
  authors' opinion, is that it provides an easy-to-use
  guide to solving a complex problem. Moreover,
  because the II is typically applied to answer the ques-
  tion "How high should we raise the pH?," it will work
  quite well under some circumstances. Because higher
 pH typically reduces copper corrosion by-product
 release, the index occasionally works by accident.
 Thus, when used in the context of calculating the
 maximum pH allowable in a given system to avoid
 problems with calcite precipitation, the LI still has
 considerable value.
    Recommended corrosion control strategies.
 Without considering inhibitor dosing .(viable but
 beyond the scope of this work) or temperature (which
 is uncontrollable),  the best advice is.;to.raise pELIn
 particular, over the pH range of about 7JO-8.0, adverse -
 effects from bicarbonate are significantly reduced by
 even small (+ 0.2 pH units) increases in pH.
    Interestingly,  the authors' analysis indicates that
 the method of increasing pH is important. Consider
 three different options including NaOH  (caustic),
 CaOH2 (lime), or aeration-CO2 stripping. For the
 caustic and the lime up to a pH of about 8.5, the OH~
 added will increase alkalinity:

                OH- + CO2 -* HCO3-

   In contrast, raising pH by stripping CO2 increases
pH but does not alter alkalinity, because the reaction
producing OH~ consumes bicarbonate (which is sub-
sequently regenerated):

          HCO3- -* CO2 (stripped) -^ +  OSr

   The significance of this effect on predicted copper
solubility depends on the initial alkalinity and pH of
the water. Consider a hypothetical water with initial
92 JOURNAL AWWA

-------
  alkalinity of 250  mg/L as
  CaCO3, initial calcium hard-
  ness of 100 mg/L as CaCO3,
  and temperature of 25°C. If
  the final pH is  raised to 7.2
  with lime or caustic, soluble
  copper is predicted to be 3.8
  and 3.4 mg/L if the initial pH
  is 6.5 or 6.7, respectively (Fig-
  ure 10). If pH is raised to 7.2
  using aeration, soluble copper
  is predicted to be only 2.8
  mg/L regardless of initial pH
  (because alkalinity does not
  change upon aeration). This
  represents  a  17-26 percent
  enhancement in copper cor-
  rosion control (at a given final -
  pH) using aeration compared
  with lime or caustic.
    More important, however, is that pH may be
  increased to higher values without precipitating cal-
  cite if aeration is used. In the waters modeled in Fig-
  ure 10 and  for lime addition, calcite is supersatu-
  rated at only pH 7.2 or 7.3, depending on the initial
 pH. If pH is raised using aeration, calcite is not super-
 saturated until pH 7.6. Thus, aeration provides flex-
 ibility to raise pH to higher values without precipi-
 tating calcite.
    Quantitatively, increasing pH to the point of cal-
 cite supersaturation  (pH 7.6) using aeration is pre-
 dicted to yield soluble copper of 1.3 mg/L (Figure
 10). If pH is also  raised to the point of calcite super.-
 saturation using lime, soluble copper is predicted "to
 be 3.9 and 2.8 mg/L at an initial pH of 6.5 and  6.7,
 respectively. Therefore, aeration offers a predicted
 improvement of 53-66 percent compared with lime
 if pH is raised to the point of calcite saturation. This
 improvement may be attributed to the higher possi-
 ble pH as well as the lower final alkalinity. Because the
 use of caustic would increase alkalinity without con-
 comitant increases in calcium (as with lime dosing),
 the pH at which calcite becomes saturated using caus-
 tic is between that observed for lime (lowest pH) and
aeration (highest pH). ,
   Although  the  described benefits of aeration  are
expected to apply to many high-alkalinity, low-pH
                                              Predicted equilibrium soluble copper as a.function of final pH

   	r"-—"- ~"2' "•* "umai (.else acicUIUn Will 3CtU-
aliy reduce pH and thereby worsen copper corrosion
problems. Nevertheless, aeration is an attractive, low-
cost option deserving of testing at many utilities.

Conclusions
   • At constant pH, soluble copper corrosion by-
product release in relatively new copper plumbing
  is a linear function of bicarbonate concentration
  (alkalinity).               '         .
     • Qualitative and quantitative trends in soluble
  copper corrosion by-product release  are consistent
  with predictions based on Cu(OH)2 equilibrium.
     • The sensitivity of copper solubility to alkalin-
  ity (bicarbonate), expressed as mg additional soluble
  copper/mg alkalinity as CaCO3 added, depends on
  pH and temperature. Copper carbonate complexes
  are most significant in cold and low-pH waters.
     • CO2 concentrations are a surrogate for cop-
  per-carbonate complexation capacity in many
  waters, providing a possible mechanistic basis to
  correlations between dissolved CO2 and copper cor-
  rosion problems.
    • The conventional  Langelier index or Larson's
  ratio approach to. controlling copper corrosion  by-
 product release is demonstrably inaccurate.   :
    • Compared with lime or caustic addition, raising
 pH by aeration (CO2 stripping) has inherent advan-
 tages for copper corrosion control in low-pH, high-
 alkalinity waters. These advantages are attributed to
 lower final alkalinity and a reduced likelihood of cal-
 cite precipitation.

 Acknowledgment
   The authors acknowledge the financial assistance.
 of the AWWA Research Foundation (AWWARF) and
 the National Science Foundation (NSF) under Grant
 No. BCS-9309078. Any opinions, findings, and con-
 clusions or recommendations expressed in this mate-
 rial-are those of the authors and do not necessarily
 reflect the views of the NSF or AWWARF. Portions of
 this work were based on a survey by the Water Indus-
 try Technical Action Fund. Special thanks to Gregory
 Kirmeyer, Tim Chinn, Donna Dodrill, and the many
 utility personnel responsible for development of this
 database. The authors acknowledge insightful con-
versations with Werner Wolf, Tom Holm, Eilen Vik,
Jonathan Clement, Darren Lytle, Russell Taylor, and
David Nicholas that helped shape this work.
                                                                                     MARCH 1996 93

-------
   References
    1. CRUSE,
       VJWJ.AvsyAV'Uj.v/j., yjy/j*jk*wj.vuui j. vyi.ciLn.v- rvaLv-j. ova-,- .. •.; v
       t&^.fintfflatCorrosiqtiofW^                   ?* '; *• (1988); '
       ^w- /-™™,~,t,,r~ Boo^^-b Rept., AWWARF  19.  VAN DEN HO'VEN, T.J.J. Personal comm. (1993).
                  -.,-;:,.,...r _ .  1^ain-?n€er"?lfflffe; £.. ;?{l-; ^n-osiEV, I.JETAL. Breakdown of Passive Film on
                 l--Universi^t-,KarJsra                             £«;u;*.v>,,.,v,., o-i,,*:— ^—^_r_-   „ -,
                 ffS!iiEkrSa'-i^-.-:Cj-i& BJ */•*;* V*%rfc-ii8£3y5:-§-'ri%«' ,5£"1~ "• '•£
                 ^J^ai^^ife^^iisi^SSisii^^ifei^sSgssss -J
    and:
    Institut
    WlS
 2. EDWARD'S^;
                            ^^f^ll^s^ntt^LriibMAs; JlopT. &.TILLER, A.K. Formation and
    AKKKK^oo^;/5 ii^^yy4):;;.v;X7::.U.;;'.;  •.  ,.;.:r Breakdown^bf Surface Films on Copper in
 3- ^^^i^Jf^^R^P^ft^:'^6^ & 26547 r:;. f; Sodium Hydrogen Carbonate and-Sodium Chlo-
    ^"^43i^^^;£^ii:pvg^>£Jv^^;?::i. i?; ^v^:i i vnde;Solutions L Effea of Anion Concentrations.
 4. Standard M&nofeJOTtn&Examin^
   ^^^^l?^®?^^^!^                                BROSSARD, L.; & MENARD, H.
    ton, D.Cr (18th ed:;-1992):' •    -                   -      _.-•-.-.
                                                       Copper Dissolution in NaHCO3 and NaHCO3 .
                                                       NaCl Aqueous Solutions at pH 8. Jour. Electrochem.
                                                       Sac., 139:1:39 (1992).
                                                       MATTSSON, E. & FREDRKSSON, A.M. Pitting Cor-
                                                       rosion in Copper Tubes—Cause of Corrosion and
                                                       Countermeasures. British Corrosion Jour., 3:246
                                                       (1968),
                                                       COHEN, A. & MYERS, J.R. Mitigating Copper Pit-
                                                       ting Through Water Treatment. Jour. AWWA,
                                                       79:2:58 (Feb. 1987).
                                                       Moss,  G. 6- POTTER, B.C. An Investigation of the
                                                       Green-jWater Problem in Auckland, New Zea-
                                                       land, and A Discussion of Possible Remedies.
                                                       CSIRO Restricted Internal Rept., Victoria, Aus-
                                                       tralia (1984).
                                                      . REIBER, S. Personal communication (1994).
                                                  27.  ROTHBERG, M.R. ET AL. Computerized Corrosion
                                                   ;    Control. Jour. AWWA, 85:7:26 (July 1993).
                                                  28.  LANGELIER,  W. The  Analytical Control of Anti-
                                                       Corrosion Water  Treatment. Jour. AWWA,
                                                       28:1500 (1936).
                                                  '29.  SCHOCK, M.R. S- LYTLE, D.A. Effects of pH, Carr
                                                  .  '  bonate, Orthophosphate,.and Redox Potential.  .
                                                     , on Cuprosolvency. Proo NACE Corrosion/95, ,
                                                   .   Mar. 26-31/1995, Orlando^ Fla.
                                                  30. LARSON, T. Laboratory Studies Relating Mineral  •
                                                   ,  . Quality of Water to Corrosion of Steel and Cast
                           .  	_^......	_,   '    Iron. Corrosion, 41:285 (1958)!
    Cations. Wiley-Interscience, New York (1976)'.	31.' WERNER, W. ETAL. Untersechuzen zur Flachekor-
13. VUCETA, J, &--MORGAN, J.J. Hydrolysis of Cu(n).   .   ,rion in Trinkwasserteitijen aus Kupler. Wasser.
    T Twwnl f*}- f)rf0sft*stst *">-*7yt T /1 O*7T\        "  :"	• * "+ *t       -. ~^ ~ *,	'	..
  10.
 11
 12
 5. EDWARDS, M; S- FERGUSON; J.F. Accelerated Test^
   ing of Copper Corrosion. Jour. AWWA, 85:10:105
   (Oct. 1993).
 6. EDWARDS, M. ET AL. Role of Inorganic Anions,
   NOM,':and ;Water"-lreatment Processes in Cop-
   per Corrosion. Final Proj. Kept., AWWA, Denver,"
   Colo, (in'press): '•-''    '        :
 7. Database for AWWA Water'Industry Technical"
   Action Fund Project: Initial Monitoring Expe-
   riences of Large Utilities Under USEPA's Lead
   and Copper .Rule:  AWWA, Denver, Colo.
   (1993)'.! •' '^•--•••.' .  -
8. Brad Segal, Bbulder Water Department. Personal
   communication (1993);       '
9. SCHECHER/W.D'. MINEQL+: A Chemical Equilib-
   rium Program for Personal Computers. The Proc-
   tor and Gamble Co:, Cincinnati, Ohio (19.94).
   SCHOCK, M~R:/--LYTLE; D.A.;''S- CLEMENT,  J.A:
   Effects of pH, DIG, Qrthophosphate, and Sul-
   fate on Drinking Water Cuprosolvency. Res.
   Kept., USEPA;'ORD,' Risk Reduction Engrg.
   Lab., EPA/ 600/R-95/085, Cincinnati, Ohio
   (June 1995)i--^-:*^-..--., :,;•; •.,...  . . •:<••.  •::  ..- • •
   EDWARDS^ :M.;: FERGUSON, J.F.; &• REEBER; S. On the
   Pitting Corrosion of Copper. Jour. AWWA; 86:7:74
   (July 19.93)vJ-"r::'-:1.'::''-;.-•'•''• :'"' •'• '-"••   '"
   BAES, C.F..jR\ &•" MESMER, R.E.' The Hydrolysis  of
                                                    23.
                                                    24.
                                                    25
                                                   26
     Limnol. &Oceanog., 2:742 (1977).
 14. ScHiNDiERi:p;;'"REiNERii'M!; & GAMSJUGER, H. Lus-
     lichkeitskonstahten iind reie Bildungsenthalpien
     von Cu2(OH)2CO3 (Malachit) und Cu3(OH)2
                                                   ~'Abwassen, 135:2:92 (1994).
 15.
 16.
 17.
     (CO3)2 (Azurit) bei 25°C. Helvetica Chim.Acta,
     1:1845 (1968);}-' "  -'••:"'      ;   '   ;
    ZIRINO; Av^yAMAMOTO, S. A pH-Dep'endeni"
    Model for'jhe Chemical Spedation of Copper,'
    Zinc, Cadmium, and Lead in Sea Water. Limnol.
    & Oceanoijr.'; 17:661 (1972).
    DUBY, P. The thermodyhamic Properties of Aque-
    ous Inorganic Copper Systems. INCRA Mono-
    graph IV, IhtL .Copper Res. Assn., New York (1977).
    DODRILL, D. & EDWARDS; M. Corrosion Control
    Based on Utility Experience. Jour. AWWA, 87:7:74
    (July 1995).
                                                                 About the authors: Marc Ed-
                                                                 wards is an assistant professor in
                                                                 the Department .of Civil Engineer-
                                                                 ing, University of Colorado (CU),
                                                                 Boulder, CO 80309. He is a gradu-
                                                                 ate of the University ofWashington
                                                                 (Seattle) with PhD and MS degrees
                                                                 and of the University of Buffalo
                                                                 (New York) with a BS in biophysics.
                                               Michael R. Schock is a research chemist with the US Envi-
                                               ronmental Protection Agency, 26 W. Martin Luther King
                                               Dr., Cincinnati, OH 45268. At the time of this research,
                                               Travis E. Meyer was a research assistant at CU.
94 JOURNAL AWWA

-------
       Copper Corrosion and Iron Removal Plants
           Lih-in W. Rezania, P.E. and William H. Anderl, P.E.
                   Public Health Engineers

               Section of Drinking Water Protection
                Minnesota Department of Health
                   121 East Seventh Place
                  St. Paul, MM 55164-097
 ABSTRACT

 INTRODUCTION

 ssrir^sTr^^^
                                 a corrosion control
     *.
COPPER CORROSION IN MINNESOTA
All copper exceedances in Minnesota came from groundwater s^

5SS5Ssr:^^S£SSMr^

-------
Scaling Water

tanglier Index (LI) and calcium carbonate precipitation potential (CCPP) are two
indexes commonly used to predict the scaling tendency or corrosivity of a water
Scaling water conceptually is non-corrosive, unfortunately, it is no so in Minnesota
as far as copper corrosion is concerned. Seventy-five percent of the water
systems with a copper exceedance are having a positive CCPP in their treated
waters.
High Dissolved Inorganic Carbon

Dissolved inorganic carbon (DIG) and PH are the two key parameters associated
with cuprosolvency. The high dissolved inorganic carbon content in finished water
is blamed for causing the copper exceedances in Minnesota. Theoretical dissolved'
inorganic carbon contents ranging from 200 mg/L to 900 mg/L as CaCO3 were
predicted for all systems using the Rothberg, Tamburini & Winsor (RTW) Model
The majority of them have DIG between 500 and 800 mg/L as CaCO, or between
60 mg C/L and 95 mg C/L in their treated water.


Iron Removal Treatment Plants

Iron removal treatment is used across the State of Minnesota by about one-third of
the municipal public water systems in Minnesota. In most cases/the treatment
process consists of aeration/oxidation, filtration, chlorination, and fluoridation  It is
notable that among those 142 public water systems exceeding the copper action
level, 83% of the medium-size (Figure 1) and 57% of the small-size systems
(Figure 2) are iron removal plants. It is believed that the aeration/oxidation step  in
the iron/manganese removal process results in a more corrosive finished water due
to the higher dissolved oxygen levels and/or the content of oxidizing agents such
as potassium permanganate, and chlorine in the water.
IRON REMOVAL PLANTS D.O. STUDY

A study on copper corrosion of iron removal treatment systems in Minnesota was
conducted ,n summer 1993, by the Public Water Supply Unit. Nineteen (19) iron
removal filtration plants were studied for their finished water dissolved oxygen
levels associated with the aeration/oxidation process they use and their reported
90th percentile copper values.  This study focused on the way oxygen is
introduced into the water and demonstrated a strong association of 90th percentile
copper levels with the dissolved oxygen levels (figure 3)

-------
 Gravity Plants
       f' t  /em°V   flltratl°n Pl3ntS are the ones wlth c°PPer Problems. Most
       f.ltratjon systems use a combination of steps such as tray aerator, spray
aerator, grav.ty drop, flow splatter or splitter to aerate the water.  Some also use
oxidizing agents such as Chlorine and KMnO4 oxidizing agents in this
 0mananeS!, °XLdatlon. steP and followed by a gravity filtration process to
           ,OX,  .   Prec'P'tates. Finished  water dissolved oxygen levels of gravity
           l  lants are
 irorp    ,,  .             .              r  ssove oxygen le
 iron removal  plants are generally at or near the D.O. saturation point.

 In winter 1993, eight more systems were added to the study.  This study
 concluded that gravity systems are most susceptible to copper corrosion  The
 average , finished water dissolved oxygen level from fourteen (14) gravi™ plants
 was 6.5 mg/L and the average 90th percentile copper level was 2 35 rna/L  rL

           01"" r93™ Carb°n W3S 71 m9 C/L  Thetrge dissolved oxygen
            erCent'le C?Per l6Ve!S' and diss°lved inor9anic carb°n »evel from  9
             pressure plants are 1.52, 0.74, and 76 mg C/L, respectively.


 Pressure Plants

 Pressure systems use compressed  air and/or oxidizing  agent(s) to precipitate iron
 and manganese and a pressure filtration process to remove the precipitates

 exceed" 0° moT T^ "l *?* ***** *"** ^ ^^ #™ S* do not
 fivaromn™   °3" •    noted that among the 27 iron removal plants studied, the
 five compressed air aerated plants  resulted with the highest average dissolved

         Carb0neVeI °f87
 henedaTH     .                                                e in
the injected a.r. The eight non-aerated pressure plants on the other hand have the
 owest average DIG of 65 mg C/L.  Average 90th percentile copper valCe for these
two groups are .88 mg/L and 0.60 mg/L, respectively.                      '
Phosphate Treatment Survey
                  °f the ir°n rem°Val SVStems in Minnesota also treat with
   t                 C°rrOSf0n 3nd SCale C°ntroL  ln 1994' « Phosphate
treatment survey study conducted by the Public Water Supply Unit confirmed that
iron removal systems are susceptible to copper corrosion even with the presence of
phosphate inhibitors.  This survey study found the only exceptions "are svstems

-------
 SEASONAL VARIATION IN COPPER LEVEL

 In winter 1 994, 35 water systems with no first round problems exceeded the
        3A    lVeL ,H!?her COPPSr tap levels were tested in a significant number of
          Z"  £  *  6m exPerienC6d inc^ases from a less than detection limit
  <0 05 r          *                                  s   an  eecton  mt
 (<0.05 mg/L) in the f.rst round monitoring to a greater than the 1.3 mg/L copper
 ae&on level m the second testing.  Winter heating has been blamed for causing the
 elevated copper levels. The Minnesota Department of Health, Public Water Supply
 Unrt, has investigated the seasonal variation issue in following efforts.
 Demonstration Tests
        w taps were collected seasonally and analyzed for copper concentrations
        homes for an e.ghteen months period. This testing program found six
        xhibrted a clear seasonal pattern in tap copper levels, peak in midwinter
 and big drops for the rest of the year with the lowest levels tested in summer and
 round mo ^ demons,trated the on»Y elevated copper level tested was the second
 round monitonng sample, collected in January and February 1994. One site shows
  e peaks occurred in summer months.  These tests strongly demonstrate the
 conrprnfS th °°PPer ^^ '" *e f'rSt draw taPs dur|ng heating season and raised
 concerns of the corrosion control treatment issue for water systems caught up with
 a copper exceedance due to this seasonal phenomenon.


 Copper Tap Level Reproducibility

 The lead/copper monitoring results of 5311  sites were analyzed for their copper tap
 DatlT hUC   I'- S'teS W6re SeleCted °Ut of tne Lead/Copper Monitoring      P
 Database, by matching up each location with the  1st and 2nd samples taken
 approximately s,x months apart, so that a well-defined "season" can be referred to
 and the seasonal variation issue can be examined
                   Produced and Analyzed as presented in Table 1.  The average
«        H W!re hi9her '" the SGCOnd r°Und testin9-  About 8-5% of the sites
Pvtrpm    K 3 ^   9e m C°PPer l6Vel 9re&ter than °'5 m9/L in the two tests, with
to 7 1/1 Sn™'ngtH°PPerrJ.eV8lS I6SS than detectlon Iimits at one samP'ing and up
               °ther; Th'S percenta9e <8-5%> is significant due to its potential to
       R.PPer eXCeedance' Since 10% is ^ed as the trigger in the Lead and
Copper Rule.

-------
Data Set
No. of Sites
2603 Sites
2652 sites
55 sites
Overall Reprod
|Cu1 - Cu2
0.1 rng/L <
|Cu1 -Cu2
• 	 ! '"' ' " TT"
1st Round Monitoring
Sampled in
JuI/Aug 93
Sep/Oct 93
Nov/Dec 93
Ave. Cu
.18 mg/L
.29 mg/L
.21 mg/L
.
2nd Round Monitoring
Sampled in
Jan/Feb 94
Mar/Apr 94
May/Jun 94
ucibility in Two Tests:
| <• O.1 mg/L 	 	 «i QOA
|Cu1 - Cu2| ^ C
> 0.5 mg/L 	
).5 mg/L 	 30 2%
	 „ 	 	 R C;OA
==============^^
Ave. Cu
.21 mg/L
.35 mg/L
.28 mg/L
• — i
                                   Table 1
 TREATMENT EVALUATION

 The Lead and Copper Rule suggested that four treatment approaches be evaluated
 by every public water system exceeding an action level.  Evaluations, based on the
 S?6d.Water Ch![a°terJ??LCAS for systems with a c°PPer exceedance using both the
 EPA guidance and the AWWA's RTW Model, show that phosphate inhibi?or is the
 most prominent alternative for copper corrosion control in Minnesota.
Carbonate Passivation

Optimal water characteristics for employing this treatment approach for lead
control are low dissolved inorganic carbon  (DIG) and high PH. This treatment
application is limited to few lime softening  water systems for lead control
Calcium Carbonate Precipitation

The RTW Model was used to obtain the LI, DIG, and the CCPP values  These
values were used to assist public water systems in the evaluation of the suggested
corrosion control treatment alternatives required by the lead and copper rule  The
calcium carbonate precipitation corrosion control treatment approach was
eliminated by the majority of water systems due to the exhibited high CCPP values

-------
 corrosion contrnf "h the trea!fd W3ter PH appears t0 be beneflcial 1™ copper
 Silicate Inhibitor
 Phosphate Inhibitor


 Phosphate treatment is well received by most of the public water systems and















es^^
corrosion control for lead and copper (Figure 4).                optimal
TREATMENT EXPERIENCE FROM IRON REMOVAL PLANTS




                                              °f ^ ^ -mova,
Utility #1 - Treatment With Poly-Orthophosphate Blend



Population Served: 7,500



90th Percentile. Copper Level: 2.25 mg/L

-------
         Water Treatment Used: induced aeration, gravity filtration, fluoridation, chlorination

         Treatment Modification After Exceeding Cu A.L.: blended phosphate treatment

         Corrosion Control Treatment: using a blended phosphate inhibitor with a poly/ortho
         ratio of 1  at 1.5 - 2.0 mg/L as total phosphates; phosphate feed rate is constrained
         by costs and local wastewater discharge limits

         Treatment Outcome (Figure 5): immediate  reduction in tap copper levels was
         appreciated, the effectiveness started to level off after achieving 50% to 65%
         reduction.

         Water Quality Data:

             Parameters                   Finished Water
             PH                           7.6 - 7.9
             Calcium as CaCO3              150 - 180 mg/L
             Alkalinity as CaCO3           320 - 340 mg/L
             Dissolved Solids               350 mg/L
             Orthophosphate as PO4         0.8 - 0.9 mg/L

         Utility #1  demonstrates corrosion control treatment experience pf the majority of
         the iron removal systems in Minnesota who have started or switched to feeding
/        blended phosphates for corrosion control purposes. Copper tap level reduction of
         50%  or more can be easily achieved by feeding phosphate at 1 to 2 mg/L as total
         phosphate (orthophosphate ranging from 0.7 to 1.0 mg/L). However, the reduced
         copper levels remain at 1.0 to 1.5 mg/L, indicating the likelihood of exceeding
         copper action level in future monitoring  effort.


         Utility #2 - Calcium Carbonate Precipitation

         Population Served: 22,000

         90th  Percentile Copper Level: 4.08 mg/L

         Water Treatment Used Prior to Pb/Cu Monitoring: chlorination and fluoridation

         Water Treatment After Exceeding Cu A.L.: two new pressure iron removal plants
         have  been installed, treatment involving strip aeration, pre-chlorination, pressure
         filtration,  and fluoridation

         Corrosion Control Treatment: practicing calcium carbonate precipitation treatment
         approach; as a result of the aeration, water pH has increased from 7.5 to 8.1;

-------
theoretical CCPP for treated waters are 40 mg/L and 80 mg/L as CaC03.

Treatment Outcome (Figure 6): copper reduction above 50% has been achieved
contributing to the higher finished water pHs and CCPP values produced by the
new treatment plants; system remains to face the challenge in meeting the 1 3
mg/L copper action level                       •

Water Quality Data:

    Parameters	                Plant #1                pjant #2
    Raw Water pH:                7.5                   7 7
    Finished Water pH:            8.1                   8.1
    Alkalinity as CaC03:          430 mg/L              390 mg/L
    Calcium  as CaC03:            260 mg/L              150  mg/L
    Dissolved Solids:               630 mg/L              490  mg/L
    Free Ammonia mg/L as N:     2.8 mg/L              3.0 mg/L

Utility #2 demonstrates that regardless of the positive (high) calcium carbonate
precipitation potentials exhibited in the finished water,  copper levels remain above
the action level.  The appreciated copper reduction can be credited to the aeration
used that has increased pH from 7.5 to 8.1. This utility also  demonstrated a very
rare case in that copper remains as a concern at pH of 8.1.  The high ammonia
concentrations in the treated water may be the reason  for their continuing  copper
problem.                                                                K
Utility #3 - Treatment With Zinc-Orthophosphate

Population Served: 49,000

90th Percentile Copper Level: 3.36 mg/L

Water Treatment Used: utility has one pressure plant and one gravity plant; copper
problems are located in areas supplied by the gravity plant; the treatment involves
cascade aeration followed by gravity filtration, chlorination and fluoridation;
potassium permanganate and chlorine are added as oxidizing agent

Treatment Modification After Exceeding Cu A.L.: zSnc-orthophosphate treatment

Corrosion Control Treatment: zinc-orthophosphate  at 0.8 - 1.3 mg/L as ortho-PO4

Treatment Outcome (Figure 7): treatment successfully reduced copper levels to
below 10 mg/L at four test sites, achieved greater than 7O% reduction in copper
tap levels showing promise in meeting the copper action level in future monitoring

                                     8

-------
Water Quality Data:                      ;

   . Parameters              Gravity Plant          Pressure Rant

   Finished Water pH          7.1-7.5             6.8-7.0
   Finished Water D.O.        8.2 mg/L             0.8 mg/L
   Alkalinity as CaCO3         270 - 380 mg/L        200 - 310 mg/L
   Calcium as CaCO3          190 - 240 mg/L        170 - 200 mg/L
   Ortho-phosphate as PO4    0.8 - 1.3 mg/L           None ••*'•
   Dlc as C                  87 mg/L              77 mg/L

Utility #3 demonstrate a successful copper corrosion control treatment experience
using zinc-orthophosphate, achieved by a gravity iron  removal system.  Copper tap
levels were reduced below 1.0 mg/L at a relatively low orthophosphate  feed rate
(0.8-1.3 mg/L) with respect to the high level of dissolved inorganic carbon and
dissolved oxygen in the finished water. This study echoed the conclusion from the
small systems phosphate treatment survey study that  orthophosphate treatment is
the most prominent corrosion control treatment option for Minnesota's iron removal
systems.
SUMMARY AND CONCLUSIONS

1. High dissolved inorganic carbons in groundwater sources and high dissolved
   oxygen in treated water from iron removal plants are causes for copper
   exceedances in Minnesota.

2. Iron removal plants using gravity filtration are most susceptible to copper
   corrosion due to the high levels of dissolved oxygen introduced fay aeration."

3. Calcium carbonate precipitation appears to be ineffective for copper corrosion
   control.  This treatment approach has very limited application in Minnesota due
   to the,high hardness and alkalinity in groundwater

4. Seasonal variation in tap copper levels was demonstrated in Minnesota, raising
   concerns of requiring corrosion control treatment for systems exceeding the
   copper action level due to this seasonal phenomenon.

5. The common  experience using blended phosphates in Minnesota is that a 50%
   reduction in copper levels can be easily achieved. However, the lowered
   copper levels remain near the 1.3 mg/L copper action level, leaving water
   systems with continuing challenge of meeting the copper action level.

-------
6. Ortho-phosphate levels at 0.8 - 2.0 mg/L have successfully reduced copper tap
   level below 1.0 mg/L for five gravity iron removal systems. A minimum ortho-
   phosphate level  of 1.0 mg/L is recommended to be maintained throughout the
   distribution for iron removal systems treating with a phosphate inhibitor for
   copper corrosion control.
REFERENCES

1.  "Copper Corrosion Study of Iron Removal Treatment Systems in Minnesota"
   Minnesota Department of Health, Section of Drinking Water Protection, Public
   Water Supply Unit, November 1993.

2.  "Corrosion Control Treatment Survey for Small-Size Community Public Water
   Systems - Summary and Conclusion", Minnesota Department of Health, Section
   of Drinking Water Protection, Public Water Supply Unit, June 1994.

3.  Rezania, L.W. & Anderl, W.H., "Copper Corrosion and Iron Removal Plants the
   Minnesota Experience", Proc. of AWWA WQTC, 1995, New Orleans, LA
                                    10

-------
<*-•
I
en

-------
I
II

              O)

-------
                                          o
                                          ==
                                          O
                                          O
                                          o

                                          I
                                          ji>
                                          o
in


O)

3
cn
(V6n) uoiiDJ|U8Ouoo jeddoo MDJQ isj

-------
    o

    O
CM
• ^^"w»


O
 P 

                                                       a



                                                       I

                                                       o

                                                       .2
                                                       a
              (V6n)
                            (spuosnoqi)
<£>




S-

3

cn

-------
     ^__ BOMB
     DO.

     


                                                       3

                                                       05

-------
    An Evaluation of the Secondary Effects of Enhanced Coagulation, With Emphasis
                                 on Corrosion Control

                Darren A. Lytle, Michael R. Schock, Richard J. Miltner
                         U.S. Environmental Protection Agency
                               NRMRL, WSWRD, TTEB
                                 Cincinnati, OH 45268

 Abstract
       The proposed Dbinfectant-Disinfection By-product (D-DBP) Rule will require many
 water purveyors to meet enhanced coagulation or softening objectives for  total organic
 carbon (TOC) removal as a means of controlling DBF precursors.  Enhanced coagulation can
 be achieved by  performing chemical coagulation at lower pH values by increasing the
 coagulant dose, adding acid, or a combination of the two. Pilot studies using enhanced alum
 coagulation revealed a number of secondary effects, many of which directly impact lead and
 copper corrosion. Aluminum, sulfate, pH, and dissolved inorganic carbon were significantly
 altered during enhanced coagulation.  These parameters can directly affect metal solubility
 and surficial pipe film stability.

 Introduction
       Under the proposed Disinfectant-Disinfection By-product (D-DBP) Rule, many utilities
 will be required to meet enhanced  coagulation or softening objectives for total organic carbon
 (TOC) removal as a means of controlling DBF precursors.1 A recent survey found that 64 percent
 of responding coagulation plants would comply with the specified TOC removal requirements of
 the proposed rule.2 Several researchers have investigated natural organic matter (NOM) control
 by coagulation and discussed the importance of pH in removal efficiency.  Qasim et al.3 found that
 TOC removal in  natural waters by coagulation and  softening was  strongly pH dependent and
 Randtke found that removal of NOM by coagulation was optimum  at pH 5.0 to 6.0.
       Most coagulation plants that will be required to  meet TOC removal requirements under the
 proposed D-DBP  Rule will likely do so by either increasing coagulant dosage, lowering the pH
 by  acidification during coagulation, or a combination of the two.  Enhanced coagulation will
 provide improved removal of DBF precursor material, as well as secondary benefits such as lower
 chlorine demand, increased disinfectant stability in  the distribution system, lower color and
 reduced substrate for microbiological growth.3  However, negative secondary impacts of enhanced
 treatment must also be addressed to insure that other water quality parameters and treatment plant
 operations are not compromised. Some of the secondary impacts that must be considered include
 sludge production, filter run time, inorganic water  quality changes, turbidity and particulate
 removal efficiency, and distribution system corrosion and corrosion  control.
       The U.S.  Environmental Protection Agency's (USEPA's) National Risk Management
 Research Laboratory (NRMRL), in Cincinnati, Ohio, in conjunction with the University of
 Cincinnati,  conducted bench- and pilot-scale tests to evaluate enhanced coagulation for optimal
NOM or DBF precursor removal  from a surface water using aluminum sulfate, more commonly
referred to as "alum".  Results of those studies directly related to the goals of the D-DBP Rule
have been previously discussed.5-6  This paper explores several secondary impacts relative to
corrosion control  of enhanced alum coagulation based on observed water parameter changes

-------
during pilot-scale studies.  Specifically, the effect of pH, dissolved inorganic carbon (DIG) and
alkalinity, sulfate and chloride (during ferric chloride coagulation), total organic carbon (TOC),
and aluminum on the corrosion and corrosion control will be discussed.

Experimental
       Based on  preliminary jar tests studies6, East Fork Lake water (Cincinnati,  Ohio) was
selected for pilot testing.  Two parallel pilot-scale treatment plants  consisting of rapid mix,
flocculation,  and sedimentation basins were used to compare conventional and  enhanced
coagulation as shown in Figure 1.  Settled water was  chlorinated  and sand filtered.   The
adjustment of pH for corrosion control was made after filtration in the conventional plant.  The
enhanced plant was split into two parallel filters where pH was adjusted before one filter and after
the other.  Final pH in all cases was adjusted to approximately 8 using NaOH. Chlorine was fed
as NaOCl.  A number of water quality parameters were monitored  at a variety of locations
throughout the plant and are given in Table 1.  Operational data including chemical doses,
temperature, head loss development and filter run time were also regularly monitored.   The
sampling described in Table 1 took place on six  days between September 7 and September 13,
1994.  East Fork Lake water was trucked to the USEPA research facility daily during pilot plant
operation.

Results
       Because the results of the pilot study have been presented elsewhere5'6 and page constraints,
data will not be presented or analyzed in great detail. A number of secondary impacts  (both
positive and negative) beyond the corrosion impacts  addressed in this report have also been
previously identified in some detail.7  Positive secondary impacts of enhanced coagulation
included: reduction in turbidity, particle counts,  coliform and heterotrophic bacterial densities,
and chlorine  residual.  Increased cost and sludge production, and potential future regulatory
conflicts such as the proposed sulfate standard were identified among negative secondary impacts.
This  document  lists  the differences  in  water  quality  between conventional and  enhanced
coagulation treatment modes that are considered important in lead and copper corrosion control.
       Table 2 shows the major average water quality parameter  measurements;'throughout the
treatment trains.  It  should be pointed out that in the case of East Fork Lake water, only 35
percent TOC  removal would be required given the TOC (4.8 mg/L)  and alkalinity (99 mg
CaCO3/L) of  this water.8  It is important to note that in these pilot studies, conventional
coagulation (29 percent removal of TOC) was compared to optimum coagulation (54 percent
removal of TOC). Enhanced coagulation in the regulatory sense  (35 percent removal of TOC)
was not studied.  Thus, findings regarding secondary  effects must be viewed as examining the
extreme shift from conventional to optimum coagulation rather  than the expected  shift from
conventional to enhanced coagulation.
       Alum doses were 44  mg/L (3.7 mg/L as aluminum) and 152 mg/L (13.3 mg/L as
aluminum) for conventional and optimized coagulation, respectively,  as shown in Table 2.
Aluminum concentrations in settled waters were lower after optimized coagulation than after
conventional coagulation (0.47 mg/L versus 0.65  mg/L) despite the  higher alum dose.  This
observation  follows  the established pH-aluminum hydroxide solubility relationship.9"13   The
location of pH adjustment was clearly the most important factor impacting aluminum residuals
after .the filters, as shown in Figure 2.  When pH adjustment was practiced following filtration

-------
  (filters Fl and F2), the lowest clearwell aluminum residuals were observed.  This results from the
  ower solubility of aluminum at lower pH values, and retention of unsettled residual alum floe by
  Ac filters.  However^ when the pH was raised to 7.95 prior to filtration (filter F3), aluminum
  solubility was increased, enabling the passage of more dissolved aluminum into the clearwell  The
  trade off was that filter run times were shorter in the lower pH waters
  fiir«f • Alun™umrle;;els were ** lowest 0« than 0.025 mg/L) following optimized coagulation,
  filtmion and pH adjustment in clearwell 2.  This can be explained by the fact that the pH of
  settied water entering filter 2 was 6.90, which is closest to the minimum solubility of aluminum
  hydroxide.   The highest aluminum levels were observed following optimized coagulation, PH
  adjustment  and filtration because the PH of the water (pH=7.95) was the highest, resulting in
  the most soluble aluminum entering the filters.  Aluminum concentrations in filter effluents were
  used  to  develop  the empirical pH-aluminum solubility  relationship shown by Figure 3  This
        Sulfate increased after both coagulation approaches as a result of the addition of alum as
 seen m Table 2.  Optimally coagulated water contained at least 40 mg SO42VL more than
 conventionally treated water. The point of pH adjustment had no significant impact on sulfate
 concentration.
        In this study, the pH needed to meet optimum enhanced coagulation conditions was 0 7
 I  "u     M6'  *?    PH dUring conventional coagulation. Without pH readjustment, the drop
  urf^Z   T V" -KCr^Sed ^ md C°Pper ****** ** Stabilization of existing pip^
 surface films in the distribution system. Thus, pilot plant finished water pH values were adjusted
 to near 8 to represent reasonable conditions under the Lead and Copper Rule and to drive DBF
 reactions under representative conditions. .
 imnnrt '" ^^f * the&ffQCi of the location of pH adjustment to aluminum residual, another
 important consideration is the impact of the pH adjustment chemical on finished water quality

      eS
 issuesnR      ,                             '                                   ng
 rnnT ; ?   ^Jjustment chenucals such as caustic soda (NaOH) or soda ash (Na2CO3)  wiU
 contnbute sodium to the finished water.   Although lime was not used in this study for pH

         im uritiel  mCreaS£ ^^ ^^^ ^ may contribute *"™™™ as a result of
d^r,  ^ ^^f Cfb°n CnQ W3S n0t ^^^ ^ «»vw»tional coagulation; however, TIC
decreased significantly from optimized coagulation (see Table 2). The amount of TIC decrease
                               mg C/L) is clearly far beyond
Corrosion impacts

hnth   ^ effeCt f PH Snd DIC °D Iead and C0pper corr<*ion. DIG has been shown to have
both a positive and negative impact on corrosion control.15'17  DIC serves to control the buffer

£™,!L* ri -^f SyStemS' ^ theref°re' SUfflcient DIC is necessary to Provid^ ^ stable pH
throughout the distnbubon system for corrosion control of copper and lead.17'19  However larger
amounts can result in increased lead and copper solubility.17'20
       The importance of pH to lead and copper solubility is well-established.  Therefore,' the
reduction of pH during enhanced coagulation will likely have to be counteracted by PH adjustment
following treatment to maintain corrosion control objectives outlined under the Lead and Copper

-------
Rule21"23  or state or  local wastewater  discharge  requirements  for  copper.    Without pH
readjustment, the drop in pH from coagulation would result in increased lead and copper solubility
and destabilization of existing pipe surface films in the distribution system, if corrosion control
is currently being practiced through pH/DIC adjustment.  Even the effective use of orthophosphate
or blended  phosphate chemicals for lead or copper control may require a higher pH than that
achieved during enhanced coagulation.15'17
       Considerable research has been devoted to the study of the aqueous speciation of copper
in natural waters and seawater, in which the role of carbonate species was extremely  important.24"30
The concept of the significance of carbonate complexation was then applied to drinking waters
with indications that high DIG levels would aggravate cuprosolvency.31'32 Recently, a variety of
new research has helped define the complicated interrelationships of pH, DIC, orthophosphate,
the formation of metastable solids, and plumbing  material age on cuprosolvency and observed
copper levels in drinking waters. 19-20-33JW
       Figure 4 illustrates the effect of DIC and pH on cupric hydroxide solubility, as would be
the case with relatively young plumbing systems.  An increase in copper solubility with lower pH
and higher  DIC is evident from the  figure. Above a pH of approximately 9.5, an upturn in
solubility is predicted, caused by carbonate and hydroxide complexes increasing Cu(OH)2(s)
solubility.  In the pH range of approximately 6.5 to 9, significant increases in copper solubility
are predicted from the addition of even small amounts of carbonate, although maximum solubility
remains less than about 0.3 mg/L.
       Figure 5 shows a solubility diagram for copper(II), corresponding to equilibrium with
either Cu2(OH)2CO3(s)  (malachite) or CuO(s) (tenorite), whichever is thermodynamically stable
at a given pH/DIC point.  This kind of situation represents the case with aged plumbing, though
the number  of years of exposure needed to achieve these stable solids likely depends greatly on
the water chemistry of the system.19-34 Several important contrasts in cuprosolvency behavior
between the case represented by Figure 4 as opposed to the assumptions behind Figure 5 should
be noted. If Cu2(OH)2CO3 (malachite) is present and capable of forming, below a pH of about
6.5 the addition of DIC  is predicted to decrease cuprosolvency, but increase cuprosolvency above
about pH 7.  There is a  small transition zone between these values where the first approximately
5 mg/L of DIC should slightly reduce copper solubility, but additional carbonate would decrease
it or  have essentially no effect.  In  an aged system, below a pH of about 8, cuprosolvency
becomes essentially insensitive to DIC above approximately 30 mg C/L.  Malachite formation
would enable attainment of 1.3 mg/L after stagnation below approximately pH 6.5 for all DIC
levels.  This is in stark contrast to the effect of pH and DIC when only cupric hydroxide is
formed, where a pH of over 7 would be necessary to stay under 1.3 mg/L for long stagnation
times at very low DIC levels, and over 7.5 for systems with high DIC.
       In a  simple system of metallic lead immersed in water containing dissolved carbonate
species, the solubility will be controlled by either the simple lead carbonate (PbCO3, cerussite)
or one of the two basic lead carbonates, Pb3(CO3)2(OH)2 (hydrocerussite) or Pb10(CO3)6(OH)6O
(plumbonacrite).  Above approximately pH  12.5,  lead hydroxide (PbtCH)^ may form .  The
conventional solubility constants for lead hydroxide reported in the literature vary over about five
orders of magnitude (log  K,,, from -19.85 to -14.9).  Samples of deposits from potable water
systems and accompanying dissolved lead concentrations  do not indicate either lead oxide (PbO,
massicot or  litharge)  or lead hydroxide to be active in governing lead solubility, though PbO is
frequently found in underlayers of surface films. Under  very oxidizing conditions, the lead(TV)

-------
 solid PbO2 may form, and it has occasionally been found in pipe deposits. PbO2 is less soluble
 than any of the Pb(II) carbonates or hydroxycarbonates.
       When only pH and carbonate concentration effects are considered, lead solubility is shown
 by Figure 6.   Detailed computations show that the minimum lead solubility  is at a pH  of
 approximately 9.8, with a DIG concentration of between 3.6 and 4.8 mg C/L (0.0003 to 0.0004
 M).15'17-31  Above approximately pH 8, the closeness of the lines and steepness of the slopes
 represent the high sensitivity to both pH and DIG. These properties, plus the reversal of solubility
 trends above  about 5 mg  C/L DIG result from  the stability of the solid  Pb3(CO3)2(OH)2
 (hydrocerussite) plus strong hydroxide and carbonate complexation. The downward-trending lines
 for pH 6 and  7  are caused by the formation of PbCO3 (cerussite), and an absence of carbonate
 complexation.  In  the lead carbonate stability region, below approximately pH 8 and above
 approximately 25 mg C/L, lead solubility is not very sensitive to either pH or DIG, although the
 trend is toward slightly lower levels as DIG is increased below about pH 7.5.  Detailed discussions
 of lead solubility controls for drinking water  are readily available elsewhere. 15-l7il8'31>41>42
       The effect of chloride, sulfate, and aluminum on lead and copper corrosion.  The
 effect of chloride, sulfate and aluminum on copper corrosion has recently come under renewed
 investigation, as  a result of the interest in controlling cuprosolvency for drinking water or waste
 water regulations.   Interesting results  were obtained from  X-Ray diffraction and Energy-
 Dispersive X-Ray (EDXA) analyses of the deposits formed on experimental pipes used in DWRD
 pipe loop studies.19-34 Significant diffraction peaks for Cu4(OH)6SO4-H2O(posnjakite) were found
 in pipes from experiments at pH 8 and pH 9 with 5 mg C/L DIG, even though sulfate levels were
 only  approximately 30 mg/L. This mineral has also been reported in  some copper pipes in a
 German study of pipe corrosion in a hospital.43  Diffraction peaks likely, corresponding to the solid
 Cu(Cl,OH)2-2H2O (calumetite) were found, particularly at pH 8 and 7.  Qualitative elemental
 analysis confirmed the presence of S, Cl and also Al on the pipe surfaces. A large peak and a
 secondary peak apparently consistent with the solid CuAl4SO4(Oir)12-3H2O (chalcoalumite) were
 found on the pH  7 specimen, but only a corresponding minor peak was found on the pH 8 sample.
 Additional peaks for Cu2(OH)2CO3, CuO, and Ci^O were also identified. From the qualitative
 elemental analysis, the Al concentration on the pipe appeared higher at pH 7, consistent with the
 general trend in solubility of many aluminum  minerals, and a statistically-significant decrease in
 aluminum in the water during the pH 7 experiment. The presence of aluminum on the pipe is also
 noteworthy because of the low Al concentration in the water in all three experimental systems (<
 0.1 mg/L), which suggests possibly a strong role for aluminum in the formation of natural
 diffusion barriers in plumbing and distribution systems.
       Examination of copper leaching results from DWRD studies at pH's > 8 show the copper
 levels were consistently above the solubility  of copper predicted by cupric hydroxide or oxide
 models when elevated levels of sulfate (> 30 mg SO42VL) were present.19-34 An example of this
effect is shown in Figure 7, representing data from 72-hour stagnation samples from pH-adjusted
Cincinnati tap water containing approximately 70-120 mg/L of sulfate.19'34  A detailed literature
 investigation suggested that above some threshold combination of pH and sulfate concentration,
metastable hydroxysulfate solids may form, rather than the more protective cupric hydroxide or
tenorite.19-34  Edwards, et al.33>4° showed that sulfate increased copper corrosion rates in water.
Rehring and Edwards44 attributed higher copper corrosion rates in enhanced coagulated waters to
additional sulfate carryover from the alum coagulant and also showed lower copper corrosion rates
from  chloride carryover  from ferric chloride enhanced coagulation. The effect of sulfate on

-------
copper corrosion is significant; however, the degree sulfate enhances copper solubility and effects
copper action level exceedences is not fully known. The greatest negative impact of sulfate on
cuprosolvency seems to be for waters with a pH over 8, based on early USEPA experiments at
low DIG.  Increases in copper solubility resulting from increased sulfate levels may more likely
be a concern for wastewater discharge exceedences of copper.
       To add to the confusion about sulfate effects, equilibrium chemical modeling shows that
there is some  potential  that the  same cupric hydroxysulfate solids that  cause increased
cuprosolvency above pH 8 may actually reduce cuprosolvency below approximately pH 7, relative
to the solubility of cupric hydroxide on young plumbing.19 This hypothesis needs experimental
verification.
       Several researchers have considered possible effects on lead solubility of simple anions,
such as nitrate, chloride,  and sulfate.  Beccaria et al.45 used a variety of electrochemical and x-ray
techniques to study the corrosion of lead in seawater, which provided an extreme case for effects
of chloride and sulfate.  During the initial passivation stage, Pb(OH)Cl(s) and Pb3(CO3)2(OH)2(s)
were found to be constituents in the deposit. After long immersion periods,  PbCO3-PbCl2(s),
Pb2O3(s), PbO(s),  and PbCl2(s) were also found. The presence of sulfate ions did not interfere
with the film formation, and sulfate ions did not precipitate to form a lead sulfate solid on the
surface.  In solutions with extremely high sulfate concentrations, some basic lead sulfate solids
were found, notably Pb3(SC>4)2(OH)2 and an unidentified compound with a 1:1 oxide:sulfate ratio.
       Lead(E[) forms somewhat weaker complexes with sulfate and chloride than with carbonate,
either of which might occasionally be found in water supplies at a sufficiently high level to impact
lead solubility.  The greatest impact, if it happened, would be at relatively low pH and low DIG
levels, where less lead is complexed by hydroxide and carbonate  species.  Calculations  were
performed for chloride, sulfate, and for both sulfate and chloride at DIG levels of 3-30. The  ionic
strength of the waters for these modeling calculations was set to 0.05 mol/L, to allow higher
chloride and sulfate concentrations more plausibly.  However, there appears to be no significant
impact of practical proportions is decided based on solubility.
       Ironically,  reducing aluminum residuals  could cause the destabilization of Al-containing
films built up in domestic plumbing over many years  of normal plant operation. These films have
been suggested as having been beneficial to lead and copper leaching from plumbing.17>42>447  One
study has observed an increase in lead mobilization apparently resulting from sloughing-off of Al-
based pipe scales.47
       Considerable recent evidence for a corrosion-reducing aluminum or aluminosilicate film
has been described in a study conducted by the Denver Water Department, where significant
precipitated coatings were inhibiting lead and copper release from corrosion control pipe loop
study test rigs.46 For utilities adding sodium silicate as a corrosion inhibitor, residual aluminum
from conventional coagulation may enable additional  protective films  to form.  The solubility and
complexation chemistry of aluminosilicates is highly complex and somewhat controversial  in its
details.48"52  However, aluminosilicate minerals clearly are important naturally-forming solids, and
are geochemically plausible for many drinking waters with a near-neutral to slightly acidic pH.
       An adverse affect of aluminum on  copper piping in hot water lines was reported by
Tunturi,53  who found pitting failures associated with deposition of aluminum hydroxide from a
drinking water with 0.3 to 0.7 mg/L.
       As  described earlier, aluminum has  a tendency  to  form  several highly-insoluble
orthophosphate compounds under certain chemistry conditions. Whether or not the formation of

-------
 such solids would occur in the distribution system if a water system was dosing orthophosphate
 is not known, and requires more investigation. Depending upon the surface interaction properties
 of any of these solids formed, the result could be an increased protection from additional surface
 film formed, increased turbidity in hot or cold water, or depletion of orthophosphate dosed and
 consequent lessening of passivation of lead or copper.
       The effects of NOM on lead and copper corrosion.  A goal of the D-DBP Rule is lower
 TOC concentrations prior to chlorination, so that lower DBF concentrations of regulated and
 unregulated DBFs will result.1 The impact of TOC on lead and  copper  solubility and in the
 structure of protective films, however,  may be significant.
       Many investigators have looked at interactions between Cu2"1" and dissolved natural organic
 matter.   Some studies suggest that while copper can form complexes with NOM ligands
 (sometimes called DOM for "dissolved organic matter"), at concentrations of copper typical of
 drinking  waters, copper  speciation  is more  likely dominated by hydrolysis or carbonate
 complexes.54'56.
       Cu2+ ion may also bind with adsorbed organic material containing appropriate functional
 groups. The binding with adsorbed organic matter seems to be stronger than direct binding  with
 surface sites on several materials tested.57  Copper present as an organic complex may  bind
 preferentially with adsorbed organic material.57-58  These studies  suggest that some reduction in
 copper concentration may be caused by adsorbed organic  material acting  as either a diffusion
 barrier, or as a less-soluble corrosion film.
       Some other studies suggest, however, that NOM may play a major role in the aqueous
 speciation of cupric ion, particularly when carbonate concentrations are low.59-60  Organic ligands
 produced by marine diatoms and  during diatom blooms have been shown to strongly complex
 copper, though usually at low copper concentration.61-62 Unsaturated organic ligands were also
 shown in experiments at pH's generally lower than drinking water  to increase the dissolution rate
 of copper metal in the presence of cupric ion, by a complicated interaction affecting the electron
 transfer rate between Cu(s) and Cu2+, and by stabilizing the Cu(I) state by  complexation.63  '
       The significance of NOM  to cuprosolvency relative to drinking water concentrations of
 copper and competing non-metals and ligands has not been conclusively determined, though it is
 an area under active investigation  by some research groups in the United States. Research into
 copper plumbing pitting has indicated that some NOM may actually alleviate the propensity of a
 water to  cause pitting attack, and possibly alter some scale  formation characteristics of uniform
 copper corrosion.64 Any effect on cuprosolvency will likely be stronger in untreated surface water
 supplies  than in ground waters having very low TOC, or historically coagulated and filtered
 surface waters.65-66
       In many potable water systems, the behavior of lead is complicated by its interaction with
organic ligands, or colloidal material, or both.67  Some other studies have found lead associated
with flakes of nonadherent corrosion products related to iron and organic material.68
       Samuels and Meranger69 found concentrated  "fulvic acid" solutions of pH 6.2 to be
aggressive toward lead in  solder and brass, and Thresh70 believed that organic material in the
water could prevent the deposition of films on lead surfaces by enhancing the solubility of lead.
Miles found that the organic content of moorland waters appeared to accelerate the initial attack
on the lead surface.71   Harrison and Laxen67 analyzed water from three localities in  the United
Kingdom and found that significant  quantities of dissolved  lead (over 50 percent) in some samples
appeared to be found  in organic  complexes.  Therefore,  simple solubility computations  that

-------
include only inorganic species may underestimate the levels of dissolved, lead attainable in a
distribution system.  Including the problem with associations with colloids or nonadherent scales,
even though the inorganic dissolved lead equilibria may be accurately predicted given enough
information, the accurate prediction of observed levels of total lead may not be possible.
       In contrast to NOM species that may enhance metal solubility, organic material that can
adsorb or otherwise adhere  to the inside surfaces of the pipe could serve to decrease  lead
solubility. Theoretical interpretations are hindered by the lack of characterization of the numerous
types of organic materials involved. Some simple experiments by Sheiham and Jackson41 on two
waters with varying degrees of natural color suggested that only highly colored  waters might have
a significant impact on lead corrosion. They emphasized that this conclusion must be treated with
extreme  caution because only two  waters' were  tested  and because  the organic component
characteristics might vary  from place to place.   Campbell has unambiguously expressed the
opinion  that at least some NOM species provide an  effective  natural inhibitor for copper
dissolution,65'66 and that flocculation can remove the protective substance.72
       In drinking water systems,  the presence of extensive amounts of metal piping as a source
creates  a considerably different environment than natural  water or aquifer  systems where
NOM/copper complexation has been  studied most extensively.  In addition to pH and  ionic
strength, which have been widely acknowledged to be important in complexation studies with
NOM, consideration must also be given to the considerable competitive role that cupric carbonate,
hydroxide, and hydroxide/carbonate complexes must play,  particularly as the pH increases above
7. Further, depending on source water and disinfection conditions, the redox potential of drinking
waters can vary over a wider range than  natural systems, and a variety and quantity of solids may
exist that are also not present in ground waters, lakes and sea water.
       The behavior of organic ligands can be modified by their association  with or adsorption
on particle surfaces, and some natural ligands, such as fulvic acid, may enhance metal  removal
by precipitation.73'74  NOM species interactions with surfaces and dissolved metal ions are greatly
affected  by  the concentrations of other metals and ligands in the ionic medium, with their own
competitive complexation and surface-binding effects, and the relative rates  of reactions of the
NOM ligand species with trace metals as opposed to major ions such as calcium.74'75  Aluminum
in conjunction with sulfate may be  an important agent in assisting particle formation and possible
surface adsorption of NOM.76  Adsorption films on actual pipe surfaces may take dozens of years
to form, so they are not amenable to laboratory study. Studies are further complicated by the fact
that the natural ligand or adsorbing molecules are difficult to identify and quantify, and may be
altered by isolation from their natural water chemistry matrix.

Conclusions
       A number potential  secondary impacts on lead and copper corrosion have been identified.
Specific negative impacts on lead  and copper corrosion include: (1) Possibly reduced protection
from NOM and Al-containing diffusion barrier or passivation films,  (2) Increased corrosion
control costs from additional pH adjustment and possible need for corrosion inhibitor additions,
and (3) Increased sulfate levels causing increased cuprosolvency under some pH/DIC conditions.
       In this study, the use of ferric chloride or other iron-based coagulants was not investigated
in similar detail to alum.  A general prediction about its effect is that  chloride effects on copper(II)
or lead(II) corrosion, at the likely chloride levels,  should be insignificant.
       The removal of NOM and absence of aluminum will cause the potential for passivation or

-------
diffusion barrier film destabilization as noted above.  The long-term practical impact of this
potential is unknown, as it depends on the unknown tendency for reversibility of the different
existing pipe deposits in the distribution systems.
       In contrast to the generally negative anticipated impacts of enhanced coagulation there are
two anticipated benefits.  First, if a water supply has particular NOM species that are effective
complexants for lead and copper, reduced metal levels might ultimately result after complete
implementation of the combination of coagulation changes and corrosion control. Second, there
may be reduced aluminum residuals in the distribution system, if those levels are of concern for
reasons of health, turbidity control, or possible hot water pitting enhancement.

Acknowledgments
       The authors thank the managers of East Fork Lake State Park for their cooperation  in
providing water and information for the pilot plant investigation.  They also acknowledge the
University of Cincinnati undergraduate and graduate students, USEPA co-op students, technicians,
and chemists, and DYNCORP/TAI contract staff for their contributions to pilot plant operation,
data collection, data analysis,  and  report  preparation.   This  work was funded through a
combination of USEPA in-house funds and cooperative agreement CR 816700 between USEPA's
Drinking Water Research Division and the University of Cincinnati.

Disclaimers
       Mention of commercial names does not constitute endorsement or recommendation by the
agency. The views expressed in this paper are those of the authors,  and do not necessarily reflect
USEPA policy.

References

1. Pontius, F. W. "D/DBP Rule to Set Tight Standards". Journal of the American Water Works
       Association, 85:11:22 (1993).
2. Randtke, S. J.   A Comprehensive Assessment  of  DBP  Precursor Removal by Enhanced
       Coagulation and Softening. AWWA Annual Conference, New York, NY, June 19-23, 1994.
3,. Qasim,  S.  R. "TOC  Removal by Coagulation and  Softening". Journal of Environmental
       Engineering, 118:3:432(1992).
4. Randtke, S. J. "Organic Carbon Removal by Coagulation and Related Process Combinations"
       Jour..AWWA, 80:5:40 (1988).
5. Tryby, M. E., Miltner, R. J. & Summers, R. S. TOC Removal as a Predictor of DBP Control with
       Enhanced Coagulation.  Proc. AWWA  Water Qualtity Technology Conference,  Miami, FL,
       November 7-11, 1993.
6. Miltner,  R. J., et al.  The Control of DBPs by Enhanced Coagulation. Proc. AWWA Annual
       Conference, New York, NY, June 19-23, 1994.
7. Lytle, D. A., et al.  An Evaluation of the Secondary Effects of Enhanced Coagulation. Proc.
      AWWA Specialty Conference on Enhanced Coagulation, Charleston, S.C., Dec. 5-6, 1994.
8. USEPA.  "Draft  Guidance Manual for Enhanced Coagulation and Enhanced Precipitative
       Softening", Office of Ground Water and Drinking Water, Washington, D.C., 1993).
9. Qureshi, N. & Malmberg, R. H. "Reducing Aluminum Residuals in Finished Water".  Journal of
       the American Water Works Association, 77:10:101 (1985).

-------
 10. Letterman, R. D. & Driscoll, C. T. "Survey of Residual Aluminum in Filtered Water". Journal
       of the American Water Works Association, 80:4:154 (1988).
 11. Simpson, A, M. "Aluminum, Its Use and Control in Potable Water". Environmental Technology
       Letters, 9:907 (1988).
 12. Jekel, M R. & Heinzmann, B. "Residual Aluminum in Drinking Water". AQUA, 38:281(1989).
 13. Letterman, R. D. & Driscoll, C. T. Control of Residual Aluminum in Filtered Water. AWWA
       Research Foundation, Denver, CO, (1994).
 14. Schock, M.  R. & George, G. K.  Comparison of Methods for Determination of Dissolved
       Inorganic Carbonate (DIG). Proc. AWWA Water Quality Technology Conference, Orlando,
       FL, November 10-14, 1991.
 15. AWWARF. Lead Control Strategies. AWWA Research Foundation and AWWA, Denver, CO,
       (1990).
 16. Gregory, R. & Jackson, P. J.  Central Water Treatment to Reduce Lead Solubility. Proc. AWWA
       Annual Conference, Dallas, TX, June 10-14, 1984.
 17. Schock, M. R. "Understanding Corrosion Control Strategies for Lead".  AWWA, 81:7:88 (1989).
 18. Schock, M. R.  "Internal Corrosion and  Deposition Control". Ch. 17 In:  Water Quality and
       Treatment: A Handbook of Community Water Supplies. McGraw-Hill, Inc., New York,
       (Fourth ed., 1990).
 19. Schock, M. R., Lytle, D. A. & Clement, J. A.  "Effect of pH, DIG, Orthophosphate and Sulfate
       on Drinking Water Cuprosolvency", U.  S. EPA Office of Research and Development,
       Cincinnati, OH, EPA/600/R-95/085.1995).
20. Schock, M. R. & Lytle, D. A. The Importance of Stringent Control of DIG and pH in Laboratory
       Corrosion Studies:   Theory and Practice.  Proc.  AWWA Water Quality Technology
       Conference, San Francisco, CA, November  6-10, 1994.
21. Lead and Copper.  Final Rule. Fed. Reg. 56:110:26460 (June 7, 1991).
22. Lead and Copper.  Final Rule Correction. Fed,  Reg. 56:135:32112 (July 15, 1991).
23. Lead and Copper.  Final Rule Correction. Fed.  Reg. 57:125:28785 (June 29, 1992).
24. Paulson, A. J. & Kester, D. R. "Copper(II) Ion Hydrolysis in Aqueous Solution". J. Solution
       Chem., 9:4:269 (1980).
25. Symes, J. L. & Kester, D. R. "Thermodynamic Stability Studies of the Basic Copper Carbonate
       Mineral, Malachite". Geochim. Cosmochim. Acta, 48:2219 (1984).
26. Rickard, D. T. The Chemistry of Copper in Natural Aqueous Solutions. Acta Universitatis
       Stockholmiensis, Contr. Geology 23:1:64 (1970).
27. Mann & Deutscher, R. L. "Solution Geochemistry of Copper in'Water Containing Carbonate,
       Sulfate and Chloride Ions".  Chem. GeoL, 19:253 (1977).
28. Lindsay, W. L. Chemical Equilibria in Soils. John Wiley & Sons, New York, New York, (1979).
29. Byrne,  R.  H. & Miller, W. L. "Copper(II) Carbonate Complexation in Seawater".  Geochim.
       Cosmochim. Acta, 49:1837 (1985).
30. Byrne,  R. H., Kump, L. R. & Cantrell, K. J. "The Influence of Temperature and pH on Trace
       Metal Speciation in Seawater". Marine Chemistry, 25:163 (1988).
31. Schock, M. R. & Gardels, M. C. "Plumbosolvency Reduction by High pH and Low Carbonate--
       Solubility Relationships".  Journal of the American Water Works Association, 75:2:87 (1983).
32. Schock, M. R. Treatment or Water Quality Adjustment to Attain MCL's in Metallic Potable
       Water Plumbing Systems. Plumbing Materials and Drinking Water Quality: Proceedings
       of a Seminar, Cincinnati, OH,  May 16-17,1984, 1985.

-------
 33. Edwards, M, Meyer, T. & Rehring, J. P.  Effect of Various Anions on Copper Corrosion Rates.
       Proc. AWWA Anrnial Conference, San Antdhio, TX, June 6-10, 1993.
 34. Schock, M. R., Lytle, D. A. & Clement, J. A.  Modeling Issues of Copper Solubility in Drinking
       Water. Proc. ASCE National Conference on Environmental Engineering Boulder CO July
       11-13,1994.                                                           '
 35. Meyer, T. E. & Edwards, M. Effect of Alkalinity on Copper Corrosion. Proc. ASCE National
       Conference on Environmental Engineering,  Boulder, CO, July 11-13,  1994.
 36. Rehring, J. P. The Effects of Inorganic Anions, Natural Organic Matter, and Water Treatment
       Processes on Copper Corrosion. Master of Science, University of Colorado at Boulder
  .   :  Boulder, Colorado (1994).
 37. Schock, M. R., Lytle, D. A. & Clement, J. A. Effects of pH, Carbonate, Orthophosphate and
       Redox Potential on Cuprosolvency. NACE Corrosion/95, Orlando, FL, March 26-31, 1995.
 38. Dodrill, D. M. & Edwards, M. A General Framework for Corrosion Control.  Proc. AWWA
       Water Quality Technology Conference, San Francisco, CA, November 6-10, 1994.
 39. Dodrill, D. M. & Edwards, M. "Corrosion Control on the Basis of Utility Experience". Journal
       of the American Water Works Association, 87:7:In Press (1995). ..
 40. Edwards, M., Meyer, T. E. & Schock, M. R. "Alkalinity, pH and Copper Corrosion By-Product
       Release",  submitted manuscript(1995).
 41. Sheiham, I. & Jackson, P. J. "Scientific Basis for Control of Lead,in Drinking Water by Water
       Treatment".  Journal of 'the Institute• of'Water Engineer-sandScientists, 35:6:491 (1981).
 42. Schock, M. R.  & Wagner, I. "The Corrosion  and Solubility of Lead in Drinking Water". Ch. 4
       In:  Internal Corrosion of Water Distribution Systems. AWWA Research Foundation/D VGW
       Forschungsstelle, Denver, CO, (1985).
 43. Wagner, D., Fischer, W. & Paradies, H. H. "Copper Deterioration in a Water Distribution System
       of a County Hospital in Germany Caused by Microbially Influenced Corrosion-II. Simulation
       of the Corrosion Process in Two Test Rigs Installed in this Hospital"   Werkstoffe und
       Korrosion, 43:496 (1992).
 44. Rehring, J. P. & Edwards, M. The Effects of NOM and Coagulation on Copper Corrosion.
       Proc. ASCE National Conference on Environmental Engineering, Boulder, CO, July 11-13,
       1994.                                              . . .
 45. Beccaria, A. M., etal. "Corrosion of Lead in Sea Water". British Corrosion Journal 17'2'87
       (1982).                                                                '
 46. Lauer, W.  C. & Lohman,  S. R.  Non-Calcium Carbonate Protective Film Lowers Lead Values.
       Proc. AWWA Water Quality Technology Conference, San Francisco, CA, November 6-10
       1994.
 47. Fuge, R. & Perkins, W. "Aluminum and Heavy Metals in Potable Waters of teh North Ceredigion
       Area, Mid-Wales".  Environmental Geochemistry and Health, 13:2:56 (1991).
48. Paces, T. "Reversible Control of Aqueous Aluminum and Silica During the Irreversible Evolution
       of Natural  Waters". Geochimica et Cosmochimica Ada, 42:1487 (1978).
49. Fanner, V. C., Fraser, A. R. & Tait, J. M. "Characterization of the Chemical Structures of Natural
       and Synthetic Aluminosilicate Gels and Sols by Infrared Spectroscopy".  Geochimica et
       Cosmochimica Acta, 43:1417 (1979).
 50. Neal, C., et al. "Aluminum Solubility Controls in Acid Waters: the Need for a Reappraisal".
       Earth and Planetary Science Letters, 86:105  (1987).

-------
51. Neal, C. & Williams,  R.  J. "Towards  Establishing Aluminum Hydroxy Silicate Solubility
       Relationships for Natural Waters". Journal of Hydrology, 97:347 (1988).
52. Farmer, V. C. & Lumsdon, D. G. "An Assessment of Complex Formation between Aluminum
       and Silicic Acid in  Acidic Solutions".  Geochimica et Cosmochimica Acta,  58:16:3331
       (1994).
53. Tunturi, P. & Ylasaarl, S.  A Special Case of the Pitting Corrosion of Copper in a Hot Water
       System.  Proc. 5th Scandanavian Corrosion Conference, Copenhagen, 1968.
54. Wilson,  D.  E.  "An Equilibrium Model  Describing the  Influnce of Humic Materials on the
       Speciation of Cu2*, Zn2+, and Mn2+ in Freshwaters". Limnol Oceanogr., 23:3:499 (1978).
55. Cabaniss, S. E. & Shuman, M. S. "Copper Binding by Dissolved Organic Matter: I. Suwannee
       River Fulvic Acid Equilibria". Geochimica Cosmochimica Acta, 52:185 (1988).
56. Cabaniss, S. E.  & Shuman, M. S. "Copper Binding by Dissolved Organic Matter: II. Variation
    =   in Type and Source of Organic Matter". Geochimica Cosmochimica Acta, 52:195 (1988).
57. Davis, J. A. "Complexation of Trace Metals by Adsorbed Organic Matter".  Geochimica
       Cosmochimica Acta, 48:679 (1984).
58. Hirose, K. "Chemical Speciation of Trace Metals in Seawater: Implication of Paniculate Trace
       Metals". Marine Chemistry, 28:267 (1990).
59. Holm, T. R. "Copper Complexation by Natural Organic Matter in Contaminated and
Uncontaminated Ground Water". Chemical Speciation andBioavailability, 2:63 (1990).
60. Holm, T. R. & Curtiss, C. D., III. Copper Complexation by Natural Organic Matter in Ground
       Water. Chemical Modeling of Aqueous Systems II, Los Angeles, CA, 1990.
61. Fisher, N. S. & Fabris,  J. G. "Complexation of Cu, Zn, and Cd by Metabolites Excreted from
       Marine Diatoms". Marine Chemistry, 11:245(1982).
62. Zhou, X., etal. "Production of Copper- Complexing Organic Ligands During a Diatom Bloom:
       Tower Tank and Batch-Culture Experiments". Marine Chemistry, 27:19 (1989).
63. Brown, F. R. & Fernando, Q. "Kinetics of the Dissolution of Copper Metal in Aqueous Solutions
       Containing Unsaturated Organic Ligands and Copper(II)".  Talanta, 38:3:309 (1991).
64. Edwards, M., Ferguson, J. F. & Reiber, S. H. "On the Pitting corrosion of Copper". Journal of
       the American Water Works Association, 86:7:74 (1994).
65. Campbell, H. S. "Corrosion, Water Composition and Water Treatment".  Wat. Treat. Exam.,
       20:1:11(1971).
66. Campbell, H. S. & Turner, M. E. D. "The Influence of Trace Organics on Scale Formation and
       Corrosion". Journal of the Institute of Water Engineers and Scientists, 37:1:55 (1983).
67. Harrison, R. M. & Laxen, D. P. H. "Physicochemical Speciation of Lead in Drinking Water".
       Nature, 286:5775:791 (1980).
68. AWWARF. Internal Corrosion of Water Distribution Systems. AWWA Research
       Foundation/DVGW Forschungsstelle, Denver, CO, (Second Ed. ed., 1995).
69. Samuels, E. R. & Meranger, J. C. "Preliminary Studies on the Leaching of Some Trace Metals
       from Kitchen Faucets".  Water Research, 18:1:75 (1984).
70. Thresh, J. C. "Action of Natural Waters on Lead". Analyst, XLVTI:560:459 (1922).
71. Miles, G. "Action of Natural Waters on Lead". Journal of the Society of Chemical Industry,
       67:1:10(1948).
72. Campbell, H. S. "The Influence of the Composition of Supply Waters, and Especially of Traces
       of Natural Inhibitor, on Pitting Corrosion of Copper Water Pipes".  Proc. Soc. Water Treat.
       &Exam., 8': 100 (1954).

-------
73. Saar, R. A. & Weber, J. H. "Lead(II)-Fulvic Acid Complexes. Conditional Stability Constants,
       Solubility, and Implications for Lead(n) Mobility". Environmental Science and Technology
       14:7:877(1980).
74.  Dalang, F., Buffle, J.  & Haerdl, W. "Study of the Influence  of Fulvic Substances on the
       Adsorption of Copper(II) Ions at the Kaolmite Surface". Environ. Sci. Technol, 18:3:135
       (1984).
75.  Hering, J. G. & Morel, F. M. M. "Slow Coordination Reactions in Seawater".  Geochimica
       Cosmochimica Acta, 53:611 (1989).
76.  Snodgrass,  W. J., Clark, M. M. &  O'Melia, C. R. "Particle Formation and Growth in Dilute
       Aluminum(III) Solutions". Water Research, 18:4:479(1984).

-------
c
1
6
5
S
T* 3
•la
••S
•£
o
trt
a.
CQ
Q
g|
U
H
"to
-X
"5
ft> G
•c: G
^ «
•S

««
2
s



« *
•K '
3>
"S. e
S g-2
S? •§ n
•o 8 "S
c o &!
*s . o
« Q
o
&>
S
a
&
C.
u
/*
i" r

XX

X XXX

X XX

X XXX


X XXX

X XXX
X XXX
X XXX

X XXXX
X XX



"
XX XXXX







•o
CO _M fl^
••3 ^ ^j Jj a
« ° 13 ^ JS "2
C CM CQ ^3 "y Si

fe o o o o o "g

?>>>>> s
c2 (3 6 6 6 c3 o"
- 

-------
                OO — < (N •<*• o
                28 -: 2 o "»
                *! rr-. "H — t o\
                O   0    <   \0 O
                   • o

         5

         5
               op! 82
               O  '• O *~"
                   V
              o \q £i -* o

              o pJ S 2 ^
              00   —
                    O O\ \O <5 •  O\ T—i
                    Sj ^ c4 o   °i ^
32
                                                 o\
                                                 ov
                                                 o
               i^go§
               f^i   C3

               eo

              U '

              c3
               CO
               S
                 -
              .S  |

              ll §
              < CJ 2
       -S
       8  E

       II
       S  o
S

af
en
O
TJ
W  _
e 13
                                     oj
                                .S IH
                             •^s C  C

o  & < O O 33 5 P   -g.^ 5
                                                           1-S |   -a £

                                                           CQ '^g J2   *•* 2J

-------
                                                   EC
                                                   O
a>


2
o
                  J5
                  *><

                  'i

                  "S
 CO



"S
 d>
 tn


•8
 CJ
                                                                         
                                            O

                                           S
                                                                                              I
                                                                         S
                                                                        "o
                                                  •g

-------
0.7
0.6
0.5
£ 0.4
c
1 0.3
0.2
0.1
0.0
-
-





8.2
i




7.4



i







7.5








8.0





6.6


i


6.9 7.8
8.0




8.1




              Raw   CS   Fl  CW1  OS   F2  CW2  F3' CW3

                                 Location
Figure 2. Aluminum residual at stages of pilot plant operation using
East Fprk Lake water (CS- conventional settled, CW- clearwell, OS-
optimized settled, F- filter). The average PH values are shown above bar

-------
0.4
0.3
0.2
0.1
0.0
            90
   6.0        6.5        7.0        7.5        8.0         8.5

                               pH
Figure 3.  Residual aluminum after filtration as a
function of pH in the filter influent. Aluminum concentrations
represent all filters.

-------
W>
£
    100.0000
     10.0000
      1.0000
     0.1000
     0.0100
     0.0010
L /
 /

                                 	 pH=6.0
                                 	pH=7.0
                                 	pH=8.0
                                 —— pH=9.0
                                 •	 pH=iO.O
            0      25
              50     75      100    125
                mgC/LDIC
                                                        150
     Figure 4. Solubility of copper(II) in response to pH and DIG
     assuming equilibrium with Cu(OH)2 solid at 25° C and 1=0.02.

-------
  100.0000
   10.0000
    1.0000
    0.1000
    0.0100
    0.0010
    0.0001
                                 pH = 6.0
                           -- pH = 7.0
                           --- pH = 8.0
                           ---- pH = 9.0
                                 pH=10.0

              /
              /  ..-••
             I?. .-••'
            0
25
50      75      100
   mg C/L DIG
125      150
Figure 5. Solubility of copper(TT) in response to pH and DIG, assuming equilibrium with
Cu2(OH)2CO3 and CuO solid at 25° C and 1=0.02. This would represent a case of aged plumbing.

-------
    100.0000
     10.0000
                                              pH = 6.0
                                        -- -  pH = 8.0

                                        ----  pH = 9.0
                                                = 10.0
W)
£
1.0000 r
      0.1000  r
0.0100  ' »  « « t  t
                             i  . . .  . t
       0
                    25
50      75      100     125     150

  mg C/L DIG
 Figure 6. Solubility of lead(II) in response to pH and DIG, assuming equilibrium with
 PbC03 or Pb3(OH)2(C03)2 solids at 25° C and 1=0.02.

-------
     10.0000
W)
S
       1.0000
      0.1000
      0.0100
      0.0010
Theoretical Solubility
— Run 1 (DIC=14,pH=8.5)
- Run 2 (DIC=13,pH=7.0)
   Run 5 (DIC=18, pH=7.5)
                   r      •     ,  , n  ,
                  Experimental Data
                           Run 1
                           Run 2
                      O
  «   *	
                                        8
                                      pH
                             10
 Figure 7. Comparison of equilibrium theoretical and observed 24-hour stagnation copper levels for
 coupon study, showing large deviation from expected value for system with high pH and high sulfate
 Thick line is predicted "langite" solubility.

-------
    1.1; AY
       :tt-'H
                    KIV
Journal of the Institution of
WATER AND
ENVIRONMENTAL
MANAGEMENT



           Vol &No 3

-------
                                                                            bby ,'172
                                                                                      P. 13
                                                   If)
                                                   O
                                       r
                                      .A
                                       ^
                                       O

                                       >
                                             _    r
                                             -JO    £
                                                  P

                                                  r

1
0 O. t— t- ct B
ftilS M O 3" C
•O O O 0 >1
fT >- 3 "O >-
CTO .Q S
g 3 ° »«•«•
5 ai jr o
I 3 3 C 0
1— ft _.«•"•
O H rt Q*
& O (S %
g X-o r»3
3 C, «• p. "O
C 81 H* w
IT rr D" 3 3 O
w O to Q

ire taken out from
:alled in a closed
:orrosion rate and
ermined by oxygen
latizod monitoring
the closed loop.
Because of the lack of sufficient
the connections between different
corrosion inhibition effects of the
tests have to be carried out to
for optimization.
Test methods
To obtain correct data, tests have
where the original water quality i
H/4 ) and pipe coupons are install
as a throughflow system with the
inhibitor dosage at a flow, veloc
(tutbulent flow) for several months
Ss^s
^ H* 8>
*> * < a-
o j~5 £.«
M» 3 w* ••"
a t-tj
'i:*l*
" » 2 • 8
i°sJ
3 n a,a a.
0 ® C 1* 0
- o "> 3
tJ^C-D
01 >— OJ
3 rr f-13 H-
\ 3" •-• ft rr
B (8 >< B (S
l&^
ESSS
S «a H-
c informations on
qualities and the
e chemicals, field
i the needed data
The use of inhibitor chemicals i
and potassium salts of phosphoric
additionally limited to 2,2 ppm P
cases an optimazation of the toti
dosage chemicals is wishable with
costs and a minimization of phosphi
to reduce ecological impact, mainl
tion of lakes and rivers.
2 5 * o "
8 ™ "2 = o- ?• n
2 ™ 0 0»»
= ja o. o >- a
S g g-ogJl
SSS55&S
s"|^s&
?•"«"?-
^s^:°
•o o »> ir-1 s o
o -o n£ B o
•- O 3 ^ 3 °- 0.
l— to IB B> 1—
c a s o s H- e
l D* ri- »^K; w 3
On the other hand, red water pro
such systems and are the reason
customers (1-3). Beside that, it
water is limited in Germany to 0,2
where problems arise, the use of i
to reduce or even resolve the pr
range while relining or replacement
o o o er
& O~ 3"O 3 rr, 1—
. T !-• H'^ O m
(t o O^ 3 n 3
31— a IB
^- m ri" T O •>
often arise from
complaints of the
ntent in drinking
e. In theee cases,
ors can be helpful
i in a short time
ong terra measures.
Introduction
Drinking water distribution syst
the older one, are often made fi
ductile iron without a sufficient
tection. Despite that, damages caus
attack which leads to a perforatio
ly occuring and can be linked d
to unqualified pipe-laying with re
no sufficient flushing or unfavo
tions.
c to M. a 5 x »
•i >-* §.*"§ 3
o»a.roo3*co
C* c O rn rr rf
t-* n r* .2 c^ w
iBtn^-n-^^rf^"-
^ 5 i~= IB S
o--* <*$?.«.
Us?.R»r5
2s- -o 3 o o 3
f* »~ 2 ° n ° f
>--rrS-Up-gw =
§s ^oi^^
n> a- n o J-" ^
» ° •< 2 IT
o •o -a n s 3 3
o K-'O no 2 b
3*DIBC»W'n-3h--
O. 15 3 ••( W- K 3
>— tnuroooo^-
i - - i 3 i n *<

-------
VB>
\       UNITED STATES ENVIRONMENTAL PROTECTION AGENCY
           NATIONAL RISK MANAGEMENT RESEARCH LABORATORY
                      CINCINNATI. OH 45268

                    December 19,  1996
                                                            OFFICE OF
                                                      RESEARCH AND DEVELOPMENT
 MEMORANDUM            .      .                               '

 SUBJECT:  Seasonal Monitoring Revision

 FROM:     Michael R. Schock, ChemistfM£
           Treatment Technology Evaluation Branch
           Water Supply and Water Resources  Division

 TO:       Jeffrey B. Keropic, Environmental  Engineer
           Office of Ground Water and Drinking Water


      Pursuant to our prior correspondence and discussion about.
 removing requirements that samples be collected  in  "warmest"
 months/ I have collected a small set of papers for  your
 information and justification of the change.   I  agree  that in
 some water systems the highest lead, copper,  or  both levels  could
 be highest in the warmest months.  However, I do not think that
 the presumption that it is always the case  is supportable by
 scientific evidence.  There are many cases  where exactly the
 opposite is true, as evidenced by both fundamental  chemistry
 arguments as well as data from experiments  and water distribution
 system sampling.  Therefore, when possible, I believe  that the
 Lead and Copper Rule should be revised to allow  sampling
 flexibility to account for this phenomenon.   Of  course,  for  water
 systems required to collect two or more rounds of monitoring data
 that can encompass "seasonality," there is  no reason to make any
 change in -requirements.

      Language changes may be required in some places to modify
 statements that would imply guidance towards  biasing of
 collection times to the "warmest" months.   In such  cases,  phrases
 could simply be amended to suggest biasing  towards  seasons or
 months "where the highest levels are most likely to occur."

      For seasonally-operated water systems, a secondary
 supplemental argument is that the intent of the  Rule was to
 reduce major exposures of sensitive populations  to  lead and
 copper.  Therefore,  sampling is most appropriate when  the systems
 are actually being used.  The question then becomes when to
 sample if operation encompasses a range of months covering a
          Recycled/Recyclable • Printed wBh Vegetable OI Based Inks on 100% Recycled Paper (40% Postconsumer)

-------
                    >

range of temperatures.  There is no definitive argument on this,
because currently we can rarely predict beforehand whether a
given water chemistry coupled with physical factors, will cause
the highest values at a particular time.  Therefore, I do not
think we should attempt to be too precise in proscribing exactly
when to take the samples in a confining regulatory structure.

     As noted previously, I think the strongest line of argument
we should use is to say that the uncertainty of the time of
^worst" metal levels causes EPA to refrain from specifying
particular months for sampling across the country.  If monitoring
data from similar systems or prior monitoring or survey
experience in that particular system is available to the States,
and if the States wish to use that information to make particular
time requirements for the systems, then I think that would be
reasonable.

     There are several plausible ^theories" about why metal
levels could frequently be higher in the "winter" months, each or
some combination of which could operate in a given water system.
Some of these are:

•    The intrinsic net solubility of many minerals, especially
     carbonates, goes up as the temperature goes down.

•    Corrosion inhibitors, especially orthophosphate, may react
     more slowly at lower temperature, so passivating film
     formation is less effective in colder water.

•    Corrosion inhibitors and other treatment chemicals may be
     more viscous at lower temperatures.  Therefore, the chemical
     feed rates may be lower when cold.

•    Many pipes are near heating systems, and in the winter the
     operation of the heating systems causes the pipes to be
     hotter.  Plus, the change in temperature could, also disrupt
     the existing protective films in the pipes built up over the
     earlier months of more stable temperatures  (this is an
     argument advanced by several staff members of Minnesota
     DOH) .

•    Dissolved oxygen levels are often higher in colder waters,
     resulting in enhanced concentrations of primary oxidants in
     the water  (when added to chlorine species).  This causes
     more rapid increases in metal levels through enhanced
     oxidation during short standing-times  (6-16 hours).

     The information I have attached is as follows, with brief
descriptions of their significance.

-------
i^..   Colling,  et.  al.  Jour.  JWEW (1992) paper showing field
 data where lower lead levels are obtained at higher temperatures
 in  some orthophosphate-dosed hard water systems in the UK (page
 262,  Fig.. 3) .

 -2.   Edwards/Schock/Meyer Jour.  AWNA (1996)  paper showing
 theoretical arguments why copper levels should be higher at lower
 temperatures in many cases.  This paper cites some data from
 Boulder {not published)  and utility data (see next one) .

^3.   Dodrill/Edwards Jour.  AWWA  (1995)  paper (pp.  79-80) showing
 temperature trends not readily apparent in major utility
 monitoring data.  There was a suggestion of  a trend of lower
 levels at higher temperatures especially for lower alkalinities.
 This paper brings up another interesting point that because of
 the different structures of houses and where the first-draw liter
 of  water res-ides in the plumbing, major temperature trends should
 not be expected.  Some unpublished work Chet Neff and I did many
 years ago (which could be reproduced with the assistance of some
 standard heat-transfer equations) suggested  that waters in lead
 and copper pipes quickly  (minutes to only an hour or two) take on
 the temperatures of their surroundings.

^#.   Rezania/Anderl conference paper (1996)on observations of
 higher Cu levels in winter in Minnesota sampling (page 4) .  They
 have several arguments to explain this, with some in the paper,
 and some noted above in my discussion.

 5.   Some supplemental laboratory test data by doctoral student
 Loay Hidmi, University of Colorado, Dept. of Civil Engineering.
 This shows higher copper release at pH 7 for lower temperature in
 water of same alkalinity.  Compare Cu level  at most times from
 Chart 1 to Chart 4.  This may not be citeable.

 6.   Some data from a couple of water systems in MN collected by
 U.   Colorado.  Note highest Cu levels are usually associated with
 "winter" month sampling.   (Probably not citeable)

 7.   Some other experimental data from Marc Edwards7 group at U..
 Colorado  (currently confidential but possibly citeable with
 permission from Marc) showing soluble copper being higher in cold
 water loops than in hot water loops.

      Interestingly, a review of  the original preamble  (Page
 26524-26525) shows that EPA had  information showing field data
 was equivocal about any particular month or period having higher
 or lower metal levels.  The choice was apparently made to
 essentially "weight" the  evidence from some unspecified  studies
 more highly than others,  and the proposal to use July  through
 September was kept  (after adding June).  I believe, therefore,

-------

that this proposal is not really particularly new, but does help
argue that the increased inconvenience of requiring sampling in
specific months does not pay off in tangible public health
protection.  It also provides an option for States that have good
reason to believe a particular time period would be preferable
for sampling to use their discretion to designate such a time
constraint.

     If you need additional research to find other articles, or
if you would like me to help with some preamble language, let me
know.

Attachments

cc:  Judy Lebowich
                                                                V

-------