Anda di halaman 1dari 215

COPPER:

ENVIRONMENTAL DYNAMICS
AND HUMAN EXPOSURE ISSUES

Prepared for:
The International Copper Association

By:
P. G. Georgopoulos*, A. Roy*, M. J. Yonone-Lioy +, R. E. Opiekun*, and P. J. Lioy*

*Environmental and Occupational Health Sciences Institute (EOHSI)


Piscataway, NJ 08854
+
Nu Horizon Enterprises, Inc.
Cranford, NJ 07016

March 2001

EOHSI is a joint project of

UMDNJ-Robert Wood Johnson Medical School and Rutgers The State University of New Jersey
COPPER:
ENVIRONMENTAL DYNAMICS AND HUMAN EXPOSURE ISSUES

Table of Contents

1 OBJECTIVE AND OVERVIEW...................................................................................... 1


1.1 Objective .........................................................................................................................1
1.1.1 What This Document Contains ...................................................................................... 5
1.1.2 What Should Come Next................................................................................................ 6
1.2 Description of the Relational Database Management System for Copper Environmental
Distribution and Exposure Studies (RDMS-CEDES) ......................................................7
1.2.1 Using the RDMS-CEDES CD-ROM ............................................................................... 8

2 INTRODUCTION: COPPER AND MAN ...................................................................... 15


2.1 Copper and Technology: From the Copper Age to the Information Age .......................15
2.2 Copper and Biology: Essentiality and Toxicity ..............................................................15
2.2.1 Health Effects in Humans............................................................................................. 16
2.3 Copper and Policy.........................................................................................................19
2.3.1 Guidelines and Recommendations for Copper Intake ................................................. 20
2.3.2 Notes on Recent Regulatory Activities ......................................................................... 21

3 A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL


PROPERTIES OF COPPER ........................................................................................ 27
3.1 Physical and Chemical Attributes of Copper .................................................................27
3.2 Biological Functions of Copper .....................................................................................29
3.2.1 Biochemical Functions ................................................................................................. 29
3.2.2 Physiologic Functions .................................................................................................. 32

4 ENVIRONMENTAL RELEASES OF COPPER ........................................................... 39


4.1 Atmospheric Releases ..................................................................................................39
4.2 Releases to Wastewater ...............................................................................................40
4.2.1 Copper in Cooling Systems.......................................................................................... 42
4.2.2 Releases to Land ......................................................................................................... 43
4.2.3 Databases for Environmental Copper Releases .......................................................... 43

5 ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER.............. 49


5.1 General Concepts .........................................................................................................49
5.2 Atmospheric Dynamics of Copper.................................................................................51
5.3 Hydrospheric Dynamics of Copper ...............................................................................52

COPPER: Environmental Dynamics and Human Exposure Issues


Page iii
5.3.1 General Discussion ...................................................................................................... 52
5.3.2 Fluxes of Copper in the Hydrosphere .......................................................................... 53
5.3.3 Chemistry of Copper in the Hydrosphere..................................................................... 57
5.3.4 Drainage from Urban Areas ......................................................................................... 60
5.3.5 Copper in Groundwater................................................................................................ 61
5.3.6 Mathematical Modeling of Metal Speciation................................................................. 61
5.3.7 Copper in Seawater ..................................................................................................... 63
5.3.8 Copper in Sediments.................................................................................................... 63
5.4 Copper in the Lithosphere and Pedosphere .................................................................66
5.4.1 Soil (Pedospheric) Dynamics of Copper and Soil-Water Interactions .......................... 67
5.5 Environment-Biota Interactions: Bioconcentration and Bioaccumulation ......................70
5.5.1 Aquatic Biota ................................................................................................................ 70
5.5.2 Terrestrial Biota............................................................................................................ 73
5.5.3 Copper in the Anthroposphere ..................................................................................... 73

6 COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES: OVERVIEW OF


SELECTED FIELD STUDIES ...................................................................................... 81
6.1 Atmospheric Concentrations .........................................................................................81
6.1.1 Precipitation ................................................................................................................. 84
6.1.2 Fog ............................................................................................................................... 84
6.2 Copper Concentrations in the Hydrosphere..................................................................84
6.2.1 Marine Water................................................................................................................ 85
6.2.2 Estuarine Water ........................................................................................................... 87
6.2.3 River Water .................................................................................................................. 87
6.2.4 Lake Water................................................................................................................... 89
6.2.5 Groundwater ................................................................................................................ 90
6.2.6 Sediments .................................................................................................................... 91
6.3 Copper Concentrations in Soils and Terrestrial Biota .................................................100
6.3.1 Copper at Hazardous Waste Sites............................................................................. 104
6.4 Copper Levels in Indicator Biota .................................................................................104
6.4.1 Sewage Sludge .......................................................................................................... 107

7 ASSESSING HUMAN EXPOSURE, DOSE AND RISK ............................................ 109


7.1 Exposure Potential and Pathways ..............................................................................109
7.2 Environmental Exposures ...........................................................................................109
7.2.1 Inhalation.................................................................................................................... 109
7.2.2 Ingestion..................................................................................................................... 110
7.2.3 Dermal........................................................................................................................ 110
7.3 Occupational Exposures .............................................................................................110

COPPER: Environmental Dynamics and Human Exposure Issues


Page iv
7.4 Dietary Exposures to Copper: Drinking Water and Food ............................................112
7.4.1 Copper in Food .......................................................................................................... 112
7.4.2 Copper in Drinking Water........................................................................................... 115
7.5 From Exposure to Dose: Bioavailability of Dietary Copper .........................................122
7.5.1 Interactions of Copper with Other Components of the Diet........................................ 122
7.5.2 Copper Biokinetics and Metabolism........................................................................... 124
7.5.3 A Note on Unusually Susceptible Populations (Based on ATSDR, 1990) ................. 127
7.6 Modeling Health Risks Due to Copper Exposures ......................................................128

8 BIOMARKERS OF COPPER EXPOSURE AND EFFECTS ..................................... 135


8.1 Biochemical Indices of Copper Status ........................................................................135
8.1.1 Biomarkers of Exposure Susceptibility and Effect...................................................... 135
8.2 Copper Biomarkers Used in Recent Human Health Studies.......................................138
8.3 Copper Biomarkers and Exposure Markers in Populations ........................................141
8.3.1 Copper Levels in Human Hair .................................................................................... 142
8.3.2 Copper Levels in Blood and Serum ........................................................................... 143

9 CONCLUSIONS AND RECOMMENDATIONS FOR FUTURE RESEARCH ............ 149

10 APPENDIX A: A COMPILATION OF COPPER DATA IN FOOD, DRINKING WATER,


AND AIR .................................................................................................................... 151

11 APPENDIX B: A COMPILATION OF COPPER DATA SOURCES .......................... 159

12 APPENDIX C: ACRONYMS, ABBREVIATIONS AND GLOSSARY OF TERMS..... 167

13 REFERENCES........................................................................................................... 173

COPPER: Environmental Dynamics and Human Exposure Issues


Page v
List of Tables

Table 1: International Regulatory Practices for Copper in Drinking Water .................................................24


Table 2: Copper Containing Proteins Found in Humans.............................................................................34
Table 3: Selected Industrial Copper Compounds and their Properties.......................................................35
Table 4: Copper Releases to Water and Land, 1987 to 1993 (in pounds)..................................................44
Table 5: Summary of Copper Concentrations in Environmental Media as Reported in the Copper
Sourcebook 1998 (Harrison, 1998)..............................................................................................75
Table 6: Copper Concentrations in Environmental Media and Biota as Reported in The Handbook of
Trace Elements (Pais & Benton Jones, 1997).............................................................................76
Table 7: Copper Concentrations in Environmental Media and Biota as Reported in Nriagu (Nriagu, 1979b)
.....................................................................................................................................................77
Table 8: Sediment Component Classes......................................................................................................78
Table 9: Tissue and Body Copper Levels in Healthy Adults and Adults with Wilson's Disease ...............145
Table 10: Copper Content of Human Tissues and Body Fluids ................................................................146
Table 11: Concentrations of Copper in Air ................................................................................................151
Table 12: Concentrations of Copper in Water ...........................................................................................153
Table 13: Copper Content of Selected Foods ...........................................................................................154
Table 14: Copper Content of Selected Foods per 100 Grams..................................................................156

COPPER: Environmental Dynamics and Human Exposure Issues


Page vi
List of Figures

Figure 1: Main Switchboard of RDMS-CEDES ...........................................................................................10


Figure 2: Reports Switchboard of RDMS-CEDES ......................................................................................11
Figure 3: Criteria for Selection of Articles in RDMS-CEDES.......................................................................12
Figure 4: Forms Switchboard of RDMS-CEDES.........................................................................................13
Figure 5: Sample Report from RDMS-CEDES Showing Descriptors and Data from Selected Articles......14
Figure 6: Locations of Copper Mines in the U.S. ........................................................................................45
Figure 7: Locations of Environmental Release of Copper in the United States ..........................................46
Figure 8: Environmental Releases of Copper in the United States (by County). ........................................47
Figure 9: Conceptual Multilevel-Multiscale Model of Environmental and Human Exposure Dynamics of
Copper .........................................................................................................................................79
Figure 10: Prototype Implementation of a Model of Global Distribution and Fluxes of Copper in the
Environment, Using the STELLA (Structured Experimental Learning Laboratory with Animation)
Simulation Software .....................................................................................................................80

COPPER: Environmental Dynamics and Human Exposure Issues


Page vii
Acknowledgments

The authors wish to thank the International Copper Association, NYC, NY (#TPT0569A and B-
99) for funding this review, and in particular Dr. Scott Baker of ICA for his insights and
encouragement. The authors also wish to thank Dr. Gustavo Lagos and Dr. Herbert Allen for
reviewing an early version of the manuscript and providing most valuable comments and
suggestions. Finally, appreciation is extended to Ms. Linda Everett for assisting with the
literature database and the preparation of the final version of this manuscript, to Mr. Ioannis
Georgopoulos and Dr. Vikram Vyas for providing useful information on publications and
databases relevant to this effort, and to Mr. Michael Hennelly of ICA for his editorial comments.

COPPER: Environmental Dynamics and Human Exposure Issues


Page viii
Chapter 1
OBJECTIVES AND OVERVIEW

1 OBJECTIVE AND OVERVIEW


A DISCUSSION OF THE OBJECTIVES AND STRUCTURE OF THE PRESENT
MONOGRAPH AND THE ACCOMPANYING RELATIONAL DATABASE
MANAGEMENT SYSTEM

1.1 Objective
The objective of this work is to provide a selective review and evaluation of currently available
information on environmental and biological chemodynamics (i.e. physicochemical
transformations and transport dynamics) of copper and of its levels and distribution (observed or
estimated) in environmental media and human tissues.

An essential component of this evaluation effort has been the development and testing of an
appropriate structure for the flexible organization and management of the above information.
Such a structure has been implemented computationally in the form of a prototype, user-
oriented, relational database, named Relational Database Management System for Copper
Environmental Distribution and Exposure Studies (RDMS-CEDES). This easily expandable
and upgradable prototype database (available in MS Access format) provides access to data
summaries, literature references and other information relevant to studies focusing on copper
levels in environmental media and human tissues, and on the processes affecting these levels.

This monograph accompanies and supplements the above prototype relational database,
providing both (a) an overall background of issues relevant to copper release and distribution in
the environment and to associated potential human exposures, and (b) overview, summaries and
discussion of the specific information contained in RDMS-CEDES.

The selective review attempted in this work provides a basis for developing an evolving
conceptual and eventually quantitative model, integrating both phenomenological and
mechanistic aspects of copper dynamics and environmental exposure potential. It is also
expected to provide a scientific basis for the design of future research to fill identified gaps in
data, specifically needed to test hypotheses that can establish source-to-human receptor
relationships.

The above mentioned model of copper dynamics would provide a tool for identifying and
evaluating the influence of

biogeochemical processes, and

human activities

on the distribution and accumulation of copper in different environmental media.

This tool can be expected to enable establishment of a rational basis for quantifying estimates of
exposure and dose to human populations, to specific individual cases of concern, and to other
relevant components of ecosystems.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 1
Chapter 1
OBJECTIVES AND OVERVIEW

It should be noted that clearly it is not the objective of this work to provide an exhaustive review
of available data on copper in the environment, nor to review the biological impact and effects of
copper on humans and on terrestrial and aquatic ecosystems. The Copper Research Information
Flow (CRIF) Project maintains the International Copper Association (ICA) Reference
Collection - the most extensive and current collection of published materials on copper as it
relates to human health and the environment. CRIF is supported by the International Copper
Association and based in the Department of Earth and Ocean Sciences at the University of
British Columbia (UBC) in Vancouver, Canada. The Project now contains over 74,000 citations
on copper, provides access to major online databases (e.g., Chemical Abstracts, Toxline,
Medline, Biosis, NTIS, Embase) and 5,000 e-journals. The database is updated and augmented
regularly to ensure currency. Led by Dr. Brenda Harrison at UBC, the service can be accessed
by contacting Dr. Harrison at brendaharrison@shaw.ca. Readers interested in an introduction to
the ecological role and effects of copper can consult excellent recent reviews on these subjects
(Lewis, 1995a; Lewis, 1995b; Ecological Planning and Toxicology Inc., 1998). An introduction
to issues associated with copper toxicity and related risks, covering literature sources up to the
late 1980s, is provided in the 1990 ATSDR Toxicological Profile for Copper (TP-90-08). Good
summary reviews of the role of copper as a micronutrient in the human diet are available (see
e.g. Turnlund, 1999; Bogden & Klevay, 2000; WHO-IPCS, 1998; National Institute of Medicine,
2001).

Additional useful information can also be found in these references:

Adriano, 1986, Trace Elements in the Terrestrial Environment. 533 pp. Springer-Verlag,
NY. (Chapter 6. Copper. pp. 181-218.)

Howell & Gawthorne, 1987, Copper in Animals and Man. Vol. A. 125 pp. Vol B. 140 pp.
CRC Press.

Kies, 1989, Copper Bioavailability and Metabolism. Vol. 258 in Adv. Experimental
Medicine and Biology. 307 pp. Plenum Press, NY.

Linder, 1991a, Biochemistry of Copper. 525 pp. Plenum Press, NY.

Linder, 1991b, Nutrition and metabolism of the trace elements. in Nutritional


Biochemistry and Metabolism: With Clinical Applications. Elsevier Science Publishing,
Amsterdam, The Netherlands.

Finally, since 1997, a number of new documents have been published which contain useful
review/summary information on copper:

He et al., 1997, Spatial and temporal patterns of acidity and heavy metals in predicting
the potential for ecological impact on the Le An river polluted by acid mine drainage. Sci.
Total Environ.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 2
Chapter 1
OBJECTIVES AND OVERVIEW

Kennish, 1997, Practical Handbook of Estuarine and Marine Pollution. 524 pp. CRC
Press, Boca Raton.

O'Dell & Sunde, 1997, Handbook of Nutritionally Essential Mineral Elements. 692 pp.
Marcel Dekker, NY. (Chapter 8, Copper by E.D. Harris. pp. 231274.)

Pais & Benton Jones, 1997, The Handbook of Trace Elements. 223 pp. St. Lucie Press.

Richardson, 1997, Handbook of Copper Compound and Applications. 432 pp. Marcel
Dekker, NY.

Riveros-Rosas et al., 1997, Personal exposure to elements in Mexico City air. Sci. Total
Environ. 198:79-96.

Tobias et al., 1997, Establishment of the background levels of some trace elements in
soils of NE Spain with probability plots. Sci. Total Environ. 206:255-265.

WHO-IPCS, 1998, Copper - Environmental Health Criteria 200. World Health


Organization and the International Programme on Chemical Safety

Budd et al., 1999, The Keweenaw current and ice rafting: Use of satellite imagery to
investigate copper-rich particle dispersal. J. Great Lakes Research 25 (4):642-662.

Joseph, 1999, Copper: Its Trade, Manufacture, Use, and Environmental Status. 451 pp.
ASTM International.

Kerfoot & Nriagu, 1999, Copper mining, copper cycling and mercury in the Lake
Superior ecosystem: An introduction. J. Great Lakes Research 25 (4):594-598.

Kerfoot et al., 1999, Anthropogenic copper inventories and mercury profiles from Lake
Superior: Evidence for mining impacts. J. Great Lakes Research 25 (4):663-682.

Kerfoot & Robbins, 1999, Nearshore regions of Lake Superior: Multi-element signatures
of mining discharges and a test of Pb-210 deposition under conditions of variable
sediment mass flux. J. Great Lakes Research 25 (4):697-720.

Kolak et al., 1999, Nearshore versus offshore copper loading in Lake Superior sediments:
Implications for transport and cycling. J. Great Lakes Research 25 (4):611-624.

Landner & Lindestrom, 1999, Copper in Society and in the Environment: An Account of
the Facts on Fluxes, Amounts and Effects of Copper in Sweden. Second revised edition.
329 pp. Swedish Environmental Research Group.

Leone & Mercer, 1999, Copper Transport and its Disorders: Molecular and Cellular
Aspects. Vol. 448. New York, NY: Kluwer Academic/Plenum Publishers.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 3
Chapter 1
OBJECTIVES AND OVERVIEW

Mansilla-Rivera & Nriagu, 1999, Copper chemistry in freshwater ecosystems: an


overview. J. Great Lakes Research 25 (4):599-610.

DiToro et al., 2000, The Biotic Ligand Model, Copper in the Environment and Health.
New York, NY: International Copper Association.

Eisler, 2000, Handbook of Chemical Risk Assessment: Health Hazards to Humans,


Plants, and Animals. Vol. 1. Metals. 738 pp. Lewis Publishers. Boca Raton. (Chapter 3:
Copper. pp. 93-200.)

Jeong et al., 2000, Release of Copper from Mine Tailings on the Keweenaw Peninsula. J.
Great Lakes Research 25 (4):721-734.

Landner et al., 2000, Copper in Sewage Sludge and Soil, Copper in the Environment and
Health. New York, NY: International Copper Association.

Multhaup & Hermann, 2000, Copper in the Pathogenesis of Neurodegenerative


Disorders: A Literarature Summary, Copper in the Environment and Health. New York,
NY: International Copper Association.

National Research Council, 2000, Copper in Drinking Water. 147 pp. National Academy
Press, Washington, D.C.

Samet, 2000, A Technical Guide for the Study of Acute Gastrointestinal Effects of
Copper in Drinking Water: Methods for Public Health Investigators, Copper in the
Environment and Health. New York, NY: International Copper Association.

Georgopoulos et al., 2001, Environmental copper: Its dynamics and human exposure
issues. Journal of Toxicology and Environmental Health Part B, 4:341-394.

Lagos, 2001, Corrosion of Copper Plumbing Tubes and the Release of Copper By-
Products to Drinking Water - A Literature Summary, Copper in the Environment and
Health. New York, NY: International Copper Association.

Parametrix Inc. & EPT, 2001, Acclimation and Adaptation of Terrestrial Organisms to
Metals in Soil. New York, NY: International Copper Association.

National Institute of Medicine, 2001, Dietary Reference Intakes for Vitamin A, Vitamin
K, Arsenic, Boron, Chromium, Copper, Iodine, Iron, Manganese, Molybdenum, Nickel,
Silicon, Vanadium, and Zinc (Chapter 7 - Copper). Washington D.C.: Institute of
Medicine, Food and Nutrition Board.

The special issue of American Journal of Clinical Nutrition (AJCN, 1998) should also be
mentioned here. It contains the scientific reports presented and discussed at the International
Conference on Genetic and Environmental Determinants of Copper Metabolism, sponsored by

COPPER: Environmental Dynamics and Human Exposure Issues


Page 4
Chapter 1
OBJECTIVES AND OVERVIEW

the National Institutes of Health (NIH) and the University of Chile, held at the NIH Stone House,
March 18-20, 1996. The purpose of the conference was to critically review genetic and
environmental factors that determine copper metabolism, as well as their implications for copper
deficiency and excess in humans.

Finally, it should be noted that up-to-date information regarding many issues and activities
involving copper (including production, economic, environmental, etc. data) can be accessed
through the web pages, and the numerous links thereof, of the Copper Development Association
(http://www.copper.org ), the International Copper Association (http://www.copperinfo.com) and
the International Copper Study Group (http://www.icsg.org). Also, more recent information on
evolving methods and databases for assessing copper exposures can be found in the forthcoming
report A Framework and Data Sources for the Assessment of Exposures to Copper
(Georgopoulos et al., 2002). More details are provided on the website of the Center for Exposure
and Risk Modeling (CERM), in the section devoted to copper (http://www.CERM.org/copper).

1.1.1 What This Document Contains


In order to provide a conceptual road map to the organization of this monograph, we include a
brief description of the contents of each chapter.

The current chapter sets forth the objectives and organization of this monograph and of
the accompanying relational database management system for copper environmental
distribution and exposure studies (RDMS-CEDES). In addition, it provides up-to-date
information sources unavailable when this document was written, and recommendations
for what needs to be done next.

Chapter 2 discusses the history of copper use and summarizes its human health effects, in
terms of essentiality, deficiency and excess (toxicity). It also discusses worldwide
governmental copper regulation/policy/guidelines for pollutant discharge, human
exposure and dietary intake.

Chapter 3 provides a brief review of coppers essential physical, chemical and biological
attributes, as a background for the subsequent discussion of environmental and biological
copper chemodynamics.

Chapter 4 concerns copper production, use and disposal.

Chapter 5 provides a brief overview of environmental and biological copper


chemodynamics, including the atmospheric, hydrospheric, and soil-water interactions, as
well as environment-biota interactions.

Chapter 6 provides an overview of selected field studies on copper levels in the


environment, from household dust to marine sediments, and in human tissue.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 5
Chapter 1
OBJECTIVES AND OVERVIEW

Chapter 7 discusses human exposure, dose, and risk assessment under various conditions
and by various routes, and copper health-risk modeling.

Chapter 8 concerns biomarkers of copper exposure and effects, and

Chapter 9 sets forth conclusions and recommendations for future research.

1.1.2 What Should Come Next


During the development of this report it was realized that a very significant component of
existing data on copper have not been incorporated in the peer-reviewed literature. There are
several reasons for this, the main one being that important information on copper is often
collected through studies not focusing on copper. As a result, data directly related to copper often
do not become the subject of specific analysis and/or do not get published in the peer-reviewed
literature; rather, they can only be found in federal and state agency reports and/or in electronic
data files. To address this problem, a number of databases containing temporal and spatial
information on copper distribution were identified, mostly from such United States federal
agencies such as USGS, the Centers for Disease Control, and US EPA. Such databases are:

USGS Water Quality Monitoring Network (WQN) MT2 Data


USGS National Geochemical Atlas Data
US NOAA Ocean Resources Conservation and Assessment (ORCA) Data Trends and
Status, Mussel Watch and Benthic Surveillance
US EPA Environmental Monitoring Assessment Program (EMAP)
US EPA Toxics Release Inventory
US EPA NHEXAS
US EPA AIRS
US EPA SDWIS
US EPA STORET
CDC NHANES II/III
Geochemical Atlases of Hungary, Slovakia, Sweden, etc.
SWAD (European Surface Water Database)

In most instances, analysis of the data in the above databases has not yet appeared in the peer
reviewed literature.

Appendix B of this document provides a summary description of the contents of these databases.

It is our strong recommendation that an effort be made to retrieve and organize material from
these databases, with an appropriate focus on copper.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 6
Chapter 1
OBJECTIVES AND OVERVIEW

1.2 Description of the Relational Database Management System for Copper


Environmental Distribution and Exposure Studies (RDMS-CEDES)
The prototype RDMS-CEDES was developed using, as its starting source of information, the
Copper Information Sourcebook 1998. The Sourcebook lists 4,746 articles (i.e. the results of the
1997 search selected for inclusion in the ICA Reference Collection, which, as mentioned earlier,
contains more than 50,000 items) spanning a wide array of scientific disciplines, encompassing
human health, ecology, agriculture, the environment, etc. In addition, the Sourcebook also
contains tabular compilations of the gross aspects of the copper monitoring data (e.g., min-max
ranges, etc.) found in the articles cited. Owing to the wide variety of studies cited in the
Sourcebook, not all were of equal value for the purposes of this monograph. Further, there is
considerable variability in the types and quality of reported copper concentration data. As a
consequence of such informational limitations and constraints, it was decided to concentrate on
copper-related data reported within the context of the design of individual studies. The database
provided on the CD-ROM attempts to offer a perspective on the content of a subset of the
literature cited in the Sourcebook that met the criteria (provided in the final draft of the CD-
ROM) of focus, relevance, and study design, needed for inclusion in the present assessment.

More specifically, the articles listed in the 1998 Sourcebook were first screened based upon their
titles, and over 400 articles, or about 8%, were identified as potentially containing information
on environmental copper that met the selection criteria of focus and relevance. Thus, the selected
articles are a subset of those used to compile the tables of environmental monitoring data in the
1998 Sourcebook. These 400 plus articles were then reviewed, and evaluated, and 112 of them
were selected based on the following criteria:

adequate definition of study design;


full characterization of the analytical methods used in the study;
inclusion of adequate summaries; and

explicit discussion of the conclusions derived from the data, especially those which enhance
understanding of spatial and temporal trends in environmental copper distribution.

Salient information from these articles was summarized in RDMS-CEDES. It should be noted
that of the limited number of studies that met the evaluation and analysis criteria, only six were
designed specifically to study copper-related problems or issues. It should be further noted that a
number of the reviewed articles were found to be relevant to more than one media, therefore
supporting the approach of using a relational database as a convenient tool for cross-referencing
information in these articles. The relational data management approach also provides flexibility
in structure that can prove to be advantageous for future alternative classifications of the data.

The articles selected for inclusion in the prototype RDMS-CEDES were characterized according
to:

Focus,
Type of Study,
Geographical Location,

COPPER: Environmental Dynamics and Human Exposure Issues


Page 7
Chapter 1
OBJECTIVES AND OVERVIEW

Setting,
Polluted or Pristine Environment,
Environmental Media Studied,
whether Flux was considered,
Temporal Scale of the Study and

Spatial Scale of the Study.

The Focus field indicates whether the focus of the paper was mainly:

copper (CU),
inorganic ions (II),
toxic heavy metals (Boon, 1994),
organics (O), or

essential trace metals (TM).

The type of study is also denoted in the database by classifying research as:

analytical (A),
statistical (S),
modeling (M),
field research (F),
laboratory-based research (L), or

toxicological study (T).

Location was also used as a category in the database, classifying each study by continent. The
continent identifiers are defined in the Copper Key.txt file located on the CD-ROM. In
addition to geographical information, the database further subdivides each study by

setting (e.g., agricultural, rural, urban, etc.),


condition of the environment (e.g., polluted or pristine),
type of media investigated (e.g., air, plants, water, etc.) and

whether or not environmental flux was considered in the study.

The Spatial Scale and Temporal Scale fields in the database provide an indication of the scope of
the study, and the resolution of data in space and time. All of these categories are fully defined in
the Copper Key.txt file on the CD-ROM.

1.2.1 Using the RDMS-CEDES CD-ROM


The RDMS-CEDS CD-ROM requires the user to have Microsoft Access (97 or 2000) software.
To activate the database, the user should double-click on the file copper_select_frontend. Once

COPPER: Environmental Dynamics and Human Exposure Issues


Page 8
Chapter 1
OBJECTIVES AND OVERVIEW

the database is loaded, the user is presented with a Main Switchboard page (see Figure 1)
through which the database can be accessed and sorted by the criteria mentioned above (see
Figure 3). The Create/Print option allows the user to create reports that are formatted for
printing, while the View Copper Data option allows the user to view copper references and data
interactively.

Selecting the Create/Print Reports option allows the user to


o view and print all of the references selected from the 1998 Copper Information
Sourcebook for this study, and
o search the 112 selected references using criteria available through pull-down menus
on a form. If the user allows an asterisk to remain in place of any given criterion, all
articles for that criterion will be displayed.
Selecting the View Copper Data option allows the user to view selected references
based on the Focus and Environmental Media attributes of the reference. Double-clicking
any of the selected references allows the user to view a summary of the data in the
reference.
Selecting EXIT will allow the user to return to the Main Switchboard.

Please note: All variables are defined in the Copper Key.txt file on the CD-ROM.

The following figures (Figure 1 to Figure 5) show screenshots of the switchboards and sample
forms and reports of the prototype RDMS-CEDES.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 9
Chapter 1
OBJECTIVES AND OVERVIEW

Figure 1: Main Switchboard of RDMS-CEDES


This screen appears when the database is initially opened. Select Create/Print Reports to create
a printable list of references. Select View Copper Data to view the data in one or more of the
selected papers reviewed in this report.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 10
Chapter 1
OBJECTIVES AND OVERVIEW

Figure 2: Reports Switchboard of RDMS-CEDES


This screen appears when Create/Print Reports is selected in the Main Switchboard. Select
Search Selected Articles to create a list of references of the selected papers reviewed in this
study. Select References: Copper Information Sourceboook 1997 to create a list of all
references in the 1997 Sourcebook.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 11
Chapter 1
OBJECTIVES AND OVERVIEW

Figure 3: Criteria for Selection of Articles in RDMS-CEDES


This screen appears when Search Selected Articles is selected in the Report Switchboard
Screen. Create a list of References of the articles reviewed in this report based on criteria chosen
from the drop-down lists. Choosing * in the drop-down list will select all articles for those
criteria.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 12
Chapter 1
OBJECTIVES AND OVERVIEW

Figure 4: Forms Switchboard of RDMS-CEDES


This screen appears when View Copper Data is selected in the Main Switchboard. Select
View Selected Articles and Data to choose articles for which data are to be viewed.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 13
Chapter 1
OBJECTIVES AND OVERVIEW

Figure 5: Sample Report from RDMS-CEDES Showing Descriptors and Data from
Selected Articles
This screen appears when View Selected Articles and Data is selected in the Forms
Switchboard page. Double-click any reference in the top list of references to view the
environmental and human data in the reference.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 14
Chapter 2
INTRODUCTION: COPPER AND MAN

2 INTRODUCTION: COPPER AND MAN

2.1 Copper and Technology: From the Copper Age to the Information Age
Copper, along with gold, is one of the first metals utilized by humans, about 10,000 years ago,
owing to the natural occurrence of its elemental form readily available lumps or leaves in
exposed rock formations (native copper).* In fact, the fashioning of simple tools by hammering
and heating native copper signaled the end of the Stone Age. The series of steps required for
early humans to proceed from a knowledge of the requirements for melting metal to the
intentional practice of ore reduction by heat in the presence of carbon is still open to speculation,
as is the fabrication of copper alloys, believed to have occurred around 3000 to 2000 B.C. The
discovery of coal or charcoal firing rather than wood, which was necessary to obtain sufficiently
high temperatures to melt the native metal, eventually led to the mining of copper ores. From
early workings discovered in the Sinai, mining is known to have begun in about 3800 B.C.
Mines operating in Cyprus around 3000 B.C. were later taken over by the Roman Empire, and
the metal product was called cyprium, later simplified to cuprum, the origin of the Latin name
still used for the metal. It is a fact that for many centuries human technological advancement
basically reflected progress in the processing and utilization of copper. Subsequently, iron and its
alloys dominated metal utilization though copper usage remained extremely important; indeed
copper and its alloys have always been used extensively for plumbing infrastructure, cooking
utensils and artwork. It should be mentioned that recognition of the biological impact of copper
and its compounds (see Section 2.2) took place quite early in history; the ancient Egyptians used
copper salts as biocides.

Because of coppers excellent electrical conductivity, the advent of electricity and of


telecommunications expanded its range of applications. Todays information age has given
copper new status as a high tech metal thanks to its potential for use in microelectronics
applications, thus replacing aluminum in the manufacturing of upcoming generations of
computer microprocessors. (It is interesting to note that the code name of the successor of the
ubiquitous Pentium processor in personal computers has been code-named the coppermine
processor, to emphasize its high-tech potential. In fact the name reflects marketing
considerations since the particular processor is not copper-based at this point.)

2.2 Copper and Biology: Essentiality and Toxicity


Copper is an essential micronutrient for plants and animals, including humans; it is involved in
the function of several enzymes and other proteins needed in a wide range of metabolic
processes. At the same time, high levels of copper can be detrimental to life, thus providing a
means for controlling unwanted organisms. The biological benefits of exposure to copper reflect
this balance between essentiality and toxicity.

*
This accumulation of elemental copper via normal geologic processes is due to its low standard reduction potential of +0.158 V,
significantly below hydrogen in the electromotive series.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 15
Chapter 2
INTRODUCTION: COPPER AND MAN

The recognition of the impact of copper on biological processes took place quite early in human
history. Indeed, it is believed that both the ancient Egyptians and the Chinese were using copper
salts for therapeutic purposes. More detailed information on such practices is available from
around 400 B.C., when Hippocrates prescribed copper compounds for pulmonary and other
diseases. The use of copper compounds to treat disease peaked in the 19th century and then
declined when the treatments were unsuccessful.

2.2.1 Health Effects in Humans


It is beyond the scope of this monograph to review the rapidly expanding knowledgebase on the
biological significance of copper, and the variety of human health effects associated with either
deficient or excessive copper intake. Instead, relevant facts pertaining to the historical evolution
and the current status of the understanding of copper toxicology and health effects, based
partially on Turnlands study (Turnlund, 1999), are presented to provide perspective for the non-
specialist. Table 3 provides a representative list of copper compounds with summary information
on their toxicological characterization (typically, relative to occupational settings). The reader
interested in further details can consult the extensive specialized literature documented in the
1998 ICA Sourcebook; a brief introduction to the health effects associated with copper
essentiality and toxicity can be found in the 1990 ATSDR Toxicological Profile for Copper; for a
more up-to-date summary one can consult a more recent review, such as Turnlands study
(Turnlund, 1999), the Reports of ICA Project No. 223, and the WHO-IPCS 1998 Environmental
Health Criteria 200 document on copper (WHO-IPCS, 1998). It should be mentioned that, in
general, deficiency is considered to be a greater concern than toxicity.

Late 19th century: Copper was identified as a normal constituent of blood, and its toxicity
was described;
By 1900, an anemia that could not be prevented by iron supplements had been observed
in animals kept on a whole milk diet. In 1928 it was reported that this anemia in rats was
controlled by iron only when copper supplements were also given. Experiments in
several animal species produced similar results suggesting that copper-deficiency anemia
occurs in all species.
Human disease was first linked to copper metabolism shortly after Wilson's disease was
described in 1919, long before 1953 when the condition was recognized as a genetic error
of metabolism.
As early as 1930, a relationship between anemia in humans and copper deficiency was
suspected, but because copper supplements improved hemoglobin synthesis in only some
cases, the hypothesis was not well accepted. Conclusive evidence of copper deficiency in
humans was not substantiated until 1964.
Menkes' disease, another anemia disorder, was described in 1962 and was recognized as a
copper absorption disorder in 1972. Since about 1950, an increasing number of diseases,
not specifically disorders of copper metabolism, have been associated with altered,
usually increased, levels of copper in blood or other tissues.
Numerous studies have demonstrated that copper is required for
o infant growth,

COPPER: Environmental Dynamics and Human Exposure Issues


Page 16
Chapter 2
INTRODUCTION: COPPER AND MAN

o host defense mechanisms,


o bone strength,
o red and white blood cell maturation,
o iron transport,
o cholesterol and glucose metabolism,
o myocardial contractility, and
o brain development
An official dietary copper recommendation for an estimated safe and adequate daily
dietary intake was first introduced in 1979 and modified in 1989.
Toxicity resulting from excess copper intake has been observed in numerous studies in a
variety of animal species, such as sheep, cattle, pigs, rats, and poultry, as well as in
humans.

Copper Deficiency
Copper deficiency can result in the expression of an inherited defect such as Menkes disease or
in an acquired condition. Acquired deficiency is a clinical syndrome that occurs mainly in
infants, although it has also been described in both children and adults. This deficiency can be
the consequence of decreased copper stores at birth, inadequate copper supply, inadequate
copper absorption, increased requirements, and increased losses (Amsden et al., 1978). Clinically
evident copper deficiency is a relatively infrequent condition in humans. The most frequent
clinical manifestations of acquired copper deficiency are anemia, neutropenia (compromised
immune response), and bone abnormalities that include osteoporosis and fractures (Shaw, 1992;
Olivares & Uauy, 1996a; Olivares & Uauy, 1996b).

Copper Excess: Acute Toxicity


Acute copper toxicity is infrequent in humans, and usually is a consequence of

accidental consumption by children,


ingestion of several grams in suicide attempts,
application of copper salts to burned skin,
drinking water from contaminated water supplies, or
consumption of acidic food or beverages that were stored in copper containers.

Acute symptoms include (Knobeloch et al., 1994; Turnlund, 1999; Araya et al., 2001)

salivation,
epigastric pain,
nausea,
vomiting, and
diarrhea.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 17
Chapter 2
INTRODUCTION: COPPER AND MAN

Vomiting and diarrhea usually prevent more serious manifestations of copper toxicity that can
include

coma,
shock,
oliguria (dimished urine secretion),
hemolytic anemia,
acute renal (kidney) failure with tubular damage,
hepatic necrosis (liver cell death),
vascular collapse, and
death.

It should be mentioned that the threshold for acute adverse gastrointestinal effects from copper in
drinking water has not been precisely established (see also WHO-IPCS, 1998 for additional
references on the matter).

Copper Excess: Chronic Toxicity


Chronic toxicity in humans is associated principally with Wilson disease, with the occurrence of
infantile cirrhosis in areas of India (Indian childhood cirrhosis), as well as with isolated clusters
of cases in other countries related to excessive copper intake (Anderson et al., 1992; Patel &
Bhattacharya, 1995; Horslen et al., 1994). The most likely explanation for chronic toxicosis
appears to be a genetically determined defect in copper metabolism combined with a high copper
intake (Anderson et al., 1992; Alderdice & McLean, 1982; Aldini et al., 1987; Olivares & Uauy,
1996a; Miller et al., 1995, Mller et al., 1996). Chronic copper toxicosis has also been observed
in dialysis patients following months of hemodialysis when copper tubing was used, and in
vineyard workers using copper compounds as pesticides.

The amount of oral copper intake required to produce toxic effects is not well established, but
liver damage in infants has been reported as potentially related to consuming water with 2 to 3
mg copper/l in early infancy (Muller-Hocker et al., 1988). An extremely wide range of oral
copper dosages, beginning at 0.07 mg/kg per day, has been associated with gastrointestinal
effects (Turnlund, 1999). In addition to Wilson's disease, certain other diseases are associated
with accumulation of toxic levels of copper in the liver and other tissues, even without excessive
intake. Subtle deleterious effects of high dietary copper have also been observed: LDL
cholesterol increased when copper supplements were given to men. On the other hand, however,
lower copper intake has been implicated with other variables such as heightened cholesterol in
some studies as a possible risk factor for cardiovascular disease. Copper plays a critical role in
neurologic diseases; there is speculation that copper-induced production of hydroxy radicals may
contribute to neurodegeneration in Alzheimer's disease (Turnlund, 1999).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 18
Chapter 2
INTRODUCTION: COPPER AND MAN

2.3 Copper and Policy


The majority of human exposures to copper are associated with the oral uptake pathway
(ingestion of drinking water and food). However, in special situations, particularly in
occupational settings, adverse exposures can occur via inhalation and dermal contact. Around the
world, policies and guidelines for copper exposure and intake attempt to balance essentiality and
toxicity.

In the United States, copper and its compounds released by human activities and/or present in
various environmental media are regulated according to the following environment and health
protection acts:

CERCLA: Under the Comprehensive Environmental Response, Compensation, and


Liability Act of 1980 (CERCLA, United States Public Law 96-510), as amended by the
Superfund Amendments and Reauthorization Act (SARA, United States Public Law 99-
499) releases of listed substances at or above their Reportable Quantities (RQs) must be
reported to the National Response Center. RQs are set on the basis of aquatic toxicity,
acute mammalian toxicity, ignitability, reactivity, chronic toxicity, and carcinogenicity,
with possible adjustment on the basis of biodegradation, hydrolysis, and photolysis. The
list of CERCLA hazardous substances and their RQs can be found in 40 CFR 302.4.
(Further information is available from the RCRA/Superfund Hotline: 1-800-424-9346).
RCRA: The Resource Conservation and Recovery Act (RCRA), administered by EPA's
Office of Solid Waste (OSW), addresses the issue of how to safely manage and dispose
of the huge volumes of municipal and industrial waste generated nationwide.
FIFRA: The Federal Insecticide, Fungicide, and Rodenticide Act (FIFRA) creates a
statutory framework under which EPA, through a registration process, regulates the
development, sale, distribution, and use of pesticides.
NPDWR: The National Primary Drinking Water Regulations (NPDWR) under the Safe
Drinking Water Act, Subparts B and G (codified in 40 CFR Part 141) list Maximum
Contaminant Levels (MCLs) for certain chemicals. The MCL is the maximum
permissible level of a contaminant in public drinking water systems. MCLs are based on
health factors, but are also required by law to reflect the technological and economic
feasibility of removing the contaminant from the water supply. (Further information is
available from the Safe Drinking Water Hotline: 1-800-424-4791).

Additionally, copper is also regulated by the Clean Water Act, through inclusion in the Priority
Pollutants List (PPL):

CWA: The Clean Water Act regulates the discharge of pollutants into waterways by
industrial, municipal, and other sources. These sources are subject to effluent limitations
based on guidelines and water quality standards. The Priority Pollutants List (PPL)
consists of approximately 125 pollutants. EPA has developed water quality criteria for all
of them.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 19
Chapter 2
INTRODUCTION: COPPER AND MAN

Table 1 provides an overview of existing international regulations and guidelines for controlling
the levels of copper present in drinking water.

2.3.1 Guidelines and Recommendations for Copper Intake


Frank copper deficiency in humans is very rare, which suggests that the current dietary intake
usually suffices to prevent copper deficiency. It has been suggested that the usual copper intake
is marginal and may not support optimal health, but data in this area are conflicting and not
sufficient to support that hypothesis. Due to the difficulty in measuring copper status via factors
such as zinc, carbohydrate and vitamin C intake, that affect copper bioavailability (see following
chapters), it is very difficult to establish exact requirements for copper intake. For this reason, the
Subcommittee on the Tenth Edition of the RDAs could not establish an RDA for copper and
instead recommended a safe and adequate range of copper intake. While copper is viewed as a
nutrient that is under effective homeostatic control in humans, adequate dietary supplementation
with copper in relation to total parenteral nutrition, and in relation to nutrient-nutrient, hormone-
nutrient and nutrient-pharmaceutical interactions remain important areas for study.

Copper depletion/repletion studies to establish the minimum requirements for healthy humans
were done only recently, with a copper intake low enough to produce systematic reduction in
copper status. Relatively few cases of frank copper deficiency have been reported, and these
were accompanied by confounding factors, such as malnutrition, malabsorption, and excessive
gastrointestinal losses. Hence, their value in establishing a minimum requirement for healthy
individuals is limited.

The most relevant example of a long-term diet containing less than the minimum copper
requirement may be the following (Higuchi et al., 1988) as reported in Turnlund (1999): An
enteral diet containing 15 g copper/100 kcal (0.56 pmol/J) produced copper deficiency in six of
six severely handicapped patients between the ages of 4 and 24 years after they had consumed
the diet for 12 to 66 months. Serum copper values of 1.8 to 7.2 mol/L (11.7-45.7 g/dL) and
ceruloplasmin values of 30 to 125 mg/l (3-12.5 mg/dL) were discovered, accompanied by other
manifestations of copper deficiency. These values increased to within the normal range after 3
months of copper supplementation. By extrapolation (though it may not be valid to extrapolate to
healthy adults, from these growing, severely handicapped individuals) copper deficiency could
be expected to develop eventually, if the diet contained 15 g copper/100 kcal, or 0.44 mg
copper/2900 kcal for men and 0.29 mg copper/1900 kcal for women (0.56 pmol/J). This
example, combined with one study in which healthy young men maintained copper balance and
status at 0.79 mg/day (12 mol/day), and another in which young men did not maintain status at
0.37 mg/day (6 mol/day) (Turnlund, 1999), suggests that the minimum copper requirement for
men is somewhere between 0.4 and 0.8 mg/day (6-12 mol/day). An adult basal copper
requirement within that range, 0.6 mg/day (9 mol/day) for women and 0.7 mg/day (11
mol/day) for men, was suggested by WHO (World Health Organization, 1996). In a recent
report, the National Institute of Medicine (National Institute of Medicine, 2001) has set a new
RDA for copper at a minimum of 0.9 mg/day for adults from food alone and an upper limit of 10
mg/day of copper sourced from a combination of food, drinking water and supplements.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 20
Chapter 2
INTRODUCTION: COPPER AND MAN

Most multivitamin dietary supplements on the market today include 2 mg of copper, the
midpoint of the Safe and Adequate Range of Intake recommended by the FNB.

NRC Recommendations
The daily dietary copper intake recommended by the U.S. National Research Council (NRC) has
been 1.5 to 3 mg (24-47 mol) for adults, and 0.4 to 0.6 mg (6-9 mol) for infants 0 to 6 months
of age, to 1.5 to 2.5 mg (24-39 mol) for children over 11 years of age (Aalbers et al., 1987).

WHO Recommendations
WHO (World Health Organization, 1996) estimated the normative requirement for women to be
0.7 mg/day (11 mol/day) and for men, 0.8 mg/day (12.5 mol/day). After adding margins of
safety to the individual requirement, including all dietary conditions, variations in usual intakes,
and individual variability, a recommendation of 1.25 mg/day (19 mol/day) was derived.

2.3.2 Notes on Recent Regulatory Activities

European Union
In 1998, Council Directive 98/83/EC on the quality of water intended for human consumption
(adopted by the Council on 3 November 1998) established a new drinking water standard of 2
mg/l for copper (from 3 mg/l). The directive entered into force on 25 December 1998; member
states were given two years to transpose the directive into national legislation and five years to
ensure that drinking water complies with the standard set.

United States
In 1998, the California Environmental Protection Agency (EPA) was asked to develop 75 Public
Health Goals (PHGs) over the following three years. The California EPA Office of
Environmental Health Hazard Assessment has already published the PHG for copper in drinking
water as 0.17 mg/l, even while acknowledging the scarcity of good data and the scientific
weakness of the study on which this value is based. This PHG is substantially lower than the
federal drinking water maximum contaminant level goal (MCLG) of 1.3 mg/l, which is itself
35% lower than the World Health Organization provisional drinking water guidance level of 2
mg/l. The California PHG, which is advisory, became effective in 1999 and the industry is
concerned about the potential for a derivative regulatory state action level. Further, under
California law, consumer notification will be required under the U.S. Lead and Copper Rule if
PHG exceedences are found during residential tap water monitoring. Simultaneously, the
California Department of Housing and Community Development (CalDHCD) is considering
whether to continue the ban on the use of cross-linked polyvinyl chloride (CVPC) plumbing
tube. To make this determination, CalDHCD is preparing an Environmental Impact Report (EIR)

COPPER: Environmental Dynamics and Human Exposure Issues


Page 21
Chapter 2
INTRODUCTION: COPPER AND MAN

that will compare the economics, technology and risks of copper and plastics. The PHG, as well
as existing and future water quality, will be given careful consideration in the EIR analysis.

In December 1999, the U.S. EPA made minor revisions to the Lead and Copper Rule, effective
April 11, 2000. These revisions, the Lead and Copper Rule Minor Revisions or LCRMR,
streamline requirements, promote consistent national implementation, and in many cases, reduce
burden for water systems. The LCRMR do not change the action levels of 0.015 mg/l for lead
and 1.3 mg/l for copper, or Maximum Contaminant Level Goals of 0 mg/l for lead and 1.3 mg/l
for copper, established by the 1991 Lead and Copper Rule (the rule). They also do not affect
the Rule's basic requirements to optimize corrosion control and, if appropriate, to treat source
water, deliver public education, and replace lead service lines. As part of the LCRMR
rulemaking process, the Agency collected additional data pertaining to exclusion of transient
non-community water systems from the Rules requirements. EPA concluded that it is still
appropriate to continue this exclusion because the Agency believes there are de minimus
(Andersson et al., 1991) non-carcinogenic adverse health effects resulting from exposure to lead
in drinking water at such systems. All water system operators and managers of community water
systems (CWSs) and non-transient non-community water systems (NTNCWSs) are potentially
affected by the LCRMR, as are state staff responsible for implementing the Lead and Copper
Rule in their state. (More information is provided by EPA's Safe Drinking Water Hotline, 1-800-
426-4791, and by the Office of Ground Water and Drinking Water web page at
http://www.epa.gov/safewater/standards.html.)

In 1997, in accordance with the Clean Air Act, the U.S. EPA signed a new national ambient air
quality standard for particulate matter and ozone, which provides increased protection against a
wide range of particulate-matter-related health hazards and could have a significant impact on
the mining and mineral-processing industries. The existing standard regulates particles that are
10 microns in size or smaller and sets an average annual concentration limit of 50 micrograms
per cubic meter and an average daily limit of 150 micrograms per cubic meter. Expected to take
effect between 2000 and 2002 (Platts Metals Week, 1997e), the new standard for particles 2.5
microns or smaller is an annual mean of 15 micrograms per cubic meter and a 24-hour mean of
65 micrograms per cubic meter.

In 1989, the Basel Convention on the Control of Transboundary Movements of Hazardous


Wastes and their Disposal came into force and has since been ratified by more than 100
countries, including the United States. However, the U.S. has not passed legislation necessary to
implement the Convention. An international Technical Working Group met four times between
September 1995 and February 1997 to consider which materials should be classified as
hazardous and, hence, affected by the various bans. The group elected to include copper scrap,
copper slags, and copper oxide mill scale in the B list, the list of materials not covered by the
Basel Convention as hazardous and, thus, not subject to any export ban. In February 1998, the
Basel Convention held its Fourth Meeting of the Council of Parties in Kuching, Malaysia, and
adopted the A list, wastes characterized as hazardous and therefore subject to regulation, and the
B list, as developed by the Technical Working Group. However, material contained in the B list
is not precluded from regulation, if it contains any of a core list of materials to be controlled

COPPER: Environmental Dynamics and Human Exposure Issues


Page 22
Chapter 2
INTRODUCTION: COPPER AND MAN

because they exhibit hazardous characteristics. (More information can be found on the web site
of the International Copper Study Group (ICSG) at http://www.icsg.org.)

COPPER: Environmental Dynamics and Human Exposure Issues


Page 23
Chapter 2
INTRODUCTION: COPPER AND MAN

Table 1: International Regulatory Practices for Copper in Drinking Water*


Regulates Guiding New
copper in Max. factors Planned Regulations
Drinking copper Mandatory (Aesthetic, Sampling changes to to be
Country Water Levels or Guideline Health) Basis for Limits Basis regulations Introduced
EU Yes 2 mg/l M Aesthetic 1998 Cu NA Yes No
(smell, taste, directive
color) 98/83/EC is
currently in
effect.
Belgium Yes 1 mg/l M Aesthetic NA Random. Possibly Based on new
(stain) Regional EU Drinking
responsibility Water
Directive
Finland Yes 2 mg/l M Health (WHO EU 1998 NA No No
Guidelines) Directive
France Yes 1 mg/l G Aesthetic Historical levels Random Yes Based on new
EU Drinking
Water
Directive
Germany Yes 2 mg/l G NA EU 1998 NA Yes Based on new
Directive EU Drinking
Water
Directive
Greece NA 2 mg/l G NA EU 1998 NA NA Based on new
Directive EU Drinking
Water
Directive
Italy Yes 1 mg/l M NA Historical levels NA Yes Based on new
EU Drinking
Water
Directive
Portugal Yes 2 mg/l M Aesthetic EU 1998 At tap after 12 NA Based on new
(taste and Directive hours standing EU Drinking
stain) Water
Directive
Spain Yes 2 mg/l G Aesthetic EU 1998 NA Yes Based on new
(taste, Directive EU Drinking
corrosion, Water
stain) Directive
Sweden Yes 0.2 mg/l M Aesthetic Scientific Sampling sites: Yes
at tap (corrosion, research Water works
after taste, stain epidemiological and consumers
flushing; and green studies, by tap. At
2 mg/l no hair) experience, EU consumers' tap,
flush Health drinking water without and
regulations, after flushing.
WHO guidelines
United Yes 2 mg/l M Health WHO Tap samples Yes Based on new
Kingdom Guidelines, EU Drinking
EU 1998 Water
Directive Directive
Argentina Yes 1 mg/l G Health NA Reviewing No No
procedures
Canada Yes 1 mg/l G Aesthetic Staining Provincial No No
(taste, stain) @>1mg/l responsibility
Recognize
essentiality of
copper.

*
This table was adapted from information provided at www.ics.org and should not be considered complete. It contains as much
reliable information as could be gathered at the time that this report was written.
Under normal circumstances EU member states are expected to amend their national regulations within two years of formal
promulgation of a new EU Directive.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 24
Chapter 2
INTRODUCTION: COPPER AND MAN

Regulates Guiding New


copper in Max. factors Planned Regulations
Drinking copper Mandatory (Aesthetic, Sampling changes to to be
Country Water Levels or Guideline Health) Basis for Limits Basis regulations Introduced
Chile Yes 1 mg/l M Aesthetic 1985 WHO Min. one per No No
Aesthetic year from
parameter source
China Yes 1 mg/l NA NA NA NA Yes 2001
Czech Yes 0.1 mg/l M Health Set in 1970's. Single sample Yes 1999/2000
Republic Based on sovietNote: when
study suggesting
copper salts are
embryotoxic needed for
effects of copper
algae control, a
limit of 1 mg/l
is permitted for
a transient
period.
Ghana Yes 1.0 mg/l NA Aesthetic WHO Guidelines NA No No
(smell, taste,
stain)
India Yes 1.5 mg/l G NA WHO 1984 Per Indian No No
Guidelines, Standards
Indian Council IS1622:1981
of Medical and IS3025
Research 1971 PartI: 1987
Japan Yes 1.0 mg/l M Aesthetic Lower than Service pop. > No No
(stain) WHO 100,000 :
provisional sampled every
levels to prevent month
staining
Malaysia Yes 1.0 mg/l G Aesthetic WHO Guidelines Sample No No
(stain, taste, analyzed by the
corrosion) Chemistry
Health (Liver Dept. of
damage from Malaysia based
long-term on
accumulation) International
Recom. Standard
Dietary Reqt: Procedures,
20-80 g/kg Preservation
body and Protocol
weight/day under ISO
Guide 25
Mexico NA NA NA NA NA NA NA NA
Australia NA 2 mg/l NA Aesthetic NA NA NA NA
(1 mg/l)
New Yes G Aesthetic Staining value Aesthetic: No No
Zealand (staining, selected from Investigation of
1mg/l) experience with complaints
Health complaints Health: First
parameters which start at sample of the
(projectile about 2mg/l; day at the end
vomiting, Projectile of the spur-
2mg/l) vomiting in unflushed
children starts at
about 4 mg/l
Norway Yes 0.3mg/l G NA NA NA No No
Peru Yes 1 mg/l G NA Copper sulfate NA No No
used as algicide
in doses of less
than 3 mg/l per
week during
normal water
treatment,
paying attention
that filtered
levels are free of
metallic
particulate

COPPER: Environmental Dynamics and Human Exposure Issues


Page 25
Chapter 2
INTRODUCTION: COPPER AND MAN

Regulates Guiding New


copper in Max. factors Planned Regulations
Drinking copper Mandatory (Aesthetic, Sampling changes to to be
Country Water Levels or Guideline Health) Basis for Limits Basis regulations Introduced
United Yes 1.3mg/l M Health LOAEL of Sample sites No No
States (NOAEL) 5.3mg/day selected based
adjusted to std. on high risk for
daily dose of 2l/d Pb
by adult. contamination.
Uncertainty Number
factor of 2 depends on
applied system size.
Venezuela Yes 2mg/l M Health WHO NA No No
recommendation
s (upper limit)
Zambia Yes Yes NA NA NA NA NA NA

COPPER: Environmental Dynamics and Human Exposure Issues


Page 26
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

3 A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL


PROPERTIES OF COPPER
A brief review of essential physical, chemical and biological attributes of copper is presented
here to provide a quick reference and the necessary background for analysis of environmental
and biological chemodynamics of copper, the subject of the next chapter.

3.1 Physical and Chemical Attributes of Copper


Since early history (see Chapter 2) copper's unique combination of properties have made it one
of the world's most important metals. These properties include among others:

appearance,
malleability,
low corrosion,
alloying ability,
high thermal conductivity, and
high electrical conductivity.

Properties of metallic copper, such as electrical conductivity and fabricability, vary markedly
with purity. Standard classifications have been defined according to processing method. For
example, ASTM B5-74 is 99.90% pure and is the accepted basic standard for electrolyte copper
wire bars.

General Chemical Properties


Copper is a transition metal (the first element of Group IB of the periodic table) with atomic
mass of 63.54 daltons (Da), it has two stable isotopes, 63Cu and 65Cu, with natural abundances of
69.2 and 30.8%, respectively.* Copper generally occurs in nature in one of four oxidation states:
copper(0), copper(I), copper(II), and copper(III); although trivalent copper is very rare. Along
with silver and gold, it is classified as a noble metal and, like them, can be found in nature in an
elemental form. Indeed, copper metal is fairly unreactive, due to its high nuclear charge, small
size, and consequent high ionization potential. Having one more electron than nickel and a
higher nuclear charge, copper has a smaller atom and more tightly bound electrons. Since it is
positioned below hydrogen in the electromotive-force series, it will not displace hydrogen ions
from dilute acid. Accordingly, copper will not dissolve in acid unless an oxidizing agent is
present. Therefore, while it readily dissolves in nitric acid and hot concentrated sulfuric acid, it
dissolves slowly in hydrochloric acid and dilute sulfuric acid, only when exposed to the
atmosphere. It is also attacked by acetic acid, and other organic acids. Copper is stable in pure
dry air at room temperature, but in damp air it forms a green patina of basic salts such as the
carbonate (and, near the sea, the chloride). This tightly adhering coat protects the underlying
*
There are seven radioisotopes of copper, most with half-lives of seconds or minutes; the two with the longest half-lives, 67Cu
(61.9h) and 64Cu (12.9h), and the two stable isotopes are used as tracers of copper metabolism.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 27
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

metal from further attack and is also prized for its appearance. Despite its low reactivity, copper
forms a wide range of compounds, and its complexes, both of copper(II) and of copper(I), are
amongst the strongest known for their oxidation states. Although copper is a soft metal, it can
form harder alloys with metals that contribute more outer electrons, such as tin and zinc. Copper,
with its high nuclear charge and small size, is found in combination with both oxygen and sulfur.
Copper is the only member of the 3d series that, in nature, occurs in the (I) state (e.g. as Cu2O or
Cu2S), as well as in the (II) state (e.g. as the basic carbonate Cu2CO3(OH)2). Copper can also
exist in higher oxidation states, but these are rare. Currently an area of very intense investigation
is that of mixed oxide ceramic superconductors which copper forms with other cations. These
function at relatively high temperatures (up to 125 K for Ti2Ba2CaCu3O10). The oxide
Yba2Cu3O7-x (nicknamed 1-2-3) appears to contain copper in oxidation states (I), (II), and
(III), but little is known little about how it functions.*

Copper(I)
Copper(I), the cuprous ion, disproportionates rapidly in aqueous solution to form copper(II) and
copper(0), the change being favored by the high hydration energy of Cu2+ compared with Cu+.
The Cu+ ion is, however, stabilized by solvents such as CH3CN, which solvate it strongly; and by
groups with which it forms insoluble solids, such as CuI and CuCN. It has been shown, however,
that copper(I) complexes may be formed in seawater by photochemical processes and may
persist for several hours (Moffett & Zika, 1987a). Cuprous compounds are generally colorless.
Copper(I) forms a number of polymeric species; tetramers are particularly favored. Copper(I)
also forms a variety of simple organometallic compounds: the explosively unstable ionic
acetylicle Cu2C2; sigma-bonded aryls and alkyls, CuR; and complexes containing pi-bonded
ethene and ethyne. The chloride combines with CO to form the 18-electron compound
Cu(CO)3Cl but no other copper carbonyl compound has been made at ordinary temperatures.

Copper(II)
Copper(II), the cupric ion, is the most important oxidation state of copper; indeed, it is the
oxidation state generally encountered in water. Cupric ions are coordinated with six water
molecules in solution; the arrangement of the water molecules is distorted in that four molecules
are closely bound to the copper in a planar array and the other two are more loosely bound in
polar position. Most cupric compounds and complexes are blue or green, and are frequently
soluble in water. Cu2+(aq) is mildly hydrolysed in near-neutral solution, forming the dimer
Cu2(OH)22+. Addition of alkali gives a precipitate of the hydroxide, which is somewhat soluble in
excess of alkali, probably forming the Cu(OH)42- anion. Cu2+ ion forms stronger complexes than
other doubly charged metal ions, and, because part of this enhanced stability is a result of ligand
field effects, it is not surprising that nitrogen ligands, which produce relatively high fields, bind
copper(II) even more strongly than do lower field oxygen donors. Copper(II) complexes are
*
This compound of copper received extensive attention following publication of reports that a spinning disc of 1-2-3 causes a
small (2%) decrease in the force of gravity. (Rossotti, 1998)

COPPER: Environmental Dynamics and Human Exposure Issues


Page 28
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

almost always blue or green. With halide ions, copper(II) shows a rich diversity, except with
iodide, which it oxidizes to I2.

3.2 Biological Functions of Copper


In general, coppers biological functions involve electron transfer catalysis, by means of its two
accessible oxidation states (given suitable ligands), which differ by one unit, and can tolerate a
variety of geometries. Copper is widely used by microorganisms, plants, and animals as a
component of many electron transfer enzymes. Detailed descriptions of these proteins and their
functions have been published (Owen, 1982a; Linder, 1991a; Linder, 1991b). A dimeric
dioxygen-bridged copper protein, haemocyanine, is used as an oxygen carrier in some
invertebrates. Other such proteins include ascorbate oxidase, carboxypeptidase A, laccase, and
uricase. Like many other functional bioinorganic proteins, their mode of action is incompletely
understood. The unraveling of these processes, together with a deeper insight into reasons for the
varied stereochemistry of Cu(II), and for the formation of so many copper dimers and cluster
compounds, expected to be subjects of research for many years.

It should be noted that copper is most often in biological systems as Cu2+; in fact, at least three
distinct types of the bound cation can be found in copper containing enzymes, often in
combination within a single protein (Owen, 1982a; Owen, 1982b):

Type 1 refers to deep blue proteins, typically copper-containing oxidases,


Type 2, characteristic of many multicopper oxidases, is not blue, but is detectable by
electron paramagnetic resonance (EPR),
Type 3, is neither blue, nor detectable by EPR.

3.2.1 Biochemical Functions


A brief description (adapted from Turnlund, 1999) of the major copper-containing proteins found
in the human organism follows; for details the reader should consult (Owen, 1982a; Linder,
1991a).

Copper-Containing Enzymes Found in Humans

Amine Oxidases
Several important amine oxidases are cuproproteins. Relatively small amounts of these enzymes
are found circulating in blood plasma, where they inactivate and catabolize physiologically
active amines such as histidine, tyramine, and polyamines. They are found in tissues throughout
the body. Their activity is elevated when connective tissue activation and deposition take place,
in such conditions as liver fibrosis, congestive heart failure, and hyperthyroidism, and during
childhood, and senescence.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 29
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

Monoamine Oxidase. Involved in inactivation of catecholamines, monoamine oxidase


reacts with substances such as serotonin, norepinephrine, tyramine, and dopamine, and is
inhibited by tricyclic antidepressant drugs.
Diamine Oxidase. A number of copper-dependent diamine oxidase enzymes are found in
cells throughout the body. Diamine oxidase inactivates histamine, acting in the small
intestine, where histamine stimulates acid secretion, and in allergic reactions throughout
the body, where histamine is released in response to antigens. It also inactivates
polyamines involved in cell proliferation, which suggests that diamine oxidase may play
a role in limiting excessive growth. Diamine oxidase activity is highest in the small
intestine. Activity is also high in the kidney, where diamine oxidase inactivates diamines
filtered from the blood, and in maternal placenta, where it may inactivate amines
produced by the fetus.
Lysyl Oxidase. Lysyl oxidase, a unique amine oxidase, functions in the formation of con-
nective tissue, including bone, blood vessels, vasculature, skin, lungs, and teeth. It acts on
lysine and hydroxylysine side chains of collagen and elastin; it eliminates the lysine of
newly formed, immature elastin and collagen, after which cross-links are formed.
Concentrations are highest during development. Long-term estrogen treatment increases
its activity, and malignant transformation decreases it.
Peptidylglycine-a-Amidating Monooxygenase. A newly identified cuproenzyme,
peptidylglycine-a-amidating monooxygenase, is involved in the synthesis of many
bioactive peptides and may be influenced by copper deficiency.

Ferroxidases
Ceruloplasmin. Ceruloplasmin, also called ferroxidase 1, is an alpha-2 glycoprotein with
a molecular weight of about 132 kDa. It contains six (possibly seven) atoms of copper per
molecule, including all three types of copper(II) atoms described previously. Four copper
atoms appear to be involved in the oxidation/ reduction reactions catalyzed by the
enzyme. The role of the other atoms is not yet understood. Ceruloplasmin catalyzes
ferrous iron oxidation and plays a role in the transfer of iron from storage to hemoglobin
synthesis sites. It also oxidizes aromatic amines and phenols. 60 to 95% of copper in
blood plasma is bound to ceruloplasmin, and this fraction appears to be relatively
constant within an individual, while varying considerably among individuals.
Ferroxidase II. Ferroxidase II also catalyzes ferrous iron oxidation. It accounts for only
about 5% of ferole in some animal species.

Cytochrome c Oxidase
Cytochrome c oxidase enzyme, present in the mitochondria of cells throughout the body, is the
terminal link in the electron transport chain. It reduces O2 to form water and permits formation of
adenosine triphosphate (ATP) in mitochondrial energy production. Cytochrome c oxidase is
considered the single most important mammalian cell enzyme, because it is rate limiting in
electron transport. It contains two or three copper atoms per molecule. The activity of this
enzyme is high in brain, liver, and kidney tissues, highest in the heart.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 30
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

Dopamine beta-Hydroxylase
Dopamine beta-hydroxylase catalyzes the conversion of dopamine to the neurotransmitter
norepinephrine in the brain. Estimates of the copper content of dopamine beta-hydroxylase range
from two to eight atoms per molecule, the most recent estimates being eight. Dopamine
beta-hydroxylase concentration is two to three times higher in brain gray matter than white
matter, and it is present in the adrenal gland, where it is required for epinephrine production.

Superoxide Dismutase
Extracellular Superoxide Dismutase (EC-SOD). EC-SOD, a copper-containing enzyme,
is present in high amounts in the lungs, thyroid, and uterus and in small amounts in blood
plasma. It functions as a scavenger of superoxide radicals and protects against oxidative
damage.
Copper/Zinc Superoxide Dismutase (SOD). Copper/zinc SOD, which contains two
copper atoms per molecule, is present in most cells of the body, primarily within the
cytosol. High concentrations are found in human brain, thyroid, liver, pituitary,
erythrocytes, and kidney. SOD erythrocyte levels are high in alcoholics and individuals
with Down's syndrome. It protects intracellular components from oxidative damage,
converting the superoxide ion to hydrogen peroxide, and requires both zinc and copper
for catalytic function.

Tyrosinase
Tyrosinase catalyzes conversion of tyrosine to dopamine, and oxidation of dopamine to
dopaquinone, steps in melanin synthesis. It is present in the melanocytes of the eye and skin and
is responsible for hair, skin, and eye color. Tyrosinase deficiency causes albinism.

Copper-Binding Proteins

Metallothionein (MT)
MTs are small nonenzymatic proteins, rich in cysteine, that are responsible for binding copper.
Each molecule can bind 11 or 12 copper atoms, as well as zinc and cadmium. They appear to
play a role in metal storage by sequestering excess metal ions, thus preventing toxicity. MTs are
found in many human tissues, with highest concentration in the liver, where metals accumulate
in MT fractions. Their presence in small amounts in blood plasma suggests that they also play a
role in copper transport, but if so, the role would be minor.

Albumin
Albumin, a protein with a molecular weight of 68,000, is the most prevalent protein in blood
plasma and interstitial fluids. Albumin binds and transports copper and may also play a role in
binding excess copper that would otherwise be toxic. Estimates of the fraction of copper in blood
plasma bound to albumin range from 5 to 18%.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 31
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

Transcuprein
Transcuprein, a recently isolated plasma protein with a molecular weight of about 270,000, binds
copper and is found in humans. It has not yet been completely characterized and its functions are
not clear, but it may play a role in copper transport. Considerably less serum copper is bound to
transcuprein than to albumin.

Blood-Clotting Factor V
Blood-clotting factor V, a nonenzymatic component of the blood-clotting process, has recently
been found to contain one atom of copper per molecule. Although this indicates that copper is
required for blood clotting, impaired blood clotting is not among the reported manifestations of
copper deficiency.

Low-Molecular-Weight Ligands
Amino acids and small peptides also carry a small fraction of the copper in the blood plasma.
Estimates range from less than 1% to 4%. Histidine, glutamine, threonine, and cystine are
examples of amino acids that bind copper in plasma, and at least one copper peptide complex,
glycyl-histidine-lysine, has been isolated from human plasma. The role of these complexes is not
known, but the copper carried by low-molecular-weight ligands is thought to exchange with
nonceruloplasmin copper in the blood. The ligands may carry copper to cells.

3.2.2 Physiologic Functions


Much of coppers physiologic activity is related to reactions catalyzed by cuproenzymes, some is
related to copper deficiency. A brief review, based on Turnlund, 1999, is included here for
convenient reference; more detailed information on the physiologic functions of copper can be
found in (Davis & Mertz, 1987).

Connective Tissue Formation. Copper, through the enzyme lysyl oxidase, is essential for
cross-linking collagen and elastin, both required for formation of strong, flexible
connective tissue. Thus, copper plays a role in bone formation, skeletal mineralization,
and heart and vascular system connective tissue integrity. Lysyl oxidase activity declines
during severe copper deficiency in weanling rats, and the resulting defects in connective
tissue formation may be responsible for the multiple effects of copper deficiency on
cardiac system integrity and bone formation. Copper depletion also results in modest
changes in lysyl oxidase activity in the skin, which, owing to its large excess there, does
not compromise function.
Iron Metabolism. Several mechanisms have been proposed for the role of copper in iron
metabolism and N-thropoiesis. Ceruloplasmin and ferroxidase oxidize ferrous iron, so
that it can be transported from the intestitial lumen and storage sites to sites of
erythropoiesis. This may explain why anemia develops with copper deficiency, through
iron accumulates in the intestinal lumen and liver. Copper may also be required for
formation of normal bone marrow cells, necessary for red blood cell formation.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 32
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

Central Nervous System. Copper plays more than one role in the central nervous system.
It is required for formation and maintenance of myelin, a protective layer covering
neurons, composed primarily of phospholipids. Phospholipid synthesis depends on
cytochrome c oxidase activity, which may explain why copper deficiency leads to poor
myelination, necrosis of nerve tissue, and neonatal ataxia in copper-deficient animals.
The role of cuproenzymes in catecholamine metabolism (conversion of dopamine to
norepinephrine by dopamine beta-hydroxylase and the degradation of serotonin,
norepinephrine, tyrainine, and dopamine by monoamine oxidase) implies a function in
normal neurotransmission.
Melanin Pigment Formation. The role of copper in the pigmentation of skin, hair, and
eyes is related to the requirement for tyrosinase in melanin synthesis. Depigmentation of
hair and skin is observed with copper deficiency in several animal species and in Menkes'
disease.
Cardiac Function and Cholesterol Metabolism. The role of copper in cardiac function
has been explored in a number of laboratory animal experiments. Cardiac myopathy and
a variety of other conditions appear when weanling, but not older, rats are deprived of
copper. Cardiac symptoms have not been reported in the few frankly copper-deficient
humans, though links to heart irregularities in humans have been suggested. Blood
cholesterol levels increase in animals fed copper-deficient diets, but results of studies on
the effects of low-copper diets on blood cholesterol in humans are inconsistent; levels
have increased in some and declined in others, and copper supplementation increased
low-density lipoprotein (LDL) in a study in adult males.
Other Functions. Other physiologic functions suggested for copper are not as well
understood as those described above. These include roles in thermal regulation and glu-
cose metabolism. Its known role in blood clotting through factor V has not yet been
clearly associated with clinical manifestations of copper deficiency. Copper is known to
be both prooxidant and antioxidant. Two key antioxidant enzymes, ceruloplasmin and
superoxide dismutase, decrease in copper deficiency and may result in impaired
antioxidant status. Recent evidence suggests a role for copper in immune function; both
low- and high copper intake influenced immune function in laboratory animals. Some
indices of immune function declined with copper depletion of humans, but were not
reversed by higher copper intake. The effect of dietary copper on these functions is still
the subject of research (Turnlund, 1999).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 33
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

Table 2: Copper Containing Proteins Found in Humans


Protein Site of Action Function
Monoamine Oxidase Brain Inactivates serotonin,
norepinephrine, tyramine, and
dopamine
Diamine Oxidases Small intestine, kidney, and other Inactivates histamine, diamines,
tissues and polyamines
Lysyl Oxidase Extracellular matrix Eliminates lysine and hydrolysine
side chains of collagen and
elastin
Ferroxidase 1I Blood plasma Catalyzes oxidation of ferrous
iron
Cytochrome c Oxidase Mitochondria of cells throughout Permits formation of adenosine
the body triphosphate (ATP)
Dopamine -hydroxylase Brain adrenal gland Catalyzes conversion of
dopamine to norepinephrine
Tyrosinase Melanocytes of the eye and skin Catalyzes conversion of tyrosine
to dopamine and oxidation of
dopamine to dopaquinone
Metallothionein Liver (and GI tract tissues) Binds copper (and other metals)
Albumin Blood plasma Binds copper
Transcuprein Blood plasma Binds copper
Blood-Clotting Factor V Blood plasma Required for blood clotting
Extracellular Superoxide Extracellular matrix, especially Scavenges superoxide radicals
Dismutase lung, thyroid, and uterus
Copper Superoxide Dismutase Most cells in body

COPPER: Environmental Dynamics and Human Exposure Issues


Page 34
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

Table 3: Selected Industrial Copper Compounds and their Properties


Name Formula Physical/Chemical Toxicity Studies Summary Toxicity
Properties Statement
Copper Abietinate Cu(C19H27O2)2 Green scales
(Cupric Abietinate)
Copper Acetate Cu(C2H3O2)2 H2O Greenish-blue, fine Oral LD50 (rat) = 710
powder or small mg/kg
crystals; mp: 115, bp:
240, d: 1.882, (anhy):
1.93
Copper Acetate, Basic Cu(C2H3O2)2-CuO-6H2O
Copper Acetoarsenite C4H6O16Cu4As6 Emerald green powder Acute tox data: Oral
LD50 (rat) = 22 mg/kg;
LD50 oral (mammal) = 18
mg/kg
Copper Ammonium Sulfate CuSO4-4NH3-H2O
Copper Arsenate, Basic Cu(CuOH)AsO4
(Cuprous Arsenate)
Copper Arsenide Cu5As2 Black crystals;
mp: decomp,
d: 7.56
Copper Arsenite (Cupric CuHAsO3 Yellowish-green
Arsenite, Sheele's Mineral) powder;
mp: decomp.
Copper Boride (Cupric Cu3B2 Yellow crystals;
Boride) d: 8.116
Copper Carbonate CuCO3-Cu(OH)2 Green powder; mp: Acute tox data: Oral
Hydroxide (Cupric decomp @ 200, LD50 (rat) = 159 mg/kg;
Carbonate) d: 4 oral LD50 (birds) = 810
mg/kg
Copper Chloride (Cupric CuC12 Yellowish- Acute tox data: Oral HIGH via oral and inhal
Chloride) brown hygroscopic LD50 (rat) = 140 mg/kg; routes. Used as a
powder; mp: 498, d: oral LD50 (human) = 200 fungicide. Also a trace
3.054 mg/kg mineral added to animal
feed.
Copper-Y-Chloroaceto Cu(C8H7-CINO)2 See copper compounds
Acetanilide acetanilide and chlorides.
Copper-8-Cunilate See copper compounds
Copper Cyanide (Cupric Cu(CN)2 Yellowish-green Acute tox data: ip LD50 HIGH via ip route. See
Cyanide) powder; (rat) = 50 mg/kg cyanides and copper
mp: decomp before compounds
melting
Copper Diazo Amino Orange crystals; mp: See copper compounds
Benzene 270 (decomp)
Copper Dichlorobenzoate A toxic material. See
copper compounds. Used
as a fungicide.
Copper Dimethyl Acute tox data: ip LD50 HIGH via ip route. See
Dithiocarbamate (rat) = 25 mg/kg carbamates and copper
compounds
Copper Fluoride (Cupric CuF2 2H2O Monoclinic blue crystal; See fluorides and copper
Fluoride) d: 2.9 compounds.
Copper Fluoroacetic Acid CuFOOCCH3 Acute tox data: Oral HIGH via oral and inhal
LD50 (rat) = 10 mg/kg routes. See fluorides
Copper Gluconate (Cupric [CH2OH(CHOH)4COO]2C Light blue, fine U. See copper
Gluconate) u crystalline powder compounds. A nutrient
and/or dietary
supplement food
additive. Also a trace
mineral added to animal
feed
Copper Hydride CuH Red-brown crystals; See copper compounds
mp: decomp @ 60, and hydrides
d: 6.38
Copper Hydroselenite Cu(HSeO3)2 Bluish-green, tiny See selenium and copper
prisms compounds.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 35
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

Name Formula Physical/Chemical Toxicity Studies Summary Toxicity


Properties Statement
Copper Hydroxide (Cupric Cu(OH)2 Blue gelatinous or HIGH via oral and inhal
Hydroxide) amorphous powder routes. A trace mineral
added to animal feeds.
Used as a fungicide.
[109] See copper
compounds
Copper Naphthenate (C6H5COO)2Cu Solid; flash p: 100F, d: Acute tox data: Oral HIGH via oral and inhal
1.055 LD50 (mouse)- 110 routes
mg/kg
Copper Nitrate (Cupric Cu(NO3)2 Blue, deliquescent Oral LD50 (rat) = 940 MOD via oral route
Nitrate) crystals; mp: 114.5, mg/kg
d: 2.047
Copper Nitride Cu3N Dark-green powder; See copper compounds
mp: decomp @ 300, d: and nitrides
5.84 @ 25/4
Copper Nitrodithioacetate Solid See copper compounds
Copper Oleate (Cupric Cu(C18H33O2)2 Brown powder Used as a fungicide. See
Oleate) or greenish-blue mass copper compounds and
oleic acid. A recog
carcinogen
Copper Oxalate (Cupric CuC2H4-1/2 H2O Light bluish-green See oxalates and copper
Oxalate) powder compounds.
Copper Oxide, Black Fine black powder; bp: Used as fungicide. Also a
(Cupric Oxide, decomp @ 1026, d: trace mineral added to
Paramelaconite) 6.4. animal feeds
Copper Oxychloride CuC12-2CuO-4H2O Emerald green to Acute tox data: Oral HIGH to MOD via oral
(Brunswick Green, Cupric greenish-black powder; LD50 (rat) = 700 mg/kg; and inhal routes
Oxychloride) mp: -3H20 @ 140. oral LD50 (human) = 200
mg/ kg
Copper-2,4-Pentanedione Cu(C6H7O2) Blue crystals;
Derivative mp: >230, bp: subl.
(Acetylacetonate of
Copper)
Copper Perchlorate Cu(ClO4)2 Crystalline Acute tox data: ip LD50 HIGH via ip route. See
mp: 60. (rat) = 29 mg/kg perchlorates and copper
compounds.
Copper Peroxide CuO2 Brown or See copper compounds
brownish-black crystals and peroxides.
Copper-3-Phenyl Salicylate C26H18CuO6 Crystalline; mp: 145. Acute tox data: Oral MOD via oral and inhal
LD50 (rat) = 520 mg/kg routes. See copper
compounds
Copper Phosphate (Cupric Cu3(PO4)2 Bluish-green powder A trace mineral added to
Phosphate) animal feeds. See copper
compounds and
phosphates.
Copper Phosphide. Syn: Cu3P2 mp: decomp, See phosphides and
Cupricphosphide. d: 6.67. copper compounds.
Mixed with KNO3 or
KC1 O3, can explode.
Copper Propargylate Solid. See copper compounds.
Copper Propionyl Acetate CuC5H9O Crystals. See copper compounds.
Copper Pyrophosphate U. A trace mineral added
to animal feeds.
See also copper
compounds and
phosphates
Copper-8-Quinolinolate. C18H12N2O2Cu Yellow-green powder ip LD50 (mouse) = 67 HIGH via ip route. An
Syn: Copper-8-Hy- Mp: decomp @ 210. mg/ kg. [3] exper () neo via oral and
Droxyquinoline sc routes. A fungicide.
See copper compounds.
Copper Resinate Cu(C20H29O2)2 Green powder See copper compounds
Copper Ricinoleate Cu2(CO2(CH2)7- Green plastic solid See copper compounds.
CHCHCH2CHOH- Softening p: 64.
(CH2)5)
Copper Sebacate Cu(CH2)8C2O4 Solid See copper compounds
and silicates.
Copper Silicate CuSiO3 Greenish crystals

COPPER: Environmental Dynamics and Human Exposure Issues


Page 36
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

Name Formula Physical/Chemical Toxicity Studies Summary Toxicity


Properties Statement
Copper Silicide Cu4Si White metallic crystals mp: 850, d: 7.53. See copper compounds
and silanes.
Copper Stearate Syn: Cu(C18H35O2)2 Light blue mp: 1250.
Cupric Stearate amorphous powder
Copper Suboxide Cu4O Olive green crystals mp: decomp
Copper Subsulfate Syn: 4CuO-SO3 Light blue powder Oral LD50 (human) = 200 HIGH via oral route. See
Cupric Sulfate, Basic. mg/ kg copper compounds and
sulfates
Copper Sulfate. Syns: Blue CuSO4 - 5H2O Blue crystals or blue, Oral LD50 (rat) = 960 HIGH via ip; MOD via
Vitriol, Blue Stone, crystalline granules or mg/ kg; ip LD50 (mouse) oral and inhal routes
Roman Vitriol powder = 33 mg./kg
mp: -4H20 @ 110,
d: 2.284.
Copper Sulfate, CuSO4-4NH3-H2O Dark blue crystals
Ammoniated. Syn: Cupric
Sulfate, Ammoniaied
Copper Sulfide. Syn: CuS Black powder See copper compounds
Cupric Suffide or crystal, and sulfides
mp: transition @ 103,
bp: decomp @ 220, d:
4.6
Copper Tellurite CuTeO3 Green solid. See tellurium compounds
and copper compounds.
Copper Tetrazol See copper compounds
Copper Thiocyanate. Syn: CuCNS White to yellowish See copper compounds
Cuprous Thiocyanate powder, and thiocyanates.
mp: 1084, d: 2.85.
Copper Trichlorophenate Cu(Cl3C6H20)2 A crystalline See chlorinated phenols
solid and copper compounds
Copper Xanthate. Syn: Cu(C3H5OS2)2 Yellow precipitate
Copper Ethyl Xantho- mp: decomp
Genate.
Copper Zinc Chromate Variable in composition. HIGH. A recog
carcinogen
Copper Zinc Sulfate Toxic. See copper and
zinc compounds. Used as
a fungicide

COPPER: Environmental Dynamics and Human Exposure Issues


Page 37
Chapter 3
A BRIEF OVERVIEW OF THE PHYSICAL, CHEMICAL AND BIOLOGICAL PROPERTIES OF COPPER

COPPER: Environmental Dynamics and Human Exposure Issues


Page 38
Chapter 4
ENVIRONMENTAL RELEASES OF COPPER

4 ENVIRONMENTAL RELEASES OF COPPER


Human economic activities, involving the production and usage of copper and copper
compounds, as well as the consumption of materials (including food) that contain amounts of
copper, result in the re-distribution of copper in different environmental media. Information on
the magnitudes and trends of environmental copper releases is summarized in this section; the
review in ATSDR, 1990 has provided the basis for this summary. Additional information on
environmental releases as well as more recent information on evolving methods and databases
for assessing copper exposures can be found in the forthcoming report A Framework and Data
Sources for the Assessment of Exposures to Copper (Georgopoulos et al., 2002). More details
are provided on the website of the Center for Exposure and Risk Modeling (CERM), in the
section devoted to copper (http://www.CERM.org/copper).

4.1 Atmospheric Releases


Copper is emitted into the air from both natural and anthropogenic sources. Since copper is a
component of the earth's crust, the primary natural source of copper is windblown dust. Other
natural sources of emission, in order of importance are: volcanoes, decaying vegetation, forest
fires and sea spray (Davies & Bennett, 1985). Similarly, anthropogenic emission sources include:
nonferrous metal production, wood production, iron and steel production, waste incineration,
industrial applications, coal combustion, nonferrous metal mining, oil and gasoline combustion,
and phosphate fertilizer manufacture. In 1980, it was estimated that only 0.04% of copper
released to the environment is to air (Perwak et al., 1980). The EPA conducted a detailed study
of copper emissions into the atmosphere to estimate exposure (Weant, 1985). The sources of
emissions and the estimated quantities of copper emitted in 103 kg/yr were: primary copper
smelters, 43-6000 (2100, most probable value); copper and iron ore processing, 480-660; iron
and steel production, 112-240; combustion sources, 45-360; municipal incinerators, 3.3-270;
secondary copper smelters, 160; copper sulfate production, 45; gray iron foundries, 7.9; primary
lead smelting, 5.5-65; primary zinc smelting, 24-340; ferroalloy production, 1.9- 3.2; brass and
bronze production, 1.8-36; and carbon black production, 13. Using the most probable emission
value for primary copper smelters and the range for other sources, estimated United States
copper emissions are 2,959,000-4,300,000 kg annually. Daily stack emission rates have been
reported for three coal-burning power plants on a kg/day/1000 MW basis (Que et al., 1982).
They are 0.3-0.7 and 2.00 kg/day/1000 MW for those using low-sulfur western coal and high-
sulfur eastern coal, respectively. Emission factors in grams of copper released to the atmosphere
per ton of product have been estimated for various industries (Nriagu & Pacyna, 1988). These
factors would enable estimation of an industry's copper emissions from its production volume.
Missing from these emission estimates is fugitive dust arising from drilling, blasting, loading,
and transporting operations associated with copper mining. The only control of fugitive dust is
the manual use of water sprays. The amount of copper and other pollutants originating from a
waste site in wind blown dust is of some concern. In one study, the amount of airborne copper
and other heavy metals deposited near a large refuse dump that received municipal and industrial
waste and sewage sludge was determined by first measuring the amount of the metal
accumulated in moss bags. The deposition rate was then determined and compared with that for
an agricultural control area. The mean copper deposition rates in the two areas were about the

COPPER: Environmental Dynamics and Human Exposure Issues


Page 39
Chapter 4
ENVIRONMENTAL RELEASES OF COPPER

same; the maximum deposition rate was twice as much near the dump as in the control area
(Lodenius & Braunschweiler, 1986). Only in a few cases has the form of copper released into the
air been determined. In general, metals released into the atmosphere will be in particulate matter
in the form of an oxide, sulfate, or carbonate. Combustion processes are reported to release
copper into the atmosphere as the oxide, elemental copper, and adsorbed copper. Cupric oxide
has been identified in emissions from steel manufacturing and in fly ash from oil- fired power
plants and open-hearth steel mills (Perwak et al., 1980). Copper associated with fine particles (<
1 m) tends to result from combustion and other high-temperature sources, while that associated
with large particles (> 10 m) is likely to originate from wind blown soil and dust (Schroeder et
al., 1987).

4.2 Releases to Wastewater


As this section illustrates, there is extensive historical information on copper releases to
wastewater. An estimated 28,848 metric tons of copper entered waterways in the United States in
1976 (Perwak et al., 1980)*. This figure represents 2.4% of the identified releases of copper to
the environment. Much of this copper is associated with particulate matter. Copper is a natural
constituent of soil and will be transported into streams and waterways in runoff due either to
natural weathering or to disturbed soil. Sixty-eight percent of releases to water is estimated from
this source. Copper sulfate use represents 13% of release to water, and urban runoff contributes
2% (Perwak et al., 1980). In the absence of specific industrial sources, runoff is the major factor
contributing to elevated copper levels in river water (Nolte, 1988). In the EPA-sponsored
National Urban Runoff Program, in which 86 samples of runoff from 19 cities throughout the
United States were analyzed, copper was found in 96% of samples, at concentrations of 1-100
g/l (ppb) (Cole et al., 1984). Of the 71 priority pollutants analyzed for, copper, along with lead
and zinc, was the most frequently detected. The geometric mean copper concentration in runoff
water was 18.7 g/l. Domestic wastewater is the major anthropogenic source of copper in
waterways (Nriagu & Pacyna, 1988). Studies in Cincinnati and St. Louis showed discharges of
copper into sewer systems from residential areas to be significant, with an average loading of 42
mg/day/person (Perwak et al., 1980). Concentrations of copper in influents to 239 wastewater
treatment plants (12,351 observations) were 0.0001-36.5 ppb and the median value was ~0.4 ppb
(Minear et al., 1981a; Minear et al., 1981b). Inputs into Narraganset Bay, Rhode Island, in
decreasing order of importance, are: sewage effluent, rivers, urban runoff, and atmospheric
fallout (Santschi et al., 1984; Mills & Quinn, 1984). Ninety percent of both dissolved and
particulate copper was from sewage treatment plant effluent discharged into the Providence
River. The range of removal efficiencies reported for pilot and full scale sewage treatment plants
suggests that removal depends strongly on plant operation or influent characteristics. The best
data on typical publicly-owned treatment works (POTWs) using secondary treatment are that 55-
90% of copper is removed in these plants with a median and mean removal of 82% (Perwak et
al., 1980). By contrast, those plants using only primary treatment had a 37% median removal
efficiency. A more recent study focused on heavy metal removal in three POTWs that received
primarily municipal sewage and which used activated sludge as a secondary treatment. The study

*
It should be noted that ATSDR, 1990 erroneously reported this estimate as 28,848 million tons of copper.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 40
Chapter 4
ENVIRONMENTAL RELEASES OF COPPER

looked at removals in both the primary and secondary treatment stage. The average removal of
soluble copper and total copper after secondary treatment was 49-82% and 83-90%, respectively.
The average copper concentration in the final effluent was 17-102 ppb (Aulenbach et al., 1987;
Stephenson & Lester, 1987). Releases from these facilities contribute about 8% of all copper
released to water (Perwak et al., 1980).

Discharges to water from active mining and milling are small, and most of the western
operations do not release any water; water is a scarce resource and is recycled (Perwak et al.,
1980). Runoff from abandoned mines is estimated to contribute 314 million tons annually
(Perwak et al., 1980). These discharges are primarily insoluble silicates and sulfides and readily
settle out. Wastewater generated from mining operations comes from seepage, runoff from
tailing piles, or from utility water used for mine operation. The amount of wastewater generated
ranges from 0-300 liters of water/metric ton of ore mined for open pit copper mines and 8-4000
liters of water/metric ton of ore mined underground. Copper concentrations in wastewater from a
selected open pit and underground copper mine were 1.05 ppm and 0.87 ppm, respectively.
Discharges from electroplating operations are either directly to water or indirectly via POTWs.
Releases from copper-containing products may be substantial but difficult to predict. Corrosion
of copper in plumbing or construction may result in direct discharges or runoff into waterways.
Copper and brass production releases relatively little copper to water. Data regarding copper
concentrations in wastewater associated with selected concentrating, smelting, and refining
operations can be found in (PEDCo Environmental Inc., 1980). Results of an EPA industrial
effluent survey show that mean and maximum levels of copper in treated wastewater from six
industries exceeded 1 and 10 ppm, respectively. These industries and their mean and maximum
discharges in ppm are: inorganic chemicals manufacturing (<1.6, 18); aluminum forming (<160,
2200); porcelain enameling (1.3, 8.8); gum and wood chemicals (1.4, 3.0); nonferrous metals
manufacturing (1.4, 27.0); and paint and ink formulation (<1.0, 60.0). Emission factors in
nanograms of copper released per liter of water outflow have been estimated for various
industries. These factors would enable estimation of an industry's copper releases if the discharge
volume were known (Nriagu & Pacyna, 1988). Effluents from power plants that use copper
alloys in the heat exchangers of their cooling systems discharge copper into receiving waters
(Harrison, 1984a). The largest discharges occur after start-up and decrease rapidly thereafter. At
the Diablo Canyon Nuclear Power Station, a very high start-up discharge containing 7700 ppb of
copper fell to 67 ppb after 24 hours (Harrison, 1984a). During normal operation at two nuclear
power stations, copper levels ranged between 0.6 and 3.3 ppb. Except for after start-up of the
cooling system, most of the soluble copper (that which passes through a 0.45 m filter) discharged
was in bound forms (Harrison et al., 1980). During normal operation, 20% of the copper released
was in the <1000 molecular weight fraction, which contains the more available copper species.
Copper sulfate is directly added to lakes, reservoirs, and ponds for controlling algae. However,
the copper concentration in the water column generally returns to pretreatment levels within a
few days (Perwak et al., 1980; Effler et al., 1980). A potentially dangerous source of chemical
release at waste sites is leachate. Leachate from three municipal landfills in New Brunswick,
Canada, did not contain copper significantly above those in control samples (St-Cyr & Crowder,
1987). Concentrations of copper in landfill leachate are 0.1-1.0 ppm (Perwak et al., 1980).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 41
Chapter 4
ENVIRONMENTAL RELEASES OF COPPER

4.2.1 Copper in Cooling Systems


Copper alloys are widely used in cooling systems because of outstanding thermal conductivity,
ease of fabrication, ease of repair and excellent resistance to corrosion. This is true whether used
in marine applications, automobile radiators or nuclear power plant cooling systems. The
versatility of copper for cooling systems is indicated by the fact that copper and brass are the
materials most often used in the replacement of automobile radiators even when the original ones
were aluminum. Environmental concerns center around copper released into cooling water when
corrosion occurs. Corrosion can occur as a result of the chemical nature of the cooling water or it
can be induced by fouling organisms. Fouling organisms cause formation of biofilms that
produce anodic and cathodic areas on metal surfaces (Gaylarde, 1989).

Elevated levels of copper have been reported in cooling water from power plant cooling systems
in both open (Harrison, 1984a) and closed systems (Harrison, 1984b). In closed systems,
released copper is contained and can be controlled. In open or single-pass systems, corrosion
products of copper condenser tubing have been associated with biological effect. The release of
excessive copper, however, is often the result of improper procedures during and after periods of
shutdown. A classic example of this was the abalone kill that occurred during initial tests of the
copper alloy cooling system in the Pacific Gas and Electric nuclear power plant on the California
coast at Diablo Canyon (Martin et al., 1977). Approximately 1,500 dead abalone (Haliotis
rufescens and H. cracherodii) were found in the discharge area of the plant following testing of
the cooling system. The problem occurred when cooling system tests were conducted after a
period of shutdown. During the shutdown, untreated seawater was allowed to remain in contact
with copper-nickel piping in the condensing system. Copper dissociated from the piping at the
startup and the first pulse of discharge water produced a total copper concentration of 1.8 mg/l
in the cooling water. This would not have occurred if water had been continuously circulated
through the cooling system or if the seawater left in the system had received proper treatment.
Primarily as a result of this event, the copper piping in the main condenser system was removed
and replaced with titanium tubing, at a considerable cost to the company and, ultimately, to its
customers.

The biological impact of copper from cooling water depends on the concentration of biologically
available metal and the tolerance of the organisms in the receiving waters. In a study of the
biological impact of a nuclear power plant freshwater effluent, (Harrison & Lam, 1982) state that
...the labile copper released from the cooling system of the H.B. Robinson Nuclear Power
Station may be implicated in the increased deformities and reduced reproductive capacity found
in the bluegill population in the adjacent cooling lake. Phelps (Phelps, 1984) reports increased
tissue copper levels in oysters shortly after the 1976 startup of a nuclear power plant in
Chesapeake Bay. Although elevated tissue copper was found at all sites in the area when they
sampled in 1979, reduced levels were found in 1981 when average salinities were higher. It is
worthwhile noting that the highest metal concentrations were found in oysters from an
unaffected site, not in oysters near the power plant.

In the concluding remarks of a comparison of two open-cycle nuclear power plants located on
the ocean, Harrison et al. (Harrison et al., 1980) comment that ...impact on marine environments
of using copper alloys in condenser tubing of power stations depends on the operating conditions

COPPER: Environmental Dynamics and Human Exposure Issues


Page 42
Chapter 4
ENVIRONMENTAL RELEASES OF COPPER

at the site and the composition of the coolant waters. The operating conditions affect the
quantities of copper released and the composition of the water affects the physicochemical forms
of copper in the water. We expect little or no effect on marine biota when these stations are
operating normally. The impact will be determined by the maximum concentration and duration
of the copper pulse and on the number of copper-sensitive organisms in the discharge zone. Such
effects could be minimized if, after shutdown of the reactor, water was circulated continuously
through the cooling system or if, during start-up, natural organic chelators were added to the
water to reduce the concentration of labile copper.

From work on copper in effluent from freshwater-cooled nuclear power plants, Harrison
(Harrison, 1984a; Harrison, 1984b) reports that Under normal operating conditions, the
differences between influent and effluent waters were generally small, and most of the copper
was in bound. Lower (Lower, 1987) reports no significant offsite effects on trace metal
concentrations from cooling water.

4.2.2 Releases to Land


An estimated 97% of copper released into the environment is to land (Perwak et al., 1980). These
are primarily tailings and overburdens from copper mines and tailings from mills. The copper in
the tailings represents the portion of copper that could not be recovered from the ore and is
generally in the form of insoluble sulfides or silicates (Perwak et al., 1980). These wastes are
disposed of in mining states. Other releases to land include sludge from POTWs, a major source
of copper released to land (Nriagu & Pacyna, 1988), municipal refuse, and waste from
electroplating, iron and steel producers, and discarded copper products (e.g., plumbing, wiring)
that are not recycled. The copper content of municipal solid waste is ~0.16%; much of this will
be landfilled directly or as residues from incineration. Emission factors in milligrams of copper
released per gram of solid waste have been estimated for various industries. These factors would
enable estimation of an industry's copper releases in terms of quantity of solid waste discharged.
Agricultural products are believed to constitute 2% of the copper released to soil (Perwak et al.,
1980).

4.2.3 Databases for Environmental Copper Releases

TRI Summary Information for Copper


From 1987 to 1993, according to the Toxics Release Inventory, copper compound releases to
land and water totaled nearly 450 million lbs., of which nearly all was to land. These releases
were primarily from copper smelting industries. The largest releases occurred in Utah. The
largest direct releases to water occurred in Tennessee.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 43
Chapter 4
ENVIRONMENTAL RELEASES OF COPPER

Table 4: Copper Releases to Water and Land, 1987 to 1993 (in pounds)
Water Land
TOTALS 1,538,148 442,082,245
Top Ten States*
UT 55,350 153,501,500
NM 0 130,682,387
AZ 2,636 104,619,532
MI 19,763 11,172,897
NY 66,57 10,017,766
MT 0 8,696,153
TN 301,417 1,208,804
MO 250 1,486,000
AL 41,213 513,536
MD 78,601 270,945

Major Industries*
Primary copper smelting 7,591 201,214,264
Other nonferrous smelting 4,414 11,317,048
Plastic materials 44,422 9,637,850
Blast furnaces, steel 156,982 3,229,752
Poultry slaughtering 0 1,249,750
Copper rolling, drawing 17,253 941,075
Industrial organic chemicals 28,936 827,356
Prepared feeds, misc. 1,038 760,094
Industrial inorganic chemicals 220,503 527,458

*
Water/Land totals include only facilities with releases greater than a certain amount - usually 1000 to 10,000 lbs.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 44
Chapter 4
ENVIRONMENTAL RELEASES OF COPPER

Figure 6: Locations of Copper Mines in the U.S.


(Source: U.S. Geological Survey)

COPPER: Environmental Dynamics and Human Exposure Issues


Page 45
Chapter 4
ENVIRONMENTAL RELEASES OF COPPER

Environmental Releases of Copper by Location


(U.S. EPA Toxic Release Inventory 1987-1996)

Figure 7: Locations of Environmental Release of Copper in the United States

COPPER: Environmental Dynamics and Human Exposure Issues


Page 46
Chapter 4
ENVIRONMENTAL RELEASES OF COPPER

Environmental Releases of Copper by County


(U.S. EPA Toxic Release Inventory 1987-1997)

Figure 8: Environmental Releases of Copper in the United States (by County).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 47
Chapter 4
ENVIRONMENTAL RELEASES OF COPPER

COPPER: Environmental Dynamics and Human Exposure Issues


Page 48
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

5 ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

5.1 General Concepts


Although, as will be discussed in the next chapter, direct exposure of the general human
population to copper is typically dominated by the ingestion pathway (food and drinking water),
it is essential to understand and eventually quantify the processes affecting transport,
transformation and accumulation of copper in the various environmental and biological media. It
will then be possible to understand and predict the preferential accumulation of copper in the
food chain as well as to describe and explain both localized and regional/global trends in copper
distributions, of importance to specific population groups.

To provide a framework for building this understanding and for the systematic identification of
knowledge and data gaps, a modular conceptual model has been developed for the flows and
accumulation of copper in environmental and biological media. Figure 10 depicts a prototype
preliminary implementation of this model in the STELLA simulation language; a more
comprehensive implementation in MATLAB/FORTRAN, that also includes EPAs geochemical
speciation model MINTEQA2* is currently under development. Technical understanding of the
physical, chemical, and biological processes controlling the behavior of pollutants in the
environment has increased significantly in the past two decades. Many of the important advances
are reflected in the quantitative mathematical models now being used to describe the influences
of competing processes or reactions on the overall behavior of pollutants. A variety of
mathematical models encompassing years of research are now available for predicting the
behavior of pollutants in various environmental settings. The metal speciation model
MINTEQA2 is a versatile, state-of-the-art example of the equilibrium solution chemistry
programs now available.

*
MINTEQA2 is a geochemical equilibrium speciation model for dilute aqueous systems. The original MINTEQ was developed
at Battelle Pacific Northwest Laboratory (PNL) by combining the fundamental mathematical structure of MINEQL, a derivative
of REDEQL, with the well-developed thermodynamic database of the U.S. Geological Survey's WATEQ3 model. MINTEQA2 is
substantially different from the original MINTEQ in the features and options available, in the manner in which calculations are
implemented, and in its thermodynamic database. Also, MINTEQA2 is complemented by PRODEFA2, an interactive program
used to create input files. The original PRODEF also was a product of Battelle PNL and has undergone extensive modification
and development as PRODEFA2. The model can be used to calculate the equilibrium composition of dilute aqueous solutions in
the laboratory or in natural aqueous systems. It can be used to calculate the mass distribution between the dissolved, adsorbed,
and multiple solid phases under a variety of conditions including a gas phase with constant partial pressure. The data required to
predict the equilibrium composition consist of a chemical analysis of the sample to be modeled, giving total dissolved
concentrations for the components of interest and any other relevant invariant measurements for the system of interest, possibly
(but not necessarily) including pH, pe, or the partial pressures of one or more gases. A measured value of pH and/or pe may be
specified as equilibrium values or MINTEQA2 can calculate equilibrium values. Also, a mineral may be specified as presumed
present at equilibrium, but subject to dissolution if equilibrium conditions warrant, or definitely present at equilibrium and not
subject to complete dissolution. MINTEQA2 has an extensive thermodynamic database that is adequate for solving a broad range
of problems without need for additional user-supplied equilibrium constants. The standard database can be easily modified if it is
found to be incomplete or inadequate for a particular problem. The empirical nature of the available metal adsorption data reflects
the fact that natural adsorbent phases often occur as mixtures of impure amorphous substances that vary widely in chemical
behavior from site to site. For this reason, adsorption data are left to the discretion and problem-specific knowledge of the user.
Seven adsorption models are available in MINTEQA2 to match the type of data available for specific problems. The discussion
that follows attempts to bring together information that will be further utilized in the iterative implementation and refinement of
this model.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 49
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

It is not always possible to separate the environmental fate processes related to transport and
partitioning from those related to transformation and degradation for a metal, its various
compounds and complexes. Part of this problem is that the form of copper is often not identified
in field studies. It is also difficult to determine when a process such as adsorption should be
treated as partitioning or transformation, since the formation of strong bonds to an adsorbent may
be construed as a transformation to new molecular species. Separating weak and strong
adsorption is awkward and not always possible.

Copper and its compounds are naturally present in the earth's crust. Natural discharges to air and
water, such as windblown dust, volcanic eruptions, etc., may be significant. Therefore, it is
important to consider the background levels that are commonly found and distinguish these from
high levels that may be found as a result of anthropogenic activity. The physicochemical form of
copper is important in considering its behavior in the environment and availability to biota. For
example, the copper incorporated in mineral lattices is inert and unlikely to have ecological
significance. The analytical methods employed in most environmental studies do not distinguish
the form of copper present; it is known how much total copper is present, but not the nature of
the copper compounds or complexes present or how labile or available they are. The available
data indicate that the lability of copper varies considerably according to its environment. In
general, most copper in soil is in mineral form or tightly bound to organic matter. The median
concentration of copper in natural water is 4-10 ppb. It is predominantly in the copper(II) state.
Most of it is complexed or tightly bound to organic matter; little is present in the free (hydrated)
or readily exchangeable form. The combined processes of complexation, adsorption, and
precipitation control the level of free copper(II). The chemical conditions in most natural water
are such that, even at relatively high copper concentrations, these processes will reduce the free
copper(II) concentration to extremely low values. Sediment is an important sink and reservoir for
copper. In relatively clean sediment, the copper concentration is <50 ppm; polluted sediment
may contain several thousand ppm of copper. The form of copper in the sediment will also be
site-specific. Organics (humic substances) and iron oxides are the most important contributor to
binding of copper by aerobic sediments. However, in some cases, copper is predominantly
associated with carbonates. In anaerobic sediment, copper(II) will be reduced to copper(I) and
insoluble cuprous salts will be formed.

The largest release of copper by far will be to land, and the major sources of release are mining
operations, agriculture, solid waste, and sludge from POTWs. Mining and milling contribute the
most waste. However, these wastes will largely be in a mineral form that is unavailable to biota
and not likely to be transported or transformed. Most of the copper in soil is apparently tightly
bound to soil components and may not be available for uptake. While experiments show that
copper does not leach significantly from soil, levels of copper as high as 2.8 ppm have been
found in some groundwater (Page, 1981). Copper is released to water as a result of natural
weathering of soil and discharges from industries and sewage treatment plants; most of this
copper is attached to particulate matter. Copper compounds may also be intentionally applied to
water to kill algae. Of special concern is copper that gets into drinking water from the water
distribution system. When the system has not been flushed after a period of disuse, the
concentration of copper in tap water may exceed 1.3 ppm, the EPA drinking water limit.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 50
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

Copper is emitted into the air naturally from windblown dust, volcanoes, and anthropogenic
sources, the largest of which are being primary copper smelters and ore processing facilities. It is
associated with particulate matter. The mean concentration of copper in the atmosphere is 5-200
ng/m3.

The general population may be exposed to high concentrations of copper from drinking water
that has picked up copper from the distribution system (both from the water treatment plant and
in the home). Contact with copper may also result from using copper fungicides and algicides.
Many workers are exposed to copper in agriculture, industries connected with copper production,
metal plating, and other industries. At this time, copper has been identified at 210 out of 1177
National Priorities List (NPL) hazardous waste sites in the United States.

Table 5, Table 6 and Table 7 provide a summary compilation of information on the observed
levels of copper in various environmental and biological media; such information is essential in
the model development and implementation process, for testing purposes, and for providing
necessary initial and boundary conditions in the absence of more specific information for a given
application.

5.2 Atmospheric Dynamics of Copper


The primary source of anthropogenic emissions of copper into the atmosphere are due to copper
mining and smelting activities. The contribution of other anthropogenic sources such as the
combustion of fossil fuels, waste incineration, and the production of other metals is relatively
minor (Nriagu & Pacyna, 1988).

Changes in the rate of emission of copper into the atmosphere have been estimated based upon
an analysis of an approximately 3 km long ice core from Greenland (Hong et al., 1996a; Hong et
al., 1996b). The rate of anthropogenic emissions of copper from mining and smelting ranged
from 0.3 to 2.3 tons/year between 500 B.C. and 1850 A.D., which indicates that the total
emissions of copper into the atmosphere was approximately 2-fold that of the natural emission
rate. However, the industrial revolution brought a dramatic increase in the rate of copper
emissions, which jumped from 1.5 tons/year in 1850, to 16.2 tons/year in 1950. Since that time
the increase in emissions has been more gradual, and at present it is approximately 22,500
tons/year. In 1989, natural emissions of copper into the atmosphere are estimated to be
approximately 2,800 tons/year for interglacial conditions (Nriagu, 1988; Nriagu, 1989), and
therefore total atmospheric emissions copper are currently approximately 10-fold the natural rate.

Few data are available regarding the chemical forms of copper in the atmosphere and their
transformations. In the absence of specific information, it is generally assumed that metals of
anthropogenic origin, especially those from combustion sources, exist as oxides. Metallic species
are attacked by atmospheric oxidants, resulting in the formation of oxides. As these oxides age,
sulfatization may occur. In Arizona, atmospheric copper originating from smelters was strongly
correlated with sulfur (Schroeder et al., 1987).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 51
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

Copper is released to the atmosphere in the form of particulate matter or adsorbed to particulate
matter. It is removed through gravitational settling (bulk deposition), dry deposition (inertial
impaction characterized by a deposition velocity), washout by rain via attachment to droplets
within clouds (Schroeder et al., 1987), and rainout (scrubbing action below clouds (Schroeder et
al., 1987). Removal rate and distance traveled from the source will depend on source
characteristics, particle size, and wind velocity. Gravitational settling governs the removal of
large particles (> 5 m), whereas smaller particles are removed by the other forms of dry and wet
deposition. The importance of wet to dry deposition generally increases with decreasing particle
size. The scavenging ratio (ratio of the copper concentration in precipitation [ppm] to its air
concentration [g/m3]) for large particles displays a seasonal dependence that reflects their more
effective scavenging by snow than by rain (Chan et al., 1986). Copper from combustion sources
is associated with sub-micron particles. These particles remain in the troposphere for an
estimated 7-30 days. In that time, some copper may be carried far from its source (Perwak et al.,
1980). Metal deposition is characterized by large temporal and spatial variability. Estimated
copper deposition rates in urban areas are 0.119 and 0.164 kg/ha/yr for dry and wet deposition,
respectively (Schroeder et al., 1987). Bulk deposition reportedly ranges from 0.002-3.01
kg/ha/yr. For rural areas, bulk deposition reportedly ranges from 0.018-0.5 kg/ha/yr, and wet
deposition 0.033 kg/ha/yr. The washout ratio is 114,000-612,000 g/m3 rain/g/m3 air [(140-751
g/kg rain)/(g/kg air)]. In southern Ontario, Canada, where the average concentration of copper
in rain was 1.57 ppb during 1982, 1.36 mg of copper was deposited annually per square meter as
a result of wet deposition (Chan et al., 1986). For central and northern Ontario, the mean
concentrations of copper in rain were 1.36 and 1.58 ppb, respectively, and the annual wet
depositions averaged 1.13 mg/m2 in both instances.

Available data on atmospheric concentrations of copper are questionable because of possible


contamination from the high-volume sampler motors. To derive the value for the atmospheric
copper pool, a uniform airborne copper level of 0.5 ng/m3 has been assumed on the basis of
Nriagus compilation of ambient copper levels at remote locations. The size of this copper pool
may have been overestimated since the concentration has been assumed to be independent of
altitude. In fact, several studies have shown that the metal aerosol concentrations generally
decrease as altitude increases. The total flux of copper to the atmosphere was estimated to be 75
E6 kg/year, about 75% or 56 E6 kg from anthropogenic sources.

From the atmospheric copper pool and the global emission estimate, Nriagu (Nriagu, 1979d)
estimated the residence time of copper in the air to be (2.6 E9)/(75 E9) years or about 13 days.
This value is somewhat higher than the 2 to 10 days generally regarded as typical for airborne
particulates in the unpolluted troposphere.

5.3 Hydrospheric Dynamics of Copper

5.3.1 General Discussion


The hydrosphere consists of environments dominated by water, and is comprised of marine,
estuarine, river, and other freshwater environments. In each of these environments, copper is

COPPER: Environmental Dynamics and Human Exposure Issues


Page 52
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

present in the dissolved (< 0.4 m in diameter), suspended particulate, or sedimentary phases.
Only a fraction of the copper in the dissolved phase exists as free or hydrated Cu2+, the balance
being complexed with organic or inorganic ligands, or associated with a colloidal phase (> 3000
Da), comprised of microparticles and macromolecules too small to settle under gravitational
force.

It is generally agreed that the toxicity of dissolved copper with respect to aquatic biota is
dependent upon the concentration of free Cu2+ ion. In addition, evidence from studies in which
pH, and therefore the relative proportions of dissolved copper species differed, suggests that
inorganic complexes of copper, such as hydroxy and carbonato complexes may also be involved
in copper toxicity (Allen & Hansen, 1996). Dissolved copper bound to organic ligands however,
are not thought to be involved in the toxicity of copper. At constant pH, the ratio of the
concentrations of free copper ion and inorganic complexes is usually constant (typically around
2.5%), because the inorganic ions with which copper forms complexes are usually in
stoichiometric excess or are buffered, and are therefore relatively constant. Organic ligands form
strong complexes with copper but are generally present in concentrations only slightly higher
than copper. Therefore, their capacity to buffer the concentration of free copper ion is limited,
resulting in a nonlinear relationship between dissolved copper and free copper ion. An important
implication of this nonlinearity is that although the concentration of both dissolved copper and
organic matter in a body of water are increased by the introduction of effluent, the concentration
of Cu2+ may actually decrease. A quantitative framework for characterizing the bioavailability of
copper to aquatic biota is provided by the Biotic Ligand Model or BLM (DiToro et al., 2000).

The capacity of organic ligands in natural marine water to bind copper can range from 0.5 to 35
g/l, depending upon the amount of ligand present, and its binding constant. Discharge of
effluent enriched in both copper and organic ligands could result in either increased or decreased
copper ion concentrations, depending upon the extent to which the additional organic ligand in
the effluent can buffer the copper.

5.3.2 Fluxes of Copper in the Hydrosphere


Copper present in the hydrosphere comes from several types of sources. Lewis (Lewis, 1995b)
summarized them as follows:

Minerals in soil and soil parent material (e.g. weathered rock) that form sediments and
suspended particles in water. Copper is a common element and occurs naturally in rocks
and soils either as native copper or as a mineral. Weathering, physical disintegration and
chemical decomposition of exposed parent material forms soils and causes the release of
soluble copper and copper-containing compounds to water. River transport provides a
means of delivering this, as well as particles containing copper, to lakes and coastal
marine waters. Although parent material is an important source of micronutrients for
soils, copper mostly occurs as a minor component except in areas of mineralization.
Extraction of copper from soil parent material into a dissolved state. This process can
occur by water percolating through soil or from particles after they are introduced into an
aquatic system. It can occur when an oxidized form of copper is on the surface of the

COPPER: Environmental Dynamics and Human Exposure Issues


Page 53
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

particles, but not when it is in the interior matrix, less accessible to oxidation and
dissolution.
Biological particles, including both living and dead organic material and products of
organism decay. Organisms acquire copper to satisfy their metabolic requirements. In
doing so they remove copper from the water or sediments where they live. When
organisms die, they form copper-containing particles which ultimately contribute to the
mineral and organic load of the sediments. Suspended particles can scavenge (e.g.
plankton) or chemically bind (e.g. organic matter) copper from the water and initially
immobilize it in bottom sediments. However, this biological removal from the water is
subject to remobilization by chemical or biological activity within sediments.
Hydrothermal systems in which heated or chemically altered water are found.
Hydrothermal sites occur infrequently. They are sulfide-rich areas that have been
responsible for the formation of porphyry copper deposits over geologic time and are
commercially exploited today. Metal-rich sediment associated with hot brine from the
Atlantis II deep in the Red Sea is reported to contain 500 ppm of copper depletion in the
water around active hydrothermal sites. This is believed due to precipitation of
copper-containing sulfide minerals. Boyle (Boyle, 1979) points out that excess copper
found at hydrothermal locations must come primarily from a seawater source rather than
just leaching from rocks.
Copper input from sediments introduced into the overlying water column from benthic
sediments (bottom sediments). Geochemical conditions affecting this input vary widely.
As a result, there are inherent difficulties in providing accurate estimates of input. One
estimate (Nriagu, 1979d) of input of regenerated or refluxed copper for the entire
ocean is 4.7 x 108 kg/yr.
Anthropogenic inputs, either directly into the water or leached after deposition on land.
These include industrial and municipal effluents as well as antifouling coatings, pesticide
residues, manure and sludges.
A major source of aerosol copper is anthropogenic, from industries like power plants that
use fossil fuels. Other sources include windblown dust and particles injected into the
atmosphere by volcanic activity. Aerosol deposition is most pronounced in the Northern
Hemisphere, a result of the greater concentration of industrial activity there compared
with the Southern Hemisphere (Soudine, 1989).

Using a simplified conceptual model, Nriagu (Nriagu, 1979b) calculated copper fallout into the
oceans, assuming a dry deposition velocity of 0.4 cm/sec, an average atmospheric copper
concentration of 0.5 ng/m3, and an oceanic surface area of 3.6 E18 cm2. The calculated dry
deposition rate (6.34 E9 g) was then multiplied by two to convert it into a total deposition
(including rainfall) of 13 E9 g over the oceans. This estimate suggests that about 20% of the
copper released each year into the atmosphere is deposited in the oceans. (Weiss et al., 1975)
determined the post 1945 rate of copper deposition in the Greenland glacier to be 19 to 39
ng/cm2 year. If these data are extrapolated to the world oceans, the resulting annual atmospheric
deposition of copper is found to be 68 to 140 E9 g; these figures, in fact, exceed the total annual
emission of copper to the atmosphere. Nriagus value for the atmospheric fallout of copper over
the oceans, however, is roughly twice the 5 E9 g/year given in a National Academy of Sciences

COPPER: Environmental Dynamics and Human Exposure Issues


Page 54
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

(1975) report. Clearly, all these estimates should only be interpreted as preliminary attempts to
quantify attributes of a very complex system; the specific values of such estimates reported in
this and other chapters of the present document should therefore also be seen from the
perspective of an exploratory analysis.

The following is a list of estimates of annual global copper fluxes through the aquatic
environment (Martin & Windom, 1991):

Input to ocean margins (both natural and anthropogenic copper).

Soluble atmospheric flux 15 x 106 kg/yr


Riverine
Dissolved 58 x 106 kg/yr
Particulate 1500 x 106 kg/yr
Biological removal in margins 3 x 106 kg/yr

Input to open ocean

Output from margins to open ocean 145 x 106 kg/yr


(Estimated total retention in coastal margins 1428 x 106 kg/yr)
Atmospheric input to open ocean (natural and anthropogenic) 7 x 106 kg/yr
Total removal in open ocean 220 x 1106 kg/yr

The following residence times in years:

Ocean margins 6 x 102 years


Deep ocean 1.5 x 103 years

were calculated using the average composition of deep-sea clay and a deep-sea clay
accumulation rate of 1.1 x 1012 kg/yr.

Several comments need to be made about these estimates of copper fluxes:

Estimates of riverine fluxes from land are based on particulate and dissolved metal
concentrations and river flow rate.

Estimates of removal of riverine sediment inputs in the ocean's margins range from 80%
to 95%. This would include particulate copper that remained in the sediment as well as
any dissolved copper that became associated with the sediment in the river estuary.

The atmospheric inputs to the oceans in Martin and Windom (Martin & Windom, 1991)
are from the 1989 World Meteorological Organization report by the Joint Group of
Experts on the Scientific Aspects of Marine Pollution (Joint Group of Experts on the
Scientific Aspects of Marine Pollution (GESAMP), 1989). This report gives an estimate
of 52 x 106 kg/yr for the maximum total atmospheric flux of copper to the world's oceans.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 55
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

The soluble flux (dissolvable in seawater) global value for copper is given as 30 x 106
kg/yr or twice the input into coastal waters.

Nriagu (Nriagu, 1979d) estimated the total copper reservoir in lakes and streams using an
average copper concentration of 2 g/l. According to Nriagu, this value corresponds to the
average concentration in small Norwegian lakes (Henriksen & Wright, 1978) and in the Great
Lakes (unpublished results by the author). By comparison, the copper burden of glaciers is 3.3 E
12 g, assuming an average copper concentration in ice of 0.2 g/kg (Silver et al., 1975).

From the freshwater biomass value of 57 E12 g and mean copper content of 5 g/g the copper
burden is calculated to be about 3 E8 g. This comprises about 0.5% of the total copper in lakes
and rivers. The total net productivity for lakes and rivers accounts for about 1.0 E15 g of dry
matter. If the mean copper concentration in living biomass is also assumed to be 5 g/g, the
amount of copper cycling each year through freshwater biota is estimated to be 5 E9 g; this
represents about 10% of the freshwater copper pool. It follows that the residence time for copper
in the freshwater biota is 22 days.

Nriagu (Nriagu, 1979d) was not able to assess the sinks for copper in freshwater ecosystems.
Basically, lakes and rivers are transient features, and it seems unlikely that freshwater sediments
play an important role in the global copper economy.

Gibbs (Gibbs, 1977) value (6.3 E12 g/year) for copper in river runoff was adopted by Nriagu
(Nriagu, 1979d). According to Gibbs, only 1% (6.1 E10 g) of the copper transported by rivers to
the oceans is in soluble form. Of the solid copper load in rivers, it was shown that 85% was
associated with particulate crystalline phases, 5.7% was bound to metal hydroxide coatings on
particles, 4.5% was associated with organic material, and only 3.5% was adsorbed onto
suspended particulates. Gibbs (1977) also compared the concentration of copper in the
transported material to the average crustal copper abundance and found a high ratio (9 to 13),
pointing to continental depletion of copper. This observation is consistent with Bowens (Bowen,
1966) data, showing that the average copper content of soils (20 g/g) is less than the average
crustal value. The ratio of average copper concentration in the crust to that in pelagic sediments
ranges from about 3 (Atlantic clays) to over 20 (North Pacific clays), further indicating
continental copper depletion. High geochemical copper mobility at earth surface conditions
presumably has enabled erosional removal of copper from the continents.

The oceanic copper pool (about 2 E14 g) was estimated to be less than that stored in sediment
pore waters (5 E15 g). All reported oceanic copper profiles show a steady increase in
concentration down to the seafloor. This suggests substantial regeneration of copper from
seafloor sediments, where the copper concentration in the pore waters has been found to be about
15 g/kg (Boyle, 1979). Boyle also estimates the flux of copper from the sediments to the
bottom waters to be 0.13 g/cm2-year, a value that corresponds to 4.7 E11 g/year for the entire
ocean. Although copper regenerated from the sediments amounts to only 7.5% of total riverine
input, it is about an order of magnitude higher than copper delivered to the oceans in soluble
form. Thus sediments must be regarded as an important source of available copper in the deep
oceans.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 56
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

The annual input into oceans of dissolved copper derived from antifouling paints has been
estimated at 2.1 E10 g (National Academy of Sciences, 1975). The same report also estimates
industrial discharges of copper into freshwater environments to be 1.4 E10 g/year. The amounts
of industrial wastes and sewage sludge dumped into the ocean by the United States have been
given as 3.4 E12 and 5.0 E12 g/year, respectively. Worldwide waste discharges into the oceans
may be assumed to be three times those of the United States, or, 25 E12 g/year. If the average
copper content of sewage sludge and industrial wastes is taken as 670 g/g, copper input
associated with waste discharge is 1.7 E 10 g/year, a value comparable with ombrogenic input of
copper to the oceans.

The deep ocean residence time for copper was calculated by Nriagu (Nriagu, 1979a) to be only
1500 years, considerably shorter than times reported in the literature, such as 20,000 (Goldberg,
1965), and 4100 years (Bewers & Yeats, 1977). The major cause of this discrepancy has to be
the use of higher average oceanic concentrations in previous estimates.

The total amounts of copper in living marine biomasses were estimated from biomass values of 2
E14 and 3 E15g for plants and consumers, respectively. On the basis of the compilation in the
two volumes edited by Nriagu (Nriagu, 1979d) the average copper contents of marine plants and
organisms were estimated to be 3.5 and 1.0 g/g dry weight, respectively. Thus the copper
burdens of living plants and organisms in the oceans are 7 E8 and 3 E9 g, respectively. The
particulate organic carbon content of the oceans has been given as 3 E16 g. Taking the C-O-N-S-
P ratio in the particulate organic matter as 106-106-16-2-1 gives a value of 7 E16 g for oceanic
mass particulate organic matter. If the copper content of particulate organic matter is assumed to
be 7 g/g, the copper burden in this reservoir is 50 E10 g. From the dissolved organic pool of 1.6
E 18 g and the C-O-N-S-P ratio given above, the mass of oceanic dissolved organic matter is
estimated to be 3.5 E18 g. Since little information is currently available on dissolved organic
matter copper concentration, the size of this copper pool cannot be estimated.

Total primary oceanic production may be assumed to be 6 E16g/year dry matter. If the same
copper concentration as in living marine plants (i.e., 3.5 g/g) is assumed, annual consumption
of copper by marine biota is 21 E10 g. Thus the residence time for copper in marine biota is (3.7
E9)/(2 1 E10) years, or about 1 week. The rapid rate of turnover is not surprising since copper is
essential to all life forms.

Oceanic sinks for copper remain essentially intractable. To a first approximation, the flux of
copper from ocean to sediment should equal the sum of inputs from the atmosphere (1.3 E10 g),
waste disposal (1.7 E10 g), and the river-suspended load (624 E10 g), that is, about 6.3 E12
g/year. The flux rate for copper and other trace metals obtained using the marine sedimentation
rate was only 35% of the rate obtained on the basis of the stream supply.

5.3.3 Chemistry of Copper in the Hydrosphere


The chemistry of copper affects its distribution and final fate after release into aquatic
ecosystems, and its availability and potential toxicity to biota. The physicochemical form of
copper is of major importance in determining the nature of its interaction with chemicals and
particles and its role in the sedimentary cycle.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 57
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

Copper occurs in one of four oxidation states with copper(I) and copper(II) being the most
common. Copper(I) is found primarily in anaerobic conditions and is readily oxidized to
copper(II). Ionic copper is chemically labile, reacting with a variety of inorganic and organic
ligands, as well as with many types of particulates. As a result, the element is found in soluble,
colloidal and particulate forms. Assessment of potential biological effects of anthropogenic
copper must consider metal concentration, speciation, and the changes in its chemistry that will
occur after the metal is introduced into the environment. Although mass budgets of copper may
be of interest in determining elemental loading, the values must not be applied directly to the
determination of biological effects.

The biological availability of copper is affected by natural processes, particularly


complexation/chelation and sorption/desorption (Harrison, 1985). The picture is further
complicated by biological processes such as changes in metal speciation of organically bound
metals caused by bacterial degradation. These processes must be taken into account when
considering the nature of copper in natural environments and when attempting to apply
geochemical models to predict the effects of anthropogenic copper. Often, organisms are used to
assay water quality and the biological effect or availability of metals in aquatic systems. Cairns
(Cairns, 1990a; Cairns, 1990b) reviews the development of biomonitoring in aquatic ecosystems,
commenting on the importance of multispecies, rather than single species, testing.

Inorganic and Organic Ligands


Copper forms complexes with hard bases (ammonia, carbonate, chloride, hydroxide, nitrate and
sulfate). It reacts strongly with sulfur (to form insoluble sulfides) and with sulfur-containing
ligands. Labile copper species include ions, ion pairs, readily dissociable inorganic and organic
complexes, and easily exchangeable copper sorbed on colloidal inorganic or organic matter.
Copper forms stable cuproorganic complexes with a number of organic anions, amino acids,
amino sugars, alcohol, urea, etc.). It may also bind to high-molecular-weight organic material
and occur on or in colloids. Much of the high-molecular-weight material is termed humic
substance, a diverse group of refractory organics derived from the degradation of particulate
organic materials (primarily plant material) in terrestrial and aquatic environments.

Since labile dissolved copper readily associates with a variety of organic and inorganic
substances, its chemistry is related to the behavior of dissolved and particulate inorganic and
organic constituents of water. In seawater and in many freshwater systems, copper speciation is
now recognized to be dominated by complexation with naturally-occurring organic ligands. The
extent of complexation can vary widely. Examples of the importance of naturally-occurring
organic ligands in metal speciation are numerous and include (Lewis, 1995a):

A study of the Tamar Estuary (U.K.) in which Van den Berg et al. (Van den Berg et al.,
1990) found very low cupric ion activities as a result of the complexation of copper by
organic ligands. Van den Berg (Van den Berg, 1992) comments that copper occurs fully
complexed by organic material in the North Sea.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 58
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

A report by Gardner and Ravenscroft (Gardner & Ravenscroft, 1991) that in the Tweed
Estuary in England there was a substantial excess of complexing capacity over total
dissolved copper across the salinity range that was measured. They also report changes in
the copper affinity of ligands in different regions of the estuary.

In North Pacific waters, Coale and Bruland (Coale & Bruland, 1988) found over 99% of
the total dissolved copper associated with organic ligands at depths less than 200 m.

Oganesyan and Babayan (Oganesyan & Babayan, 1987) found that in the surface waters
of the Sevan Lake Basin of Armenia, approximately 90% of the copper was complexed,
40% of it in compounds with a molecular weight of 9,000 or more (translation).

Bugenyi and Lutalo-Bosa (Bugenyi & Lutalo Bosa, 1990) examined possible effects of
excess copper from a copper mining region in the Western Rift Valley of East Africa. In
a saline lake, they found that detrimental effects were reduced by both organic and
inorganic agents.

Organic metal-complexing ligands are produced as products and byproducts of metabolism as


well as of the breakdown of biological material. The chemical properties of the ligands can
change as a result of changes in the nature of the organisms producing them. A variety of
naturally-occurring agents collectively classified simply as dissolved organic matter (DOM)
include humic substances that have demonstrated capacity to affect both copper toxicity and
copper uptake by aquatic organisms. Humic substances which complex copper and other cations
include both humic and fulvic acids. Recent work suggests that humic acids (especially the
higher-molecular-weight fractions) have an enhanced ability to bind metal ions at higher pH as a
result of their high surface potential.

Particles can provide a major source of copper in rivers, lakes and oceans, a source which can
vary seasonally and over longer periods. Since many types of particles can adsorb metallic ions,
the concentration of copper in particulate form has been used as an indicator of the pollution
status of a body of water. Copper reactions with particles are controlled by processes including

sorption,
desorption,
complexation (including chelation) and
coprecipitation.

Since organisms are particles, this list can also include processes associated with copper uptake
by organisms. From a physical/chemical standpoint, the surface of a particle can be complicated
and capable of interacting with a number of metallic ions as well as with other components in
water or sediment. The dynamics of the processes occurring on the surface can also change, for
example, when the particle carries a coating of organic matter. A great deal is still to be learned
about the factors that affect the uptake and exchange of metallic ions by particles. The lack of
knowledge is unfortunate since particles not only play major roles in metal transport but can also
serve as metal sinks. Chemical models do exist that attempt to address some of the problems

COPPER: Environmental Dynamics and Human Exposure Issues


Page 59
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

associated with the scavenging of metal ions by marine particulate matter (Clegg & Sarmiento,
1989).

The movement of particles in aquatic environments is determined by their size and density and
by the speed and turbulence of water currents. As a result, particulate copper may be deposited in
bed-load sediments in the immediate vicinity of a source or carried into adjacent environments.
Settling particles can account for a major flux of metals to the sediments in both freshwater and
saltwater environments. Chemical factors, such as dissolved oxygen, or physical factors, such as
the overturning of lakes in winter, will affect sedimentation rates and the fate of the copper on or
in the particles.

Effect of acidic deposition


Acidic deposition not only causes corrosion but reduces the ability of upland soils and lake
sediments to retain atmospherically deposited trace metals. Although the degree of acid rain
effect is open to debate, certainly at low levels of acidification there is evidence a decrease of pH
over a range of approximately >5 to 2.5 increases leaching of metal from soils and other
substrates, which increases soluble metal concentration in runoff water.

Behavior of copper in estuaries


Changes occur in both the concentration and chemical nature of riverine copper when it is
introduced into salt water in estuaries. From a physical perspective, the speed of a river decreases
when it enters an estuary. This allows copper-containing particles to settle to the bottom. From a
chemical perspective, river water acquires major ions when it mixes with salt water in an estuary.
These ions can have an important effect on the chemistry of copper. They react with both
particulate and what is termed dissolved (< 0.4 pm) material (e.g. humic substances) in rivers.
The flocculation of humic substances that occurs in estuaries is a result of this.

Copper in a filterable state is transformed into particulates, which often settle to the bottom. This
process results in the loss of dissolved metal and metal-containing material within estuaries (e.g.
Sholkovitz, 1980). The net effect is an increase in particulate copper and a decrease in dissolved
copper within estuaries. Sedimentation of both inorganic and organic particles retains riverine
copper in coastal sediments. Estuaries and coastal waters thus act as trace metal sinks or traps
for copper (Martin & Windom, 1991).

5.3.4 Drainage from Urban Areas


Elevated levels of copper have been reported in drainage from urban areas and roadways. As a
result, urban catchment areas and detention ponds can be enriched with copper as well as with
other metallic elements. Much of the copper will be associated with fine grained sediments.
Mesuere and Fish (Mesuere & Fish, 1989), for example, note that copper was the dominant
metallic element in a detention pond system in a parking lot near Portland, Oregon. They report
that Copper was found to be deposited in the pond sediments in a small but highly concentrated

COPPER: Environmental Dynamics and Human Exposure Issues


Page 60
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

plume (up to 130 mg kg-1), extending axially from the runoff inlet pipe. (Note: With the
banning of asbestos in automotive brake linings, copper and brass are extensively used in new
formulations. Copper is also an additive in some lubricating oils and diesel fuels.) Detention
ponds can be a useful management practice for controlling runoff from parking lot areas,
although water from detention ponds and runoff from storm drains is a source of anthropogenic
metal.

5.3.5 Copper in Groundwater


Copper enters surface and groundwater from natural and anthropogenic sources. The latter
include leachate from landfills, chemical waste sites, detention ponds and deep-well injection.
The chemistry of leachate and the degree of complexation of copper differs between the various
types of sources and also among sites. This provides a range of copper forms and species and
makes it difficult to model the chemical reactions and transport of copper in the aquatic
environment. The predictive ability of groundwater flow and transport models is limited by
uncertainties in estimating some of the field parameters. Recent work includes models to
simulate the effect of sediments in highway detention ponds on metal uptake and introduction
into ground water. Evidence suggests that detention ponds can operate for many years before
they become saturated and create the potential for groundwater contamination.

5.3.6 Mathematical Modeling of Metal Speciation


Estimates of the geochemical and biological impacts of copper require an understanding of metal
speciation. Because of the complexity of natural environments, no single generalization about the
forms of copper that might be expected will be accurate. With certain assumptions, equilibrium
speciation models can help to resolve ambiguities concerning metal partitioning in natural
environments (Campbell et al., 1988). The results of mathematical models of metal partitioning
can be difficult to apply to natural systems, however, for a number of reasons, including
problems with multiple substrate systems and localized chemical variability. A number of
speciation models have been constructed using available information about pH, inorganic cations
and anions and organic ligands. Some of the models include particulates; others relate to metals
in sediments; most provide an indication of the fate of labile metal species, which does relate to
part of the biologically available metal load, although the models do not include lipid-soluble
metal species, which may be biologically important.

The copper(I) ion is unstable in aqueous solution, tending to disproportionate to copper(II) and
copper metal unless a stabilizing ligand is present (Callahan et al., 1979). The only cuprous
compounds stable in water are insoluble ones such as Cu2S, CuCN, and CuF. In its copper(II)
state, copper forms coordination compounds or complexes with both inorganic and organic
ligands. Ammonia and chloride ions are examples of species that form stable ligands with
copper. Copper also forms stable complexes with organic ligands such as humic acids, binding to
-NH2 and -SH groups and, to a lesser extent, to -OH groups. Natural waters contain varying
amounts of inorganic and organic species; this affects the complexing and binding capacity of
the water and the types of complexes formed. In seawater, organic matter is generally the most
important complexing agent (Coale & Bruland, 1988). The formation of ligands may affect other

COPPER: Environmental Dynamics and Human Exposure Issues


Page 61
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

physicochemical processes such as adsorption, precipitation, and oxidation-reduction in water


(Callahan et al., 1979).

The major species of soluble copper found in freshwater, seawater, and a combination of the two
over a range of pHs is Cu2+, Cu(HCO3)+, and Cu(OH)2 (Long & Angino, 1977). At the pH values
and carbonate concentrations characteristic of natural waters, most dissolved copper(II) exists as
carbonate complexes rather than as free (hydrated) cupric ions (Stiff, 1971a; Stiff, 1971b).

Dissolved copper concentration depends on factors such as pH, the oxidation-reduction potential
of the water, and the presence of competing cations (Ca2+, Fe2+, Mg2+, etc.), anions of insoluble
cupric salts (OH-, S2-, PO43-, Co32-), and organic and inorganic complexing agents. If the
concentration of a particular anion is high enough to exceed the solubility of a copper salt,
precipitation of that salt will occur. The most significant precipitate formed in natural waters is
malachite [copper 2(OH) 2CO3]; other important precipitates are copper (OH)2 (and ultimately
CuO), and azurite [copper 3(OH) 2(CO3)2]. In anaerobic waters, copper 2S, copper 2O, and
metallic copper forms and settles out (Callahan et al., 1979). The combined processes of
complexation, adsorption, and precipitation control the level of free copper(II). The chemical
conditions in most natural water are such that, even at relatively large copper concentrations,
these processes will reduce the free copper(II) concentration to extremely low values.

As a result of all the aforementioned physico-chemical processes, copper in water may be


dissolved or associated with colloidal or particulate matter. Copper in particulate form includes
precipitates, insoluble organic complexes, and copper adsorbed to clay and other mineral solids.
In a survey of nine rivers in the United Kingdom, 43-88% of the copper was in the particulate
fraction (Stiff, 1971b). A study using suspended solids from the Flint River in Michigan found
that the fraction of adsorbed copper increased sharply with pH, reaching a maximum at a pH of
5.5-7.5 (McIlroy et al., 1986). The colloidal fraction may include hydroxides and complexes with
amino acids. The soluble fraction is usually defined as that which will pass through a 0.45 mm
filter; it includes free copper and soluble complexes, as well as fine particulates and colloids.

The importance of copper's association with inorganic and organic ligands will vary depending
on the pH and concentration of competing ligands in the body of water. In river water from the
northwestern United States that had a relatively high pH (7.0-8.5) and alkalinity (24-219 ppm as
CaCO3), inorganic species like CO32- and OH- were the most important ligands at high copper
concentrations. However, other species were important at low copper concentrations. On the
other hand, samples from lakes and rivers in southern Maine with a relatively low pH (4.6-6.3)
and alkalinity (1-30 ppm as CaCO3) were largely associated with organic matter. After a period
of rain in southeastern New Hampshire, inorganic constituents contributed more to the copper
binding in lakes and rivers than dissolved organic matter did (Truitt & Weber, 1981a; Truitt &
Weber, 1981b). Runoff induced by the rain had added to the inorganic load of the rivers and
lakes, as was evidenced by their pH (5.7-7.4) and alkalinity (1.7-43.4 ppm as CaCO3). A green
precipitate, confirmed to be malachite [copper 2(OH) 2CO3] formed in river water in Exeter; this
water had the highest pH and alkalinity. A computer simulation of the copper species in pond
water and artesian well water that fed the pond predicted that 98% of the copper in the artesian

COPPER: Environmental Dynamics and Human Exposure Issues


Page 62
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

well water would be bound to organic matter, whereas 88% and 63% of the copper in pond water
would be bound to organics in spring and fall, respectively (Giesy et al., 1983).

5.3.7 Copper in Seawater


Seawater samples obtained in a transect of uppermost Narragansett Bay in August 1980 were
analyzed for dissolved, particulate, and organically-bound copper to investigate the geochemistry
of copper-organic complexes (Mills & Quinn, 1984). Narragansett Bay is a partly mixed estuary
in Massachusetts and Rhode Island that receives organic matter and metals from rivers,
municipal and industrial effluents, and runoff. Dissolved copper represented 60% of the total
copper and ranged from 16.4 g/kg in the Providence River to 0.23 g/kg in Rhode Island
Sound. Analysis of the data indicated that ~75% of this copper is removed within the Providence
River. Particulate copper concentrations ranged from 2.42-0.06 g/kg and generally comprised
40% of the total copper. 14% to 70% (0.12-2.30 g/kg) of the dissolved copper was complexed
with organic matter.

Organic ligands may contain a variety of binding sites and the strength of the resulting copper
complexes will vary accordingly (see, e.g., Meyer et al., 1998). Over 99.7% of the total dissolved
copper in surface ocean water from the northeast Pacific was associated with organic ligands
(Coale & Bruland, 1988). The dominant organic complex, limited to surface water, was a strong
ligand of biogenic origin. A second, weaker class of organic ligand was of geologic origin. An
independent study showed that the copper binds to humic material at a number of sites; the
binding strength of the sites varied by 2 orders of magnitude (Giesy et al., 1986). The humic
material in the study was derived from nine surface waters in the southeastern United States.
Soluble copper in water discharged from a nuclear power station was primarily complexed with
organic matter in the 1000-100,000 molecular weight range (Harrison et al., 1980). 10% to 75%
of the discharged copper was in particulate form.

High pH decreases copper adsorption (Kester et al., 1975). It is important to specify the pH when
attempting to predict adsorption-desorption in geochemical systems. In aqueous systems, such as
the mouth of rivers, the adsorbed metals will desorb as the salinity rises, if the pH is kept
constant. Both salinity gradient values and pH gradient values are needed to accurately predict
the behavior of the adsorbed metals as they pass over the gradients.

5.3.8 Copper in Sediments


Many organic and inorganic pollutants have been released into both freshwater and marine
environments from industrial and sewage treatment facilities, as well as through atmospheric
cycling of particulate matter from power plants and vehicular emissions. During the past two
decades there has been an increased interest in understanding the characteristics and processes
associated with certain heavy metals emission and transport into, and fate in, aquatic systems.
This type of information is fundamental to examining the effects of various discharges on the
aquatic environment, their exposure to humans and, ultimately, their impact on human health.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 63
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

The following sections evaluate copper release into, and its fate in, sediment within the
hydrosphere, taking into account geographical and temporal trends when possible. They will also
characterize coppers effects upon marine, freshwater and estuarine sediments. Major influences
on the aquatic environment, natural and anthropogenic, will be treated separately.

Marine Sediment
Marine sediment consists of both particulate organic and inorganic matter which has
accumulated on the sea floor. The sediment may originate from multiple sources, and then, after
transport by wind and rain, is deposited, precipitated from solution, or secreted by organisms.
Layers of sediment on the sea floor are affected by various physical, biological and chemical
processes occurring throughout the hydrosphere. Sediments are composed of numerous materials
and can be classified into five categories based upon their source: terrigenous, biological,
cosmic, volcanic and chemical. Since the terrigenous and biological sources are by far the most
prevalent, the studies summarized focused upon marine sediment arising from them.

Terrigenous components of sediment are produced from erosion and weathering of rocks. The
eroded material is predominantly transported to the oceans by rivers, streams and wind, and in
the process, are continuously ground into finer particle sizes. In order to classify these
sedimentary materials and establish working definitions for them, their size ranges have been
classified and organized in Table 5.

The only other major category of sediment formation originates from biological processes. For
example, the skeletons of aquatic plants and animals, comprised largely of calcium carbonate
and silica, are shed during reproduction or settle after death, becoming incorporated into the
sedimentary matrix. In certain shallow tropical marine environments, various algal species
secrete a considerable amount of calcium carbonate. The distribution of biological components
within the marine sediment matrix depends upon the aquatic species present, their skeletal
history, and the depth of the water column.

Most sediment components are deposited near their point of introduction. The sand component is
deposited close to the shoreline, with coarse and medium sands remaining close to the beach area
and finer sands being swept seaward. Silt and clay materials are among the finest sedimentary
components and are carried further out to sea; the concentration of terragenous material
decreasing in a seaward direction. Examination of an east-west sediment profile across the
Pacific basin showed sediment categories distributed symmetrically. Coarse varieties of sand are
found in the surf along beach areas, while finer sand and mud comprise the sediment further
offshore. The continental rise consists of a mixture of sediments of terrigenous and biological
origin. Sediments of a biologic origin are usually found to a depth of approximately 4.8
kilometers, after which a clay-like sediment predominates. The north-south sediment profile
contains different types of biologic sediment, but closely parallels the east-west profile. Core
profiles of various sediments reveal a layered composition. Deep ocean core samples exhibit
uniform composition throughout the sample, except for sediment close to the water-sediment
interface. Samples taken from shallower water have a mixture of sediment types.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 64
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

Estuarine Sediment
Copper in the estuarine environment was examined separately due to the special nature of an
estuarine system. An estuary is a partially enclosed body of water connected to the open ocean.
Within this water body, marine water is diluted with freshwater. Estuary systems are bordered
predominantly by mudflats or marshes containing a rich diversity of animals and plants. The
estuarine environment is not only an important feeding and breeding ground for numerous
aquatic species, but is also a prime location for residential and commercial developments. Thus,
estuaries receive pollutants dumped into rivers upstream. Sediment types may vary throughout
an estuarine environment. Clay particles are also known to clump within this environment when
mixed with seawater.

Freshwater Sediments
Concentration profiles in freshwater sediment are often used to trace the history of metal
accumulation from anthropogenic sources via long-range hydrologic and atmospheric transport.
Unlike marine sediments, lake sediments are not as reliable in determining contamination trends
because they are more subject to acidification which may demobilize metals.

Copper binds primarily to organic matter in estuarine sediment, unless the sediment is
organically poor. A study evaluated the importance of different nonlithogenic components of
aerobic estuarine sediment to copper by determining copper's adsorptivity to model sedimentary
phases from artificial seawater (Davies-Colley et al., 1984). These phases included hydrous iron
and manganese oxides, clay, aluminosilicates, and organic matter. The binding affinities varied
by over a factor of 10,000 and were in the following order: hydrous manganese oxide > organic
matter > hydrous iron oxide > aluminosilicates > clay (montmorillonite). The partition
coefficients at pH 7 for the more strongly bound phases (manganese oxide, iron oxide, and
estuarine humic material), were 6300, 1300, and 2500, respectively. The affinity increased
somewhat with pH but did not vary appreciably when salinity was reduced. Considering the
compositional characteristics of estuarine sediment, the results indicate that copper binds
predominantly to organic matter (humic material) and iron oxides. Manganese oxide contributes
only ~1% to the binding because of its low concentration in sediment; the other phases are
generally unimportant. These findings concur with results of selective extraction experiments
and the association of copper with humic material (Raspor et al., 1984a; Raspor et al., 1984b).

Experiments performed in synthetic seawater and water from Biscayne Bay, Florida, showed that
a reaction occurred, the rate of which was first-order in Cl- and second-order in H2O2. The
chloride ion is thought to be required for forming stable CuClOH-. Experiments showed that as
much as 15% of copper in seawater was present as copper(I). Additionally, sunlight increases the
percentage of free copper(II). The photochemical reduction mechanism is supported by the
observation that the copper(I) concentration is highest in the surface layer of seawater and that
the hydrogen peroxide concentration increases in parallel to that of copper(I) (Moffett & Zika,
1987b; Moffett & Zika, 1987c). In addition the percentage of free copper(II) is highest on the
surface.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 65
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

Once copper(I) is formed, its lifetime is determined by its rate of oxidation to copper(II). After
Biscayne Bay water was exposed to sunlight for five hours, the copper(I) formed was oxidized to
copper(II); the half- life of the copper(I) was twelve hours. Dissolved oxygen is primarily
responsible for this reaction. Since the oxidation of copper(I) by O2 in distilled water occurs in
<6 minutes, the copper(I) apparently is stabilized in seawater by the formation of complexes. In
the presence of humic acids, the oxidation of copper(I) occurs very rapidly. In coastal water off
the Everglades in Florida, no copper(I) was detected, due to the tying up of copper(II) in organic
complexes and the high concentration of radical oxidants in the water. Shanna and Millero
(Sharma & Millero, 1988) measured the rate of copper(I) oxidation in seawater as a function of
pH, temperature, and salinity. The rate of reaction increased with pH and temperature, and
decreased with increasing ionic strength (or salinity). The results suggested that the rates are
controlled by Mg2+, Ca2+, Cl-, and HCO3 through their involvement in complex formation and
ligand exchange.

Nriagu (Nriagu, 1979d) calculated the total quantity of copper in sediments (0.8 E20 g) using
average values of copper in shales (35 g/g), sandstones (30 g/g), and limestones (6 g/g), and
on the assumption that the relative proportions of shale, carbonate, and standstone in sediments
are 75-14-11. The copper flux from sediments to land was computed on the basis of the present
denudation rate (2.4 E16 g/year) and the average copper content of sediments (30 g/g), and was
estimated to be 72 E10 g/year.

5.4 Copper in the Lithosphere and Pedosphere


The copper burden in the earth's crust down to 45 km is estimated to be 14 E20 g. Only a very
tiny fraction (~0.001%) of the lithospheric copper burden represents potentially exploitable
deposits. The copper reserves have been estimated at just 4.6 E14 g, or 3% of the potential
copper resources. The pedosphere is generally believed to be in dynamic balance with the
biosphere, atmosphere, and hydrosphere, and its copper burden (4 E15 g) is unlikely to change in
relation to the burdens of the other three spheres.

Fossil fuels, especially shale oil, represent a major copper pool of considerable environmental
significance. The method used to estimate the copper pool for dead organic biomass is very
tenuous; it assumes that the average sediment contains 0.5% organic matter and that the average
copper content of the sedimentary organic matter is 10 g/g. The important point, though, is that
the copper pool in dead biomass is several orders of magnitude higher than the quantity of
copper in living biota.

About 80% of copper emitted into the atmosphere is deposited in terrestrial ecosystems. This
was reported as being consistent with other studies showing that roughly three quarters of metal
aerosols emitted annually is deposited in land areas as a result of proximity to the aerosol sources
(Nriagu, 1979c; Nriagu, 1979d).

Nriagu (Nriagu, 1979a) calculated the total amount of copper in soils for the upper 1 m of the
soil profile, assuming a mean copper concentration of 20 g/g (Bowen, 1966). A value for the

COPPER: Environmental Dynamics and Human Exposure Issues


Page 66
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

average copper content of soil humic and fulvic acids of 350 g/g was assumed to be
representative of all organic matter in soils. The organic matter pool in the world soils (6.8 E8 g)
was calculated using carbon content in the world soils of 3 E18 g and the C-O-N-S-P ratio of
106-106-16-2-1. Thus copper associated with soil organic matter was estimated as 2.4 E15 g, or
roughly 36% of the copper burden in soils. For riverine flux of copper from land to ocean of 6.3
E012 g/year, a residence time for copper in soils was estimated to be 1000 years.

5.4.1 Soil (Pedospheric) Dynamics of Copper and Soil-Water Interactions


Adsorption of copper to soil and sediment may really be complexation and transformation.
Between pH 5 and 6, adsorption is the principal process for removing copper from water; above
pH 6, precipitation becomes more dominant (Perwak et al., 1980). Copper binding in soil may be
correlated with pH, cation exchange capacity, soil organic content, the presence of iron oxides,
and even the presence of inorganic carbon, such as carbonates. Copper may also be incorporated
in mineral lattices where it is unlikely to have ecological significance. Broad generalizations are
not possible since the situation will differ among different soils. In soils with a high organic
carbon content, however, copper will be tightly bound to organic matter.

Most copper deposited in soil from the atmosphere, agricultural use, and solid waste and sludge
disposal will be strongly adsorbed and remain in the upper few centimeters of soil. Copper's
movement in soil is determined by a host of physical and chemical interactions between copper
and soil components. In general, the copper will adsorb to organic matter, carbonate minerals,
clay minerals, or hydrous iron and manganese oxides (Callahan et al., 1979; Tyler & McBride,
1982; Fuhrer, 1986). Sandy soils with low pH have the greatest potential for leaching. In most
temperate soil, pH, organic matter, and ionic strength of the soil solutions are the key factors
affecting adsorption (Gerritse & van Driel, 1984; Fuhrer, 1986; Elliott et al., 1986; Elliott et al.,
1992). The ionic strength and pH of the soil solution affect the surface charge of soils thereby
influencing ionic interaction. When the amount of organic matter is low, the mineral content or
Fe, Mn and Al oxides become important in determining the copper adsorption. Fuhrer (Fuhrer,
1986) reported that, in oxidized estuarine sediment, adsorption of copper is dominated by both
amorphous iron oxide and estuarine humic material. Copper binds to soil much more strongly
than other divalent cations, and the distribution of copper in soil solution is less affected by pH
than other metals are (Gerritse & van Driel, 1984).

In a study of competitive adsorption and leaching of metals in soil columns of widely different
characteristics, copper eluted much more slowly and in much lower quantities than Zn, Cd, and
Ni from two mineral soils and not at all from peat soil, which contained the greatest amount of
organic matter (Tyler & McBride, 1982). Elliot et al. (Elliott et al., 1986) looked at pH-
dependent adsorption of the bivalent transition metal cations (Cd, copper, Pb, and Zn) in two
mineral soils and two soils containing considerable organic matter. Adsorption increased with
pH, and copper and Pb were much more strongly retained than Gd and Zn. Reduction in
absorptivity after removal of the organic matter demonstrated the importance of organic matter
in binding copper.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 67
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

The form of copper in soil is determined by measuring the extractability of the copper with
different solvents. This extractability is determined by the nature of the soil and the form of
copper deposited in the soil. If a relatively labile form of copper is applied, binding to inorganic
and organic ligands may occur, as well as other transformations. On the other hand, if a mineral
form is deposited, it would be unavailable for binding. The capacity of soil to remove copper and
the nature of the bound copper were evaluated by incubating 70 ppm of copper with 5 g samples
of soil for 6 days (King, 1988). Thirteen soils (21 samples) from the southeastern United States,
10 mineral and 3 organic, were included in the study. Soil samples were taken from the subsoil
as well as from the surface. The amount of adsorbed copper ranged from 36-100%, of which 13-
100% was nonexchangeable when extracted with KCl. Removal of copper from solution was
much higher with surface soils than with subsurface sandy soils; 95-100% of the copper was
removed by five mineral surface soils and all three organic soils. The percentage of copper that
was nonexchangeable was relatively high in all but some of the acid subsoils. While the fraction
of exchangeable copper was not dependent on pH in surface soils, 96% of the variation in
exchangeability was correlated with pH in subsoils. The soil/water partition coefficient for
copper was >64 for mineral soils and >273 for organic soils. Of the eight heavy metals in the
study, only Pb and Sb had higher partition coefficients than copper. Most of the copper in
Columbia River estuary sediment and soil was correlated with inorganic carbon (e.g., carbonate),
but not with the amount of extractable Fe or the organic carbon content of the sediment (Fuhrer,
1986).

The amount of ammonium acetate- and DTPA-extractable copper, in wetland soil/sediment,


resulting from atmospheric deposition from smelters in Sudbury, Ontario, showed the same
pattern as total copper, despite random variations in soil pH, redox potential, and organic carbon
(Taylor & Crowder, 1983b; Taylor & Crowder, 1983c). Thus, in this case, soil characteristics
were not the dominant factors determining extractability and availability; the form of copper
deposited was. The median concentrations of total copper, ammonium acetate-extractable
copper, and DTPA-extractable copper at 25 sample sites were 371, 49, and 98 ppm, respectively.

Within the estuarine environment, anaerobic sediments are known to be the main reservoir of
trace metals. Under anaerobic conditions, cupric salts will reduce to cuprous salts. The
precipitation of cuprous sulfide and the formation of copper bisulfide and/or polysulfide
complexes determine copper's behavior in these sediments (Davies-Colley et al., 1985). In the
more common case where the free sulfide concentration is low due to the controlling coexistence
of iron oxide and sulfide, anaerobic sediment acts as a sink for copper. However, in the unusual
situation where the free sulfide concentration is high, soluble cuprous sulfide complexes may
form, and the copper concentration in sediment pore water may then be high.

In sediment, copper is generally associated with mineral matter or tightly bound to organic
material. As is common when a metal is associated with organic matter, copper is generally
associated with fine, as opposed to coarse, sediment. Badri and Aston (Badri & Aston, 1985)
studied the association of heavy metals in three estuarine sediments with different geochemical
phases. The phases were identified by their extractability with different chemicals and termed
easily or freely leachable and exchangeable, oxidizable-organic (bound to organic matter), acid-
reducible (Mn and Fe oxides and possibly carbonates), and resistant (lithogenic). In the three

COPPER: Environmental Dynamics and Human Exposure Issues


Page 68
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

sediments, the non-lithogenic fraction accounted for ~14-18% of the total copper, and the easily
exchangeable component was 5% of the total copper. Sediment samples taken from western
Lake Ontario were similarly analyzed by a series of sequential extractions (Poulton et al., 1988)
with regard to the compositional associations of copper. The mean (SD) percentages of copper in
the various fractions were: exchangeable, 0 (0); carbonate, 0.1 (0.3); iron or manganese oxide-
bound, 0.2 (0.3); organic-bound, 40 (11); and residual, 60 (8). Another study found that 10-20%
of the copper in Lake Ontario sediment samples was bound to humic acids, and virtually all the
copper was bound to organic matter (Nriagu & Coker, 1980). The concentration of copper
associated with humic acids was 21-40 times greater than in the sediment as a whole.

To determine the factors affecting copper solubility in soil, Hermann and Neumann-Mahlkau
(Hermann & Neumann-Mahlkau, 1985) performed a study in the industrial Ruhr district of West
Germany which has a high groundwater table (10-80 cm from the surface) and a history of heavy
metal pollution. Groundwater samples were taken from six locations and two horizons of soil, an
upper oxidizing loam and a lower reducing loam. Total copper concentrations were high in the
upper soil horizons and low in the lower horizons. Copper showed a pronounced solubility only
in the oxidizing environment; in the reducing environment, solubility was low, possibly due to
the formation of sulfides. The form of copper at polluted and unpolluted sites may affect its
leachability, particularly by acid rain. The leaching of heavy metals by simulated acid rain (pH
2.8-4.2) was measured by applying rainwater to columns containing humus layers from sites in a
Swedish spruce forest both near to and far from a brass mill (Strain et al., 1984). Leaching of
copper increased considerably when water with a pH <3.4 was applied to soil from polluted sites.
Since 25 to 75% of copper entering POTWs is removed in sludge, much of which is disposed of
by spreading on land, it is important to ascertain whether copper in sludge is apt to leach into
soil. This does not appear to be the case; leachate collected from sludge-amended soil contained
12 ppb of copper (Perwak et al., 1980). In laboratory experiments, three sludges containing 51,
66, and 951 ppm (dry weight) of copper were applied to soil columns containing four coastal
plain soils. The columns were subsequently leached with distilled water at a rate of 2.5 cm/day
for a total column application of 25.4 cm of water. Only small amounts (0.01-0.87 ppm) of
copper were found in the leachate (Ritter & Eastburn, 1978). This indicates that hazardous
amounts of copper should not leach into groundwater from sludge, even from sandy soils. In
another study, soil cores taken after sewage sludge was applied to grassland for four years
showed that 74% and 80% of copper remained in the top 5 cm of a sandy loam and calcareous
loam soil (Davis et al., 1988a; Davis et al., 1988b). Similarly, copper remains in the surface layer
when it is applied to soil as a liquid. Secondary sewage effluent spiked with 0.83 ppm of copper
was applied weekly to four different soils. After one year of treatment, the concentration of
copper in the surface horizons increased greatly; 50-76% of applied copper was found in the
upper 2.5 cm, and 91-138% was found in the upper 12.7 cm (Brown et al., 1983a; Brown et al.,
1983b). In a study of accumulation and movement of metal in sludge-amended soils, field plots
received massive amounts of sewage over a period of six years. Two sludges (one containing
industrial waste), with average copper contents of 0.29 ppm and 23 ppm were incorporated into
the top 20 cm of soil in the spring; barley was grown, and after harvest, core samples of soil
taken down to 1 m. Some movement of copper into the 22.5-25 cm layer of soil was observed,
but little, if any, below this zone. The availability of the copper in soil, as determined by its
extractability with diethylenetriamine pentaacetic acid (DTPA) and nitrate, remained constant

COPPER: Environmental Dynamics and Human Exposure Issues


Page 69
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

over a four-year period at all depths. Much of the copper discharged into waterways is in
particulate matter and settles out, precipitates out, or adsorbs to organic matter, hydrous iron and
manganese oxides, and clay in sediment or in the water column. A significant fraction of the
copper is adsorbed within the first hour, and in most cases, equilibrium is obtained within 24
hours. Copper in wastewater discharged into a river leading into Chesapeake Bay, MD,
contained 53 ppb copper, of which 36 ppb were in the form of settleable solids (Helz et al.,
1975). The concentration of copper rapidly decreased downstream of the outfall; 2-3 km from
the outfall, the copper concentration had fallen to 7 ppb. The concentration of copper in sediment
downstream from the outfall was about a factor of 10 higher than in uncontaminated sediment.

5.5 Environment-Biota Interactions: Bioconcentration and Bioaccumulation


Copper is required for normal growth in organisms. It is a component of a number of enzymes,
compounds that act as catalysts in the metabolism of organisms. Copper is also involved with
several organic molecules that transport materials within the organism and has been found in
association with certain nucleoproteins and linked to genetic activity within the cell. Copper
plays an important role in photosynthesis in plants and in the production and maintenance of
connective tissue in animals and humans. Because it is an essential metal, an adequate supply is
necessary for normal metabolism.

A sample of findings from various literature studies is presented next. A useful relevant review,
that focuses on terrestrial organisms, can be found in Parametrix Inc. & EPT, 2001).

5.5.1 Aquatic Biota

Fluxes of Copper for Aquatic Organisms


Aquatic organisms acquire the copper required for metabolic processes from soluble copper in
water and in interstitial water in sediments, adsorbed copper on particles in water or sediments,
and copper in animals they ingest. The relationship between copper in the organism and in its
environment is complex and dependent on the biochemistry of the organism, as well as the
environmental chemistry. Uptake of copper can occur from water or sediments in organisms that
are photosynthetic (phytoplankton, macroalgae and aquatic plants) or chemosynthetic (some
bacteria or bacteria-like organisms). Uptake in animals can occur either directly from water or
from ingested food. Regulation of tissue metal concentrations occurs in healthy plants and
animals with metal being utilized, stored or excreted.

In aquatic environments, metal bioavailability will be affected by geochemical processes which


affect metal speciation. Metal-metal interactions (e.g. Mn-copper, Zn-copper) can also affect
both copper uptake and metabolism.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 70
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

Uptake of Copper and Localization of Copper Within the Organism


Uptake of copper is dependent on a number of metabolic processes. It may occur over the entire
surface (e.g. an algal cell) or be confined to specialized tissues or organs. Uptake across the body
surface is often the most important mechanism in organisms with high ratios of surface area to
mass (e.g. polychaete worms). However, there is increasing evidence of variation within species
in uptake due to age, sex and size of the organism. Uptake of copper from food across the lining
of the gut is affected by the pH of the gut as well as by digestive processes, both of which
mobilize organically bound metal in the food. The pH of the digestive tract varies from organism
to organism, which implies that there can be a difference between organisms in the amount of
metal mobilized from food or from ingested sediments.

Uptake of metals by plant cells is affected by the nature of the cell wall. Within the cell wall is
the plasma membrane, a nonpolar, lipid structure containing chemical agents that assist in the
transport of essential polar substances across the membrane (Luoma, 1983). The plasma or cell
membranes of both plants and animals act in a comparable manner in the transport of metals into
and out of the cell. Some generalizations can be made about the nature of metal uptake (Lewis,
1995b):

(a) Uptake and transport can occur by carrier molecules specific for a given metal or one that
appears similar to it. From work on phytoplankton and macrophytes, there is evidence
that copper and manganese compete for uptake at the same site on the cell membrane,
possibly for the same carrier molecule.
(b) Copper associated with a lipid soluble ligand can be incorporated into the lipid structure
of the plasma membrane. Although this does not guarantee uptake, it does increase the
potential for uptake.
(c) Accidental uptake and transport of copper can result from complexation by nonspecific
carrier molecules.
(d) Metals complexed with essential organic nutrients (e.g. amino acids) can be transported
by carrier molecules specific for the nutrient.

Metal reactions with transport carrier molecules are most likely to occur at the outer surface of
the cell and to involve the most labile form of the metal (e.g. form with highest free energy).
With copper, the free ion (copper2+) is the most likely to react with transport carrier molecules.
The size of the hydrated ion radius is also small which increases the chance for absorption from
the surface of the cell membrane. Direct uptake of particles containing copper may occur in some
animals, by endocytosis. Endocytosis is the engulfment of particles by the cell membrane and
occurs with some epithelial cells. The vacuoles containing these particles are then pinched off
and become digestive vesicles within the cell.

The fate of copper after it is taken into the organism depends on a number of factors, including
the requirements of the organism, its physiological state and the conditions of exposure to the
metal. Metal uptake initiates a series of metabolic responses and, as a result, copper will be
transferred to specific ligands within the organism. These will then transport the metal to sites
where it will be utilized or sites where it can be stored or eliminated. Although a variety of
organic ligands can act as transport agents, two are commonly considered:

COPPER: Environmental Dynamics and Human Exposure Issues


Page 71
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

Phytochelatin is considered to play an important role in metal transport in a variety of


plants.
Ceruloplasmin is a transport molecule that is present in a broad range of animals,
including humans.

Copper in Aquatic Food Chains


Copper is found in organisms that serve as food for other organisms. The transfer of carbon and
energy in food chains from plants to herbivores to carnivores also has the potential to transfer
metals. It has been suggested that, as a result, there will be an increase in the concentration of
metals at successively higher levels in the food chain, a process termed biomagnification.
Uptake of copper from food does occur and is an important mechanism for obtaining metal
required for metabolic processes. However, regulation, or control, of metal concentrations occurs
within organisms at each level in the food chain. Facilitated by the action of organic molecules
like metallothionein, excess copper will be transported to a site where it can be eliminated or
stored. Studies like those of Patrick and Loutit (Patrick & Loutit, 1976) imply little or no
food-chain biomagnification of copper.

The bioavailability of copper(I) has been largely ignored since soluble or complexed forms of
copper in this oxidation state have not been thought to occur in significant amounts in aerobic
environments. Investigators have recently speculated on the possibility that copper(II) can be
directly or indirectly reduced to copper(I) by photochemical processes (Moffett & Zika, 1987a).
If this should occur, it would be more likely to occur in seawater, where chloride ions might
stabilize the copper(I) through complex formation. Copper(II)- organic complexes absorb
radiation >290 nm and can undergo charge transfer reactions where the copper(II) is reduced and
the ligand oxidized. Photochemically-generated reducing agents such as O2- and H2O2 could also
reduce copper(II) to copper(I).

Fish Bioaccumulation
The bioconcentration factor (BCF) of copper in fish obtained in field studies is 10-100,
indicating a low potential for bioconcentration. The BCF is higher in molluscs, especially
oysters, where it may reach 30,000 (Perwak et al., 1980). This may be due to the fact that they
are filter feeders, and copper concentrations are higher in particulates than in water. However,
there is abundant evidence that there is no biomagnification of copper in the food chain (Perwak
et al., 1980). A study was conducted with white suckers and bullheads, both bottom-feeding fish,
in two acidic Adirondack, NY, lakes (Heit & Klusek, 1985). These lakes were known to have
received elevated loadings of copper, but the suckers and bullhead had average copper levels of
only 0.85 and 1.2 ppm (dry weight) in their muscle tissue. The biomagnification ratio (the
concentration of copper in the fish to that in potential food) was <1, indicating no
biomagnification in the food chain. Similarly, the copper content of muscle tissue of fish from
copper-contaminated lakes near Sudbury, Ontario, did not differ significantly from that of fish in
lakes far from this source (Bradley & Morris, 1986).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 72
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

5.5.2 Terrestrial Biota


No evidence of biomagnification was obtained from a study of pollutant concentrations in the
muscle and livers of ten animal species in Donana National Park in Spain (Hernandez et al.,
1985). A study of heavy metals in cottontail rabbits on mined land treated with sewage sludge
showed that, while the concentration of copper in surface soil was 130% higher than in a control
area, the elevation was relatively small in foliar samples. No significant increase in copper was
observed in rabbit muscle, femur, kidney, or liver, indicating that copper was not
bioaccumulating in the food chain (Dressler et al., 1986).

5.5.3 Copper in the Anthroposphere


The total amount of copper in the human biomass was estimated in 1979 (Nriagu, 1979a) to be 3
E8 g, using a world population estimate for that time of 4 E9 individuals and the mean copper
content of the reference human, reported as 72 mg. The dietary copper intake for the reference
human, given as 3.5 mg/day, was extrapolated to 5.2 E9 g/year for the entire human population.
According to the numbers estimated by Nriagu, the amount of copper in the human biomass, as
well as the consumption of copper by human beings, is insignificant compared to the other
copper burdens and flux rates in the biosphere. On the basis of the dietary intake quoted above,
the lower limit for the copper residence time in human beings has been estimated to be 21 to 23
days, whereas the upper limit has been estimated at 1400 days on the basis of urinary copper
output. These human residence times must be used circumspectly, considering that (within the
body) a given element may have several residence times corresponding to different organs of the
body.

Although the human metabolic requirement for copper is small, human beings represent the most
important macrobiological agent in the present-day biogeochemical cycle of copper. Through
mining and industrial activities, they are broadcasting substantial amounts of copper into the
atmosphere, hydrosphere, and pedosphere. The ecological impacts of these anthropogenic copper
loadings remain to be fully assessed.

Total (all-time) mine production of copper has been estimated at 307 million tons. About 80% of
total world copper production has been made in the twentieth century, with the amount during
the present decade estimated as 30% of the all-time figure. Most of this enormous quantity of
copper apparently is wasted on land, where the turnover time for this element in artifacts may be
considerable. Ultimately, though, copper in artifacts will be completely wasted on land, a process
which is now being hastened tremendously by the release of acid gases to the environment. The
all-time (total) copper production figure is about twice the copper burden in the top 2-cm layer of
soils and is about an order of magnitude greater than the annual copper demand for all living
land biota. The effect of this impressive translocation of copper (from the lithosphere to the
earth's surface) on the copper economy of the pedosphere and associated fauna obviously needs
to be carefully assessed.

The total (all-time) quantity of copper that has been emitted to the atmosphere is estimated at 3.2
million tons. This figure is about three orders of magnitude greater than the present-day
atmospheric copper burden. Because of the relatively short residence time for airborne copper

COPPER: Environmental Dynamics and Human Exposure Issues


Page 73
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

aerosols, it is doubtful that there can be an inexorable buildup of copper in the atmosphere. The
relevant point, however, is that the atmosphere is the important medium for the transmission of
pollutant copper to the most remote areas of the globe. Indeed, analyses of moss samples and
polar ice cores reveal a recent substantial increase in ombrogenic deposition, and hence in
airborne levels, of copper at locations far from any sources of copper emission.

If a constant land-ocean deposition ratio of 80-20 is assumed, the total quantity of airborne
copper that has fallen into the oceans is estimated at 44 E 10 g. This is small compared to either
the dissolved copper burden in the ocean or the annual riverine influx of copper. It is, however,
about two orders of magnitude greater than the annual copper demand by aquatic biota. In fact,
atmospheric loading of copper into most terrestrial and aquatic ecosystems now exceeds the
combined copper requirement of fauna and flora (Nriagu, 1979a). Obviously, atmospheric supply
of copper to any given ecosystem needs to be viewed with some concern until the relative
availabilities of native and ombrogenic copper to biota can be determined.

Copper pollution of local ecosystems has been documented in numerous reports. For example,
ambient copper levels in urban air typically range from 3 to 200 ng/m3 and are orders of
magnitude higher than values observed at remote locations (Nriagu, 1979a). In New York City
copper fallout from the atmosphere often exceeds 100 mg/m2 year. Copper contents of street
dusts and urban park soils frequently exceed 1000 g/g, while copper concentration in household
dusts has been found to be 700 to 900 g/g. Human beings have turned their cities into hot
spots of environmental copper levels. Significant copper level elevations are also commonly
encountered in ecosystems around smelters and other copper point sources.

In view of the analytical uncertainties in the data reported, it is difficult to evaluate the effects of
human wastes on copper concentrations in natural waters. However, copper contents of
sediments from rivers, lakes, and coastal waters that receive domestic or industrial waste
discharges have been shown to exceed natural background levels by factors of 2 to 100. For
example, copper contents of surficial Lake Erie sediment range from 100 to 250 g/g, while
1970s levels in Hudson River and New York harbor sediments generally exceeded 300 g/g
(Williams et al., 1978). Some of the highest copper concentrations recorded in sediments
anywhere are the 2 10 to 12,000 g/g reported in Sorfjord, West Norway, which receives
massive doses of industrial effluents. The former practice of using copper salts to control algal
growth in lakes and ponds invariably results in massive accumulations of copper in sediments,
with unknown consequences for benthic organisms.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 74
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

Table 5: Summary of Copper Concentrations in Environmental Media as Reported


in the Copper Sourcebook 1998 (Harrison, 1998)
Environmental Media Concentration Units
Atmosphere
Aerosol 0.1 382 ppt

Hydrosphere Water
Coastal Dissolved 0.06 4.3 ppb
Total 0.5 13.8 ppb
Suspended Solids 0.6 37000 ppm
Estuarine Dissolved 0.02 4.7 ppb
Total 1.2 71.56 ppb
Suspended Solids 0.38 72 ppm
Ocean Dissolved n.d. 10 ppb
Total 0.04 10 ppb
Suspended Solids 0.01 2.8 ppm
Lake Dissolved 0.1 15.6 ppb
Total 0.1 15.6 ppb
River Dissolved 0.18 3000 ppb
Total 0.5 5800 ppb
Groundwater Dissolved 0.003 70 ppb
Total 1 1160 ppb
Drinking Water Total 0.3 1352 ppb

Hydrosphere Sediments
Coastal Particulate 0.03 3789 ppm
Interstitial Water 25.5 32.7 ppb
Estuarine Particulate 0.3 2895 ppm
Interstitial Water 0.3 100 ppb
Ocean Particulate 3.1 648 ppb
Interstitial Water 22 45 ppm
Lake Particulate 0.4 796 ppm
Interstitial Water 45.6 52 ppb
River Particulate 5.3 4570 ppm

Pedosphere
Soil Total 0.01 3138 ppm
Organic 293 7634 ppm
Dust Total 2.9 176 ppm

COPPER: Environmental Dynamics and Human Exposure Issues


Page 75
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

Table 6: Copper Concentrations in Environmental Media and Biota as Reported in


The Handbook of Trace Elements (Pais & Benton Jones, 1997)
Environmental Media Concentration
Soils
Total Content in Soils 2-100 ppm; geometric mean, 18 ppm
Soluble Content in Soils <1 ppm (3-135 g/l in soil solutions)
Hydrosphere
Sea Water 8E-5 ppm
Fresh Water 0.01-2.8 ppm; reference level, 3 ppb
Biota
Plants 1-10 ppm
Animals 2.4 ppm
Common Foods (Human Diet) 3-80 ppm; raw product, 0.26-12 ppm wet weight
Cows Milk 0.3 ppm
Human Tissues
Muscle 10 ppm
Bone 1-26 ppm
Blood 1.01 mg/dm3
Kidney 1.07-4.19 ppm
Urine 6.1-30.3 ppb
Scalp Hair 26.0 ppm

COPPER: Environmental Dynamics and Human Exposure Issues


Page 76
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

Table 7: Copper Concentrations in Environmental Media and Biota as Reported in


Nriagu (Nriagu, 1979b)
Total Mass of Average Copper
Environmental Media
Medium (kg) Concentration (ppm)
Lithosphere (down to 45km) 5.7 E22 24
Sedimentary Rocks 2.5 E21 30
Shale and Clay 1.9 E21 35
Limestone 3.5 E20 6
Sandstone 2.8 E20 30
Soils (to 100 cm) 3.3 E17 20
Organic Fraction 6.8 E15 350
Fossil Fuel Deposits
Coal 1.0 E16 17
Shale Oil 4.6 E16 70
Crude Oil 2.3 E14 0.14
Copper Ore
Reserves 4.6 E11
Resources 1.5E13
Terrestrial Biomass
Plant Biomass 2.4 E15 12
Animal Biomass 2.0 E13 12
Litter 2.2 E15 60
Oceans
Dissolved, seawater 1.4 E21 0.15 ppb
Dissolved, pore water 3.2 E20 15 ppb
Particulate organic matter 7.0 E13 7.0
Dissolved organic matter 3.5 E18
Aquatic Biomass
Plants (live) 2.0 E11 3.5
Animals (live) 3.0E12 1.0
Swamps and Marshes, Biomass 6.0 E12 12
Lakes and Streams 3.4 E16 2.0 ppb
Glaciers 1.7 E19 0.2 ppb
Atmosphere 5.1E18 m3 0.5 ng/ m3

COPPER: Environmental Dynamics and Human Exposure Issues


Page 77
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

Table 8: Sediment Component Classes


Particle Classification Diameter (mm)
Very coarse sand 1-2
Coarse sand 0.5-1
Medium sand 0.25-0.5
Fine sand 0.125-0.25
Very fine sand 0.0625-0.125
Silt 0.0039-0.0625
Clay <0.0039
Mud mixture of silt and clay and various debris

COPPER: Environmental Dynamics and Human Exposure Issues


Page 78
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

resuspension
deposition
deposition, inspiration
vegetation exudates, forest fires, resuspension
permeation, entrainment
ANTHROPOSPHERE
MICROENVIRONMENTS
(residential and occupational)

HUMANS ECONOSPHERE
(individuals and populations)
FOOD (noncommercial) MINES, SMELTERS

BIOSPHERE (NON HUMAN)


FOOD (commercial) PROCESS FACILITIES emissions
TERRESTRIAL BIOTA

MICROORGANISMS ingestion WELLS & SPRINGS FOSSIL FUELS

ATMOSPHERE
PLANTS MUNICIPAL WATER Cu COMPOUNDS

AEROSOLS
ANIMALS disposal

HYDROSPHERE
HYDROMETEORS
SURFACE WATER
AQUATIC BIOTA
FRESH SURFACE WATER MARINE WATER
MICROORGANISMS
ESTUARIES GLACIERS

PLANTS
RIVERS COASTAL REGIONS

ANIMALS
LAKES OPEN OCEANS

GEOSPHERE

PEDOSPHERE (Soil/Dust) SEDIMENTS GROUNDWATER

ORGANIC FRACTION FRESHWATER SEDIMENTS MARINE SEDIMENTS VADOSE ZONE

ESTUARY SEDIMENT
INORGANIC FRACTION SATURATED ZONE

RIVER SEDIMENTS COASTAL SEDIMENTS


LITHOSPHERE
LAKE SEDIMENTS OCEAN SEDIMENTS
ORES

COAL

SHALE & CRUDE OIL

volcanic emissions

Figure 9: Conceptual Multilevel-Multiscale Model of Environmental and Human


Exposure Dynamics of Copper

COPPER: Environmental Dynamics and Human Exposure Issues


Page 79
Chapter 5
ENVIRONMENTAL AND BIOLOGICAL CHEMODYNAMICS OF COPPER

A Simple Conceptual Model of


fossil fuel combustion GLOBAL DISTRIBUTION AND FLUXES OF
COPPER
ATMOSPHERE (draft operational prototype 1999.12.07)
nonFe metal production

HUMANS inhalation
?
? fw ingestion
Fe metal production ?

mwb ingestion
? tb ingestion
other anthg emissions
?

resuspension ? dermal and pica


deposition on land MARINE BIOMASS
fwb ingestion
forest fires ?

mwb assim from mw


vegetation exudates
seasalt sprays

ANTHROPOSPHERE TERRESTRIAL BIOMASS FRESHWATER BIOMASS


deposition on oceans

fwb assim from fw


PEDOSPHERE FRESHWATER MARINE WATER

volcanic emissions
mb decay ?

tb assim soil river runoff


land waste disposal
?

tb decay sediment outflux to fw


fertilizer application sediment outflux to mw
LITHOSPHERE SEDIMENTS
?
mining
fwb accum in sed

fw sedimentation
waste disposal mw sedimentation

Figure 10: Prototype Implementation of a Model of Global Distribution and Fluxes


of Copper in the Environment, Using the STELLA (Structured Experimental
Learning Laboratory with Animation) Simulation Software

COPPER: Environmental Dynamics and Human Exposure Issues


Page 80
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

6 COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES: OVERVIEW


OF SELECTED FIELD STUDIES

6.1 Atmospheric Concentrations


Copper concentrations in air depend on site proximity to major sources, such as smelters, power
plants, and incinerators.

According to the EPA's National Air Surveillance Network report for the years 1977, 1978, and
1979, median copper concentrations were 133, 138, and 96 ng/m3, respectively, for urban
samples and 120, 179, and 76 ng/m3 for nonurban samples (Evans et al., 1984). In this study,
10,769 urban and 1402 nonurban air samples collected for 24 hours were analyzed. For 1977,
1978, and 1979, 1% of urban samples exceeded 1156, 975, and 843 ng/m3, and 1% of nonurban
samples exceeded 1065, 1396, and 645 ng/m3, respectively. The maximum urban and nonurban
copper concentrations reported were 4625 and 4003 ng/m3, respectively. Davies and Bennett
(Davies & Bennett, 1985) reported average atmospheric copper concentrations of 5-50 ng/m3 in
rural areas and 20-200 ng/m3 in urban locations. Several studies have shown that concentrations
in rural areas are considerably lower than those reported in the EPA survey. Data from many
urban locations in the United States from the mid-1970s to the late 80s show concentrations of
copper associated with particulate matter ranging from 3-5140 ng/m3 (Schroeder et al., 1987).
Remote and rural areas had concentrations of 0.029-12 and 3-280 ng/m3, respectively. Levels
reported by Schroeder et al. (Schroeder et al., 1987) are consistent with those obtained in a study
of airborne trace elements in national parks (Davidson et al., 1985). In Smokey Mountain
National Park, copper concentration in air was 1.6 ng/m3, while in Olympic National Park, where
several locations were monitored, 3.3-6.7 ng/m3 of copper was measured in the atmosphere. The
lower copper concentrations found in Smokey Mountain Park compared with those in Olympic
National Park have been attributed to greater vegetative cover and higher moisture in the former
and larger amounts of exposed rock and soil in the latter. Average copper crustal enrichment
factors, or copper concentration in air compared with the average concentration in the earth's
crust, were 31 and 76, respectively.

As part of the Airborne Toxic Element and Organic Substances (ATEOS) project for
determining patterns of toxic elements in different settings, three urban areas (Camden,
Elizabeth, and Newark) and one rural site (Ringwood) in New Jersey were studied during two
summers and winters between 1981 and 1983 (Evans et al., 1984; Lioy et al., 1987). Each site
was sampled every 24 hours for 39 consecutive days. Geometric mean copper concentrations
were 16.0-21.0, 21.0-36.0, 21.0-33.0, and 6.0-63.0 ng/m3 for Camden, Elizabeth, Newark, and
Ringwood, respectively. Copper levels measured in these industrial urban areas are considerably
lower than mean values reported in the National Air Surveillance survey [201-259 mg/m3 for
1977-1979 (Evans et al., 1984)]. A possible reason is that the samplers used in the ATEOS study
were among the few that prevented sampling of the emissions from the motor brushes. Summer
and winter maxima in the three urban areas ranged from 100.0-131.0 and 231.0-493.0 ng/m3,
respectively, and 77.0 and 29.0 ng/m3, respectively, for Ringwood. Copper follows the same

COPPER: Environmental Dynamics and Human Exposure Issues


Page 81
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

pattern as other heavy metals, in that increased copper levels are present in winter in urban areas
and in summer in rural areas. No explanation for this pattern has been offered.

Anderson et al. (Anderson et al., 1988) performed a study of the atmospheric aerosols collected
at a site in Chandler, AZ, over a 12-day period in February and March 1982. Several major
copper smelters are located ~120 km to the southeast. Particles containing >0.5% copper were
termed copper-bearing particles; 5.6% of the fine (0.4 to ~2 mm) particles collected were in
this category. The most abundant type of copper-bearing particle, representing 74% of the total,
was associated with sulfur; however, the analysis was not able to specify the form of sulfur
present. These particles were often associated with Zn, Fe, Pb, As, and Ca. Sixteen percent of the
copper-bearing particles were associated with silicon and 4% were associated with chloride. The
concentration of copper-S particles was highest when the surface and upper-level winds were
from the southeast to the east, and reached a maximum 1-2 days after the winds began to blow
from the southeast; smelters to the southeast were the probable source. Particles associated with
silicon and chlorine did not show any apparent correlation with wind and were from either a
diffuse regional source or a local source.

Aerosols
Aerosols are stable atmospheric suspensions of fine particulate matter (solid, liquid or
multiphase). Copper concentration in aerosols was measured during the rainy and dry seasons at
two rural savannah locations in western Venezuela (Morales et al., 1996). Copper concentration
in the rainy season was up to eight times greater than that of similar rural locations in eastern
Venezuela, and up to 300 times that in other parts of the world. Rainy season concentrations
were also up to 2.5 times dry season concentrations at the same location.

Copper concentration in aerosols was measured in both a moderately polluted area and an area
with background pollution levels in Hungary (Hlavay et al., 1996). Mean concentrations were
7.7 and 4.5 ng/m3, of which approximately 11% and 17% were environmentally mobile, 56%
and 57% were bound to silicates and organic matter, and the balance 33% and 26% were bound
to carbonate and metal oxides, respectively. Aerosol copper was enriched 60- and 39-fold
relative to aluminum at the moderately polluted and background sites, respectively. In
comparison, the enrichment factor at Budapest was 158; aerosols with enrichment factors > 10
are generally considered to have contributions from anthropogenic sources.

Copper concentration in suspended dust and soil, as well as in human hair and teeth, were
measured at a polluted small mountain town in Poland, along with wet and dry deposition rates
of copper (Nowak, 1993). The average concentration in suspended dust was 3.91 g/m3, and that
in the soil was 18.6 g/g, of which less than 10% was in an environmentally mobile fraction. In
comparison, the concentration of copper in soil from an unpolluted area was 8.8 g/g.

Copper concentrations in atmospheric aerosols were measured at a site in Majorca, Spain for a
period of 4 months (Mateu et al., 1996). They ranged from 7 - 19 ng/m3, and since local

COPPER: Environmental Dynamics and Human Exposure Issues


Page 82
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

anthropogenic copper sources are negligible, copper concentration variation is due to long-range
transport. The series of aerosols collected over the duration of the study were grouped by cluster
analysis of the other metals in the aerosol. Back-tracing of the trajectories of these aerosols
indicated that the origin of aerosols in a particular group tended to be similar. The mass median
diameter of copper was 0.43 um, small enough for a large fraction of copper in the aerosol to
reach the lower airways of human lungs.

Natural sources for trace elements in the atmosphere are crustal weathering, volcanic action and
the sea.

The source of copper-containing atmospheric particulate material collected at the South Pole is
derived from a source other than crustal weathering. This is an area isolated from major sources
of anthropogenic emissions in both Northern and Southern Hemispheres.

Transmission and retransmission of trace metals into the atmosphere on a global scale were
surveyed in a simplified global model framework (Nriagu, 1979a). Eroded soil particles
contribute heavily to natural copper levels, although how much of this level was due to
retransmission from anthropogenic sources could not be determined. It was estimated that
anthropogenic copper emissions into the atmosphere exceed natural levels by more than 300%.
Most anthropogenic copper emissions originate from the Northern Hemisphere.

Street and house dusts were analyzed for trace metal levels in various studies (e.g. Lioy et al.,
1987). Soil is the major contributor to the trace metals found in dust. Soil is the main component
of street dust. Other than soil, street dust sources of these metals include road traffic, industry,
weathered materials, and specific intermittent episodes. Automobile wear and tear contributes
to atmospheric copper levels as does weathered materials. House dust includes outdoor soil and
its trace element components. Copper concentrations are generally found to be higher in house
dust than in the outside soil. It cannot be determined whether the copper source was generated
within the house or whether the outside source was something other than soil. The dustiness of
the house is expected to increase the metal loading, and this also increases with the amount of
carpet wear. Also, copper levels inside the house are typically found to increase with the
proximity to roads with heavy traffic. Indeed, copper concentration in dust is generally elevated
not only in the vicinity of high industrial activity, but also, though to a lesser extent, in the
vicinity of heavy traffic. In addition, type of industrial activity has considerable impact on copper
concentrations in dust. Temporal and spatial variability in copper concentrations in street dust
were studied in Bahrain (Madany et al., 1996). Copper concentrations in areas of high traffic
density were significantly greater than concentrations in the vicinity of an oil refinery, which in
turn were greater than concentrations in urban commercial areas. Copper concentrations in the
vicinity of an aluminum smelter, and in remote areas were approximately six-times smaller than
concentrations in high traffic density areas. The association of elevated copper concentrations
with high traffic density in this study contrasts with a study in Alice, South Africa (Fatoki,
1996), where copper concentration in soil adjacent to a highway was low, and did not vary with
distance from the highway.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 83
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

In another study, copper concentration in indoor dust was found to be generally higher than that
in outdoor dust, and the ratio appears to be larger in areas with a history of industrial activity,
including copper smelting activity. This suggests that copper may accumulate in indoor dust over
a period of time. Concentrations of copper in indoor and outdoor dust was measured in Riyadh,
Saudi Arabia (Al-Rajhi et al., 1996), a region that experiences heavy dustfall due to its dry
climate and major construction activities in many areas of the city. The average copper
concentration in outdoor dust (0.04-0.08 mm) in an area with old industries, and which had
copper smelters, was approximately 300 g/g, two-fold greater than levels observed in newer
industrial areas, urban areas, suburban areas, rural areas, or near motorways.

There was a fairly strong correlation (r = 0.7) between average copper concentrations in indoor
and outdoor dust, but copper concentration as a function of particle size was strikingly different
between the two. In outdoor dust, copper concentrations decreased with increased particle size,
whereas the opposite was true of indoor dust. It is notable that copper concentration in indoor
dust was generally 2- to 3-fold higher than outdoor dust concentrations. The only exceptions to
this were in suburban and new industrial areas, where the concentrations in indoor dust were 6-
fold higher and in outdoor dust about the same.

6.1.1 Precipitation
Mean concentrations of copper in remote and rural precipitation ranged from 0.013-1.83 and
0.68-1.5 ppb, respectively, on a volume-weighted basis (Barrie et al., 1987). Although an earlier
survey referred to by these investigators yielded much higher values, 0.060 and 5.4 ppb, these
were ascribed to sample contamination. The mean concentration of copper in rain reported in an
extensive study in southern Ontario, Canada, was 1.57 (0.36 standard deviation) ppb during 1982
(Chan et al., 1986). These concentrations showed little spatial variability and agree with those
reported by Barrie et al. (Barrie et al., 1987).

6.1.2 Fog
Elevated levels of copper in fog water were observed 3 km downwind from a refuse incinerator
in Switzerland (Johnson et al., 1987). High copper concentrations were associated with low pH.
The maximum concentration, 673 ppb, occurred at pH 1.94; levels >127 ppb were associated
with pH values <3.6. Copper(II) concentrations in fog water from the central valley of California
were 1.7-388 ppb (Miller et al., 1987). The source of the copper was not investigated. High
values were recorded just as the fog was dissipating.

6.2 Copper Concentrations in the Hydrosphere


Copper levels in surface water range from 0.5-1000 g/l, with a median of 10 g/l; seawater
contains from <1-5 g/l (Davies & Bennett, 1985; Page, 1981; Yeats, 1988; Mart et al., 1984).
The results of several studies in which copper was detected in drinking water, surface water, and
groundwater are presented in Table 12: Concentrations of Copper in Water.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 84
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

The geometric mean (standard deviation) and median concentration of dissolved copper in
surface water, based on 53,862 occurrences in EPA's STORET database, are 4.2 (2.71) and 4.0
ppb, respectively (Eckel & Jacob, 1988). Higher concentrations tend to be found in New
England, the western Gulf, and the lower Colorado River (Perwak et al., 1980). An analysis of
high concentrations of copper in minor river basins reported in EPA's STORET database in 1978
revealed that sources of copper in the Gila, Coeur D'Alene, and Sacramento River Basins appear
to be primarily mining activities, especially abandoned sites (Perwak et al., 1980). The high
concentrations were generally localized at specific stations. The low pH of the surface water in
these areas was reported to exacerbate the situation. However, in another study concerning lakes
sensitive to acid rain, copper levels were relatively low (1-8 ppb range, 2 ppb mean) regardless
of pH or alkalinity (Reed & Henningson, 1984).

In a study of representative groundwaters and surface waters throughout New Jersey, in which
>1000 wells and 600 surface sites were sampled, the median copper levels in groundwater and
surface water were 5.0 and 3.0 g/l, respectively (Page, 1981). The respective 90th percentile
and maximum levels were 64.0 and 2783.0 g/l for groundwater and 9.0 and 261.0 g/l for
surface water. The pattern of contamination in surface water correlates with light hydrocarbons,
while that in groundwater correlates with heavy metals. This indicates that the sources of
contamination in the case of surface water and groundwater are probably different. The nature of
the sites with elevated levels of copper was not indicated. Experimental data demonstrate that
leaching of copper is minimal.

The copper concentration in some bodies of water evidently varies with season. In one small
pond in Massachusetts, the concentration varied from <10-105 g/l (Kimball, 1973); copper
levels were low from summer to late fall and rose to maximum levels in midwinter. This cycling
in copper concentrations is thought to be a response to biological need and uptake of copper
during the growing season and its subsequent release from biota decay.

Copper concentrations in seawater usually range from 1-5 g/l (Perwak et al., 1980). Copper
levels are lower in the Pacific Ocean than in the Atlantic Ocean and higher near the continental
shelf than in the open ocean. Copper concentrations in surface water transected on a cruise from
Nova Scotia to the Sargasso sea ranged from 57.2-210 ng/l (Yeats, 1988). The mean value in
surface water of the eastern Arctic Ocean was 93 ng/l (Mart et al., 1984).

6.2.1 Marine Water


The marine environment is comprised of all the interconnected oceans and seas. The copper in
this environment undergoes biogeochemical cycling between the biota, water, and sediments,
and the dynamics of these processes have important implications for the viability of marine life.

Copper in marine water exists in dissolved and suspended particulate phases. The concentration
of free Cu2+ in marine water is strongly dependent upon the availability of organic ligands as
well as chlorides and sulfides. Most of the dissolved copper is tightly bound to ubiquitous
organic ligands, which serve to buffer the concentration of free Cu2+. The metal binding

COPPER: Environmental Dynamics and Human Exposure Issues


Page 85
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

properties of coastal and estuarine waters were measured at four-month intervals off the coast of
central southern England (Muller, 1996; Al-Rajhi et al., 1996). The copper ligands appear to
interact mainly with the truly dissolved copper, and not with the colloidal or particulate fractions.
Significant seasonal variation in free copper concentrations were observed, concentrations in
July being larger than those in November and March.

Copper concentrations in the sea-surface microlayer, an approximately 50 m-thick layer of the


water column, is enriched by a factor of approximately twenty over the concentration in the bulk
surface mixed layer of the ocean, which extends to about 10 m in the summer in the North
Atlantic (Shine & Wallace, 1996). This enrichment occurs due to a variety of transport
processes, including atmospheric deposition, diffusion, and bubble scavenging. The sea-surface
microlayer serves as a habitat for a variety of flora and fauna, and factors affecting metals and
organic matter transport to the microlayer can have important implications for viability of the
biota in this system. An important factor mediating copper accumulation in the sea-surface
microlayer is the concentration of surface-active organic matter, which forms stable complexes
with copper, and acts as a copper reservoir at the ocean surface. This surface-active organic
matter is largely oceanic in origin (phytoplankton are a major source), although anthropogenic
sources may contribute to a significant extent in coastal areas. The flux of surface-active organic
matter from the mixed surface layer of the ocean to the sea-surface interface occurs by molecular
diffusion, as well as by scavenging of the bulk of the surface water by bubbles created by the
action of wind and waves. The molecular diffusive flux of copper bound to surface-active
organic matter is estimated to be three orders of magnitude greater than the flux due to
atmospheric deposition, bubble scavenging, or diffusion of particulate copper from the surface
mixed layer.

Riverine inputs are a major source of copper in marine water. Copper concentrations were
measured in the coastal waters of the Adriatic Sea, south of the River Po estuary (Tankere &
Statham, 1996). Copper concentrations were slightly elevated nearer the surface, and were
negatively correlated with salinity. This indicates that atmospheric and riverine inputs are major
sources of copper in this environment. The highest concentrations observed were near the mouth
of the Po River (30 nM).

The flux of copper from rivers to oceans can vary considerably with season. The total
concentration of copper and other heavy metals were measured in the coastal waters of the South
China Sea along Guangdong province (Ke & Jiang, 1995). The concentration of copper
decreased as the salinity and pH increased, and chemical oxygen demand (COD; a surrogate for
organic ligands) decreased. The average concentrations in the wet season was almost twice as
high as that in the dry season.

The concentration of copper in marine water in some regions not directly affected by riverine
inputs have actually decreased. The concentration of dissolved copper in the Gulf of Trieste was
measured at six sites over a period of seventeen years (Reisenhofer et al., 1996). For example,
the average copper concentration in 1992-93 was approximately 2 ppb.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 86
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

The flux of copper from riverine and atmospheric sources has been modeled using an advection-
diffusion model proposed earlier to describe the concentration of copper in the southern North
Sea which was extended to incorporate the particulate phase (McManus & Prandle, 1996). The
model was solved using inverse-problem techniques, and good agreement was observed between
the model predictions and observed dissolved and particulate concentrations at 100 sampling
locations in the North Sea. The log KD values used in the model to relate dissolved and
particulate concentrations ranged from 4.9-5.3.

6.2.2 Estuarine Water


Estuaries serve to trap much of the copper that flows out of rivers into oceans. A first-order mass
balance was estimated to describe the geochemical transport of copper and other trace elements
in the Firth of Clyde, adjacent to the Clyde estuary in the UK. Approximately 115 tons/year of
copper is transported out to the open sea, representing 25% of the input into the Clyde estuary.

Dissolved and non-detrital particulate fluxes of copper and other trace metals into and out of the
plume of the Humber River, in the UK, were estimated (Millward et al., 1996). Samples of SPM
were collected from estuarine, coastal plume, and coastal background waters during the winter,
spring, and summer seasons. The total flux of copper out of the estuary is estimated to be 90
kg/day in the summer and 10 kg/day in the winter, whereas the total flux into the estuary is
estimated to be 270280 kg/day, implying that a substantial amount of copper is deposited as
sediments in the estuary.

6.2.3 River Water


Rivers comprise a major component of the copper flux into oceans, the source of which is both
natural soil erosion and anthropogenic. Trace element concentrations were measured in ten major
rivers draining into the Humber estuary, in the UK (Neal et al., 1996), some of which originate in
the relatively rural northern catchments, while others flow through the southern industrial and
urban catchments. Concentrations of dissolved copper range from approximately 4-8 g/l, with
the higher concentrations being found in the rivers flowing through industrial and urban areas.
Although these concentrations are significantly higher than background levels of 1 g/l, they are
much lower than the industrial rivers of Europe, exemplified by the Rhine, in which
concentrations in excess of 30 g/l have been measured. Copper concentrations of less than 6
g/l have been recommended for the protection of salmonid fish. Copper was determined to be
released from both point and diffuse sources, based on an analysis of the variation in copper
concentrations with river flow rate.

The concentration of copper was measured in the dissolved and particulate phases along the St.
Lawrence River in Canada, to estimate the rate at which the metal is introduced into the river by
its tributaries and industries along the river (Quemerais et al., 1996). The average dissolved
concentration of copper in the St. Lawrence River was 0.7-0.8 g/l, whereas the concentrations
in its tributaries varied between 0.3 and 1.1 g/l. The dissolved concentrations in the St.
Lawrence River were not significantly influenced by the dissolved concentrations in its

COPPER: Environmental Dynamics and Human Exposure Issues


Page 87
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

tributaries due to the dilution of these concentrations by the relatively greater water flow in the
St. Lawrence River. The concentration of dissolved copper in the St. Lawrence River was similar
to the concentrations in other major, relatively pristine rivers, such as the Lena and Amazon
Rivers, and were two-fold less than the concentration in the Mississippi River. The concentration
of copper in suspended particulate matter was also relatively constant throughout the St.
Lawrence River, at 60-70 g/g, approximately two- to three-fold less than concentrations in the
Amazon (266 g/g) and in the Rhine (325 g/g), but greater than in the Mississippi (42 g/g).
Copper in the suspended particulate is associated mainly with particulate organic matter.
Transport of copper in the river is higher in the spring due to snowmelt, which also causes an
increase in the concentration of particulate matter. Approximately 50% of the copper in the St.
Lawrence was in the dissolved phase, compared to only approximately 10% in pristine rivers like
the Amazon and Yukon, and is of concern due to free Cu2+ toxicity to aquatic biota.

A review of copper concentrations in eight Chinese rivers (Zhang, 1995) indicates that the
copper concentration in suspended particulates in these rivers is in the range 26 - 89 g/g, and
that the enrichment factor (EF; ratio of copper:Al in sediments and fresh rock/soil) of sediments
was 1 or less for all the rivers for which data was available. Concentration trends in sediments
tend to be less variable than in suspended particulates, and therefore EF for sediments provides a
more long-term assessment of pollution than would EF for suspended particulates. EF ratios of
less than 1 are unexpected, and indicate that the fresh rock/soil samples chosen as the
normalization constant in the EF may not be representative of the rock/soil the sediments
originated from. Nonetheless, the low levels are still indicative of low copper pollution. The
partitioning of copper between the dissolved and particulate phases appears to be relatively
constant, as represented by log KD values ranging from 4.2 to 4.7. The total flux of copper in
riverine suspended sediments to the ocean in China is approximately 70 million kg/year, 60
million of which is accounted for by the Huanghe and Changjiang rivers. The flux of dissolved
copper to the ocean is approximately 1.7 million kg/year, of which approximately 90% is
accounted for by the Changjiang river.

The utility of aquatic moss in providing a temporally integrated, and spatially detailed measure
of copper concentration in a river was demonstrated for a stream near Stockholm, Sweden. The
concentration of copper in the moss increased from approximately 5-10 mg/kg in the relatively
pristine upper reaches of the stream to almost 40 mg/kg near the lower, urbanized section of the
stream. A definite gradient in copper concentration was observed even in the upper reaches of
the stream, and is thought to be due to copper released by corrosion in the water pipes in the
houses along the stream.

Copper concentration in the leaves and roots of common cattail growing in the wetlands adjacent
to the heavily contaminated Clark Fork River in Montana were measured (Johns, 1993).
Concentrations in the cattail roots were approximately 25 times background level downstream of
the copper mining and smelting areas, whereas concentrations nearest these areas were close to
background, possibly due to liming of the settling ponds adjacent to the sampling point.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 88
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

6.2.4 Lake Water


The concentration of dissolved copper in lakes is coupled both to biological processes, and to the
oxidation potential of lake waters. Spatial and temporal distribution of copper in the dissolved
and sediment phases were studied in two eutrophic Swiss lakes (Xue et al., 1997); Lake Greifen
becomes seasonally anoxic in its hypolimnion, whereas Lake Sempach remains oxic throughout
the year. The concentration of major inorganic ions and dissolved oxygen was similar in both
lakes. However, the concentration of dissolved copper in Lake Sempach was approximately 5-7
nmol/l, with slightly higher concentrations observed during the summer, in both the epilimion
and the hypolimnion, while the concentration in Lake Greifen was two- to three-fold higher, with
the highest concentrations occurring near the surface, in the period between winter overturn and
the summer stagnation. Furthermore, the flux of copper due to settling sediments did not vary
with depth in Lake Greifen, whereas in Lake Sempach the flux was much higher in the
hypolimnion. This suggests that the oxidative status of a lake effects the cycling of copper via
two competing processes. Under oxic conditions, the rate at which copper is released from
organic material in the sediments increases, but under these conditions manganese and iron form
oxides and settle out of solution, and in the process scavenge copper which binds to the surface
of the oxide particles.

The vertical profile and spatial and temporal variation of copper in Lakes Superior, Erie, and
Ontario were studied using ultraclean laboratory methods (Nriagu et al., 1996). Dissolved copper
concentrations in the Great Lakes ranged from approximately 700 -1100 ng/l, with the higher
concentrations occurring near the shoreline. Average concentrations remained at approximately
830 ng/l during spring and summer, possibly due to buffering by particle scavenging. Dissolved
copper concentrations in the Great Lakes were similar to those observed in Canadian arctic lakes
(100-1200 ng/l). Although the average conductivity of lake water (a measure of total dissolved
solids) increased three-fold between Lake Superior and Lake Ontario (down the drainage basin)
dissolved copper concentration was relatively uniform, the average concentrations in Lakes
Superior, Erie and Ontario being approximately 760, 870, and 830 ng/l, respectively. Although
atmospheric input of copper into the Great Lakes is substantial, 330-1470 ng/m2-yr, accounting
for 60-80% of anthropogenic input into Lake Superior and 20-70% into Lakes Erie and Ontario,
observed concentrations did not match estimated loading rates, suggesting that the behavior of
dissolved copper is not conservative, and involves cycling in the ecosystem. The principal
outflow from the Great Lakes was into the St. Lawrence River, where dissolved copper
concentrations of 150-890 ng/l have been observed.

The mean residence time of copper in these lakes ranged from 15 (Lake Erie) to 101 years (Lake
Superior). Shorter residence times imply that dissolved copper is rapidly scavenged by
suspended particles, hydrolyzed to solid or sedimentable forms, or biologically assimilated.

Spatial patchiness in copper distribution in Lake Ontario was observed to be correlated with
biological productivity, and since conductivity and pH were relatively constant, is unlikely due
to chemical factors. The effects of recent reductions in copper emissions on dissolved
concentrations is masked by the long residence times, and the measured concentrations reflect
long-term trends.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 89
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

Seasonal variation in copper concentrations in the water, sediments, and fish of Nashua
Hydrodome, a shallow lake fed by the Nile River in Egypt, was studied (Saad, 1987). Average
seasonal concentrations of dissolved copper ranged from 1-80 g/l, and the average
concentration in sediments was 130 g/g. Copper concentration in the flesh of Tilapia fish
ranged from 2-66 g/g, and was closely related to dissolved copper concentration. High seasonal
variability was attributed mainly to anthropogenic activity.

6.2.5 Groundwater
Groundwater concentrations are affected by copper leaching from soil, and by concentrations in
runoff water that eventually percolates through the soil. The concentration of copper in
groundwater measured in Tirupati, India (pop. 350,000) was found to range between 10-20 mg/l,
with higher concentrations occurring in the summer, probably due to decreased dilution from
rainwater infiltration (Mohan et al., 1996).

In the absence of dramatic change in soil properties and pollution, groundwater concentrations
can be expected to be similar to adjacent bodies of surface water. Copper concentrations were
measured in the Nile River, as well as in the groundwater and drinking water (post-filtration river
water) near the Aswan Dam (Ismail, 1996); copper concentrations were similar in all three
media, between 5-6 g/l.

Copper concentrations in leachate from an uncontrolled garbage dump in Buenos Aires were
determined before, immediately after, and four months after reclamation work at the dump site in
1992. Samples were taken from three zones within the dump site: Zone 1 consisted of recent
domiciliary garbage; Zone 2 was contained garbage at the lowest elevation of the dump, which
was in contact with the underlying aquifer for most of the year, and was intermediate in age; and
Zone 3 contained the oldest waste, dating from 1974. Concentrations of dissolved copper ranged
from 42 to 140 g/l, and adsorbed concentrations, from 33 to 135 mg/kg. Dissolved and
adsorbed copper concentration immediately following reclamation were higher than initial
concentrations, but decreased after a four-month period, with adsorbed concentrations dropping
below initial levels. The reclamation works resulted in adsorbed copper mobilization.

Copper concentrations in storm water runoff was determined at several general industrial sites,
such as auto salvage, metal fabrication, and vehicle maintenance sites, in North Carolina (Line et
al., 1996). There was a wide range of variability in the first flush runoff water concentrations
collected. The highest average concentrations, approximately 2200 ppb, were observed in a scrap
and recycling site, and the lowest, approximately 40 ppb, were observed at the vehicle
maintenance site. Runoff from all sites exceeded the North Carolina standard of 25 ppb. Some of
the sites observed, however, compared favorably with the mean concentrations of 144 ppb
recorded in the Nationwide Runoff Program.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 90
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

6.2.6 Sediments
Sediment is an important sink and reservoir for copper. In pristine areas, sediment generally
contains <50 g/g copper; the level can reach several thousand g/g in polluted areas. The mean
copper level in surficial sediment of Penobscot Bay, Maine, was 14.1 g/g (dry weight), while
that in estuaries or bays in other New England locations ranged from 4.4- 57.7 g/g (Larsen et
al., 1983). Levels reflect anthropogenic input as well as the regional bedrock mineral content.
Copper levels in sediment from 24 sites along the New Jersey coast ranged from <1.0-202 g/g,
with a mean value of 66 g/g (Renwick & Edenborn, 1983). The sediment texture varied from
94% clay to 100% sand, and copper level correlated negatively with percentage of sand.

Surficial sediment in lakes in the Sudbury region of northeastern Ontario, where several smelters
operate, decreased rapidly with increasing distance from the smelters (Bradley & Morris, 1986).
Three lakes 10 km from the Sudbury smelters contained copper concentrations in sediment
approaching 2000 ppm dry weight, over 100 times the concentration in a baseline lake 180 km
away.

Marine and Estuarine Sediments


Marine sediments have been examined for a variety of minerals and elements. Most of the
studies reviewed examine copper deposition from anthropogenic sources rather than from natural
processes.

The coastal environment has become a major focal point in the study of industrial and domestic
waste effluent discharges. Of particular interest is their contamination of marine sediment.
Research performed in Greece (Angelidis, 1995) has shown that urban effluents contain high
organic content, and high metal loading attributable to particulate matter. Metals discharges have
been shown to settle into sediment, becoming non-point sources of pollution. Sediment sampling
was conducted along a coastal area in a harbor near the city of Mytilene. Finer sediments were
found in the inner harbor area, while coarse sediments were found beyond the harbor. This is the
reverse of the sediment characteristic profile discussed earlier. Copper concentrations in the
sediment were high in the inner harbor area, and highest near sewage outfalls (30.7-58.9 g/g).
Lower concentrations of copper in sediment were found along coastal sampling stations (9.4-
17.5 g/g). High copper enrichment (i.e., 2.0 times the background levels) was also found in the
inner harbor area, as opposed to the outer harbor area, or along the coastline. Total copper in
sewage effluents was measured at 0.0460.016 g/g, and the mean copper concentration in
particles was 137.837.9 g/g of sediment. Copper was shown to exceed background in urban
effluent unlike other metals found in sewage.

Heavy metals contamination was examined at an untreated sewage system discharge point for
the two largest cities in Greece, Athens and Thessaloniki (Savvides et al., 1995). The aim of the
study was to estimate the level of contamination and determine sediment metals distribution via a
sequential extraction procedure to partition the metals into five chemical phases: exchangeable,
metals bound to carbonates, reducible metals, organic bound metals and residual metals.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 91
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

Subsequently, experiments were performed to examine the ability of sediment to release heavy
metals. Copper concentrations in the soils near Athens ranged between 0.16 and 0.26 mg/g (avg.
0.2150.036), while concentrations in sediment near Thessaloniki ranged between 0.05 and 0.26
mg/g (avg. 0.1570.083). In order to assess the amount of contamination from anthropogenic
sources, contamination factors were calculated for each of the studied metals. The contamination
level, expressed as a contamination factor (C), was calculated using the relationship: C =
(sediment metal content/natural reference sediment metal content). Contamination factors for
Athens followed the decreasing order Cr>Zn>copper>Fe>Pb>Ni and the factors for Thessaloniki
followed the order Cr>Zn>Fe>copper>Ni>Pb. Since all of these values were much higher than
unity, the result demonstrated that effluent discharges have significantly impacted sediment
quality near both cities. Sediment partitioning experiments indicated that copper was
predominant near both cities in the organic phase (Athens, 51%; Thessaloniki, 61%), followed
by the residual phase (i.e., non-exchangeable, non-carbonate, non-organic, non-reducible).
Copper extraction from sediment was determined to be independent of acid concentration, with
most copper extraction from sediment occurring within the first 10-15 minutes after contact with
HCl.

The aforementioned sediment studies were short-term in duration, lasting either days or months,
and did not allow for temporal analysis of copper contamination. A year-long study was
conducted through the Latvian Academy of Sciences in the Gulf of Riga to determine the impact
of a waste treatment facility on the level of heavy metals in the areas ecosystem (Seisuma et al.,
1993). Water, sediment, plankton and mollusks were sampled near the sewage discharge area.
Copper concentrations ranged between 0.3 and 8.8 mg/kg dry weight in the estuarine discharge
area during sampling. The highest concentration of copper in water occurred at the central
discharge point, but sediment concentrations were highest in the areas north and west of the
central discharge location. Thus, it appears that waste waters washed contaminated sediment
away from the central discharge location and transported metals, including copper, in a northerly
and westerly direction. Ocean sampling in the area revealed copper concentrations ranging
between 0.6 and 4.4 g/l. Transport of metals did not appear to affect concentrations in the
mollusk population studied, while uptake by plankton was directly related to metal
concentrations in water but not sediment. An additional analysis of activated sludge from the
treatment plant determined that healthy activated sludge actively accumulates heavy metals.
Copper concentrations within the sludge ranged between 2052-5980 mg/kg dry weight. When
the sludge is deactivated it not only fails to absorb more metals, but allows desorption to occur.

To examine the impact of sewage discharge on sediment contamination along the North
American continent, temporal trends of metal concentrations in sediment were studied near a
wastewater outfall in California (Phillips & Hershelman, 1996) between 1985 and 1992. This
outfall discharged approximately 260 million gallons per day (MGD) of treated municipal and
industrial wastewater. While average effluent flow volume increased by 12% during this
sampling period, the copper emission rate was determined to have been 84 kg/day in the 1985-
1986 sampling year, but steadily declined to 31 kg/day during the 1991-1992 sampling year.
Sampling of the sediment near the outfall found copper concentrations ranging from 20.5 to 51.6
mg/kg during the 1985-1986 sampling period and then decreasing to a range of 14.4 to 21.4

COPPER: Environmental Dynamics and Human Exposure Issues


Page 92
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

mg/kg of sediment. When effluent copper concentration was normalized by suspended solids
concentrations, it also showed a steady decline, with 1,860 mg copper/kg SS discharged between
1985-1986 and only 800 mg copper/kg SS discharged between 1991 and 1992. Sediment and
effluent concentrations were found to be significantly correlated, indicating the impact of
wastewater contamination on the sedimentary matrix.

Another study which examined heavy metals accumulation in sediment was conducted to
determine the impact of untreated sewage, industrial waste and landfill leachate discharge into
Halifax Harbor, Nova Scotia (Buckley et al., 1995). Since the early twentieth century, copper
concentrations have increased from 26 g/g in the harbor sediment to 88 g/g. A factor analysis
of various geochemical data determined that sediment contamination from waste discharges was
associated primarily with both total, and organic-bound forms of, lead, zinc and copper.
Contamination attributed to leachate from landfills and from modification of contaminants by
chemical reactions in the sediment have been attributed to acid-labile forms of zinc, nickel and
copper. Characterization of the sediment found it to exist mostly of fine-grained particles.

A study of heavy metals distribution in the Eastern Mediterranean (Ergin et al., 1996) suggested
that aluminosilicates and organic matter are effective carriers of metals, facilitating transport to
sediments. Copper concentration in surface sediment in the Gulf of Iskenderun ranged from 9
and 33 g/g. However, in order to compensate for dilution caused by carbonates, total
concentrations for various metals were recalculated on a carbonate-free basis. Upon
recalculation, copper concentrations ranged between 14 and 52 g/g, similar to those found in
average crustal rocks (crustal rock concentrations not given). This study also noted that copper
concentrations increased significantly with decreasing grain size in sediments composed of clay
or mud. This is due to the larger surface area of the small particles and their strong adsorptive
properties. The study also noted a strong correlation (r=0.84) between copper and magnesium
concentrations. A significant correlation was also found between sediment copper and zinc
concentration (r=0.87). This appeared to be due to uptake of these elements as micronutrients by
microorganisms. The major source of copper in this Eastern Mediterranean region was primarily
weathering of crustal rocks.

A study supporting the hypothesis that transport caused copper concentrations to increase with
distance offshore was conducted in Europe. Research performed in the Netherlands (Nolting et
al., 1996) showed that copper concentrations increased from the Lena Delta seaward. Sediment
copper concentration found in open waters off the coast varied between 2.1 and 20 g/g, while in
that same region, estuarine concentrations ranged between 26 and 219 g/g, increasing seaward.
The metals measured in the study (Ni, Cd, Pb, Zn) were found in highest concentration in river
sediments, while offshore concentrations were not significantly higher than background (not
given). Depth profiles of sediment cores did not reveal much variation in the concentration of
any trace metal, including copper.

While the aforementioned transport studies were of rather short duration, weeks to several
months, a longitudinal study was conducted over a seven-month period off the coast of Sydney
Australia. Its purpose was to examine the distribution of contaminants in sediment found in three

COPPER: Environmental Dynamics and Human Exposure Issues


Page 93
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

spatially discrete zones from the shoreline (Schneider & Davey, 1995). Sampling was conducted
from shore 2 km seaward, between 2 and 4 km from shore and, finally, beyond 4 km from shore.
The turbulent waters in this region are storm-dominated with wave heights recorded up to 12
meters. Bottom currents were measured up to 1 m/s in the region, whereas surface currents could
exceed a speed of 1.5 m/s. The lowest concentration of copper was measured in sediment 2 km
from shore in waters up to 50 meters in depth. Similarly low levels were also measured in the
sampling region 4 km offshore in waters that exceeded 80 m in depth. Highest concentrations of
copper were measured in the 2-4 km region offshore in sediment at depths between 50 and 80
meters (actual concentrations not provided, only contour modeling profiles). A stepwise
regression model was developed to isolate the anthropogenic contribution to sediment
contamination from natural geochemical processes. The high copper levels in the sediment 2-4
km offshore were determined to originate from the harbor area, due to the fast-moving currents
in the region. Harbor copper contribution to total offshore sediment copper was calculated at
16%; the remaining copper contamination was attributed to deep sea sewage outfalls from
urbanized coastal areas.

The previously discussed studies attempted to address the spatial distribution of copper
contamination from both sewage and industrial discharges. While extremely important, this
information does not adequately address the issue of copper accumulation throughout the depth
of the sediment matrix. Thus, information is needed on the sedimentary profile of copper
accumulation. For example, temporal and spatial variations in trace metals within bottom
sediments were examined in the Sea of Japan (Tkalin et al., 1996). Sediment core analysis was
performed in 2 cm increments from surface to 28 cm in two locations around the Sea of Japan.
Analysis of the Amursky Bay sediment core found fairly uniform copper concentrations ranging
between 5.9 and 8.3 ppm throughout the core sample. The average copper core concentration
was 6.90.7 ppm. A Ussuriysky core sample (taken in Peter the Great Bay area of the Sea of
Japan), while having a slightly higher copper concentration averaging 10.31.7 ppm, was found
to also maintain a relatively uniform concentration throughout the 30 cm depth of the core
sample. Fluxes for copper in the Amursky Bay and Ussuriysky Bay were calculated at 43 mg m-2
year-1 and 18 mg m-2 year-1, respectively. Sedimentation rates for trace metals were estimated for
the Amursky Bay and Ussuriysky Bay as 0.17 g cm-2 year-1 and 0.12 g cm-2 year-1 respectively,
indicating not much variation between the two bays.

In contrast to other core studies, a Croatian study of trace metal accumulation in sediment
(Mihelcic et al., 1996) found the highest concentrations of copper within the first 10 centimeters
of sediment within a saltwater lake (11.3-13.2 g/gm dry weight). Copper concentrations at
depths between 15 and 40 centimeters of sediment were 5-6 times lower than surface
concentrations.

The transport and fate of mine tailings in sediments were studied in a coastal fjord in British
Columbia (Odhiambo et al., 1996). Core samples collected from Alice Arm and Upper
Observatory Inlet were analyzed for metals, including copper, to establish sediment
accumulation rates. Accumulation rates of copper in Alice Arm ranged between 1.4 and 2.0
g/cm2 per year, whereas accumulation rates in the inlet varied between 0.17 and 0.76 g/cm2 per

COPPER: Environmental Dynamics and Human Exposure Issues


Page 94
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

year. Thus, sediment, including metals from local mining operations, become trapped in the Arm
after entering the waterway. Sediment copper concentrations near the mines, one located at the
head of Alice Arm and another located 32 km from the head of Alice Arm on the banks of the
Kitsault River, ranged between 56.3 and 70 g/g; copper concentrations averaged 650 g/g in
the vicinity of a smelter located about 25 km from the head of Alice Arm. Crustal copper
concentration in the mining area was determined to be 55 g/g. Core samples taken outside of
Alice Arm waterway in the Upper Observatory inlet in British Columbia, showed enrichment of
copper, Cd, Pb and Zn in the sediment. Concentration of these metals decreased with distance
from the mine. The best tracer for mine tailing dispersion was copper since, copper background
levels being very low, even low levels in the sediment could be detected.

A study to evaluate the effect of organic matter decomposition on diagenetic remobilization of


copper was undertaken in the Kalix River estuary in northern Sweden (Widerlund, 1996).
Diagenetic changes occur when certain substances in sediment undergo chemical reaction and
are transformed into rock after being buried under subsequent sediment deposits. This estuarine
basin was 10-15 meters deep and contained sediments which are deposited at a rate of
approximately 1.0 cm yr-1. Early diagenetic copper demobilization was found to be controlled
entirely by decomposition of a highly reactive organic matter fraction in the surface layer of the
sediment. Core profiles taken from the estuary system at a depth up to 36 cm revealed copper
concentrations between 5.7 and 42 nmol/gm. Sediment traps in the river measured copper
concentrations between 456 and 740 nmol/gm. Flux calculations indicated that approximately 3
percent of all copper deposited into the sediment layer is released back into the water column
throughout the year.

A European study conducted in the Mediterranean (Scoullos et al., 1996) reported copper
concentrations in sediment ranging between 18 and 52 g/g. Contrary to the results of other
studies mentioned earlier, metal contamination within core samples was seen at a surface sample;
however little information was available on sediments and core sampling since the primary focus
was on water sampling and water levels.

Along the Asian continent, heavy metal fluxes in Thailand were estimated to approximate the
movement of copper between the water column and sediments (Cheevaporn et al., 1995).
Sedimentation rates for the Bang Pakong River estuary showed sediments to be accumulating at
a rate of 0.19-0.28 gm cm-2 yr-1. Sedimentary fluxes (0.1-16.8 g cm-2 yr-1) of copper and other
metals studied were higher than the diffusive fluxes of these metals (0.01- 4.8 g cm-2 yr-1). The
mean concentration of copper in porewaters ranged from 22-100 ppb, and did not appear to
decrease with sample depth. Metal concentrations in the porewaters, however, did appear to be
higher than in waters overlying sediments. In this system, between 70% and 91% percent of
diagenetic contribution to metal flux was attributed to copper. Thus, for copper, diagenesis
contributes significantly to metal enrichment of surface sediments.

An evaluation of surficial sediments from Lake Michigan and the Virginian province (coastal
Virginia) was performed to assess proposed sediment quality criteria for copper and other metals
(Leonard et al., 1996). The concentration of acid-volatile sulfide, simultaneously extracted

COPPER: Environmental Dynamics and Human Exposure Issues


Page 95
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

metals, total metals and total organic carbon were measured in core samples taken from these
areas. Ninety-one percent of collected sediments contained measurable quantities of acid volatile
sulfide. More than seventy percent of the marine sediments had concentrations of acid-volatile
sulfide exceeding the concentration of the simultaneously extracted total metals, indicating little
bioavailable metal in the sediment. Freshwater samples contained more simultaneously extracted
total metals than acid-volatile sulfide, indicating more bioavailable metal in the freshwater
sediment region.

Heavy metal removal (transport) was studied in the Changjiang estuary in China (Shen & Liu,
1995). The annual transport of copper from the Changjiang River to the East China Sea is
approximately 8000 tons. Measurements of particulate and ion forms of copper indicated that
concentrations of the ion form of heavy metals, including copper, were quite low (0.48 g/L for
copper) compared to the particle form (11.7 g/L for copper). River water entering the East
China Sea changes both the salinity and pH of the water and leads to coagulation of the ionic
form and particulate form of the heavy metals. Very large particles result from coagulation and
enter the sediment. Most metals (e.g., zinc, cadmium) are removed in the salinity range of 10 to
23. The maximum movement of this copper complex from water into the sediment was 22%.
With changing salinity, the horizontal distribution of total copper indicated a 51% decrease from
estuary to sea.

Fresh Water Sediments


Fresh water systems and copper sediment contamination is the last category to be examined in
this chapter. Concentration profiles in this aquatic environment are often used to trace the history
of metal accumulation by the sediment from anthropogenic sources via long-range hydrologic
and atmospheric transport. As opposed to the marine environment, lake sediments are not as
reliable in determining contamination trends since they are more subject to acidification which
may cause demobilization of metals. One of the most extensively studied systems is Lake Erie,
one of the lower Great Lakes of North America. Trace element profiles of the sediment to a
depth of approximately 50 centimeters were recently obtained to determine relationships
between sediment metal concentrations and the concentration in pore water of the Central Basin
of Lake Erie (Azcue et al., 1996). The study determined that the lowest copper concentration
(approximately 30 g/g) in the sediment existed between 30 and 50 centimeters below the water-
sediment interface. The concentration of copper steadily increased from a depth of 30
centimeters through a depth of 18 centimeters below the sediment surface, coinciding with heavy
anthropogenic inputs during the 1960s (approximately 75 g/g). Concentrations of copper then
decreased steadily through the column from 18 cm to the sediment surface (approximately 60
g/g). Sediment concentrations of dissolved trace elements below the water-sediment interface
are approximately one order of magnitude greater than concentrations measured in the lake
water. Copper and other element concentrations measured within the first several centimeters of
sediment were determined not to be representative of anthropogenic deposition since this upper
layer is subject to a variety of biogenic processes and the effects of hydrologic transport.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 96
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

Heavy metals (Zn, As, Pb, copper and Cd) present in sediments were also studied in two
tributaries of the Chesapeake Bay (Gupta & Karuppiah, 1996). This was done to determine the
contributions of both point source and non-point source pollution to the estuary. Metal
concentrations were found to be higher in water near a sewage treatment facility and a poultry
farm than in sediment just downstream from a sewage treatment facility alone. Sediments in
deeper water also had lower copper concentration than sediment in shallower water, presumably
resulting from hydrologic transport and dilution.

Metal accumulation of cadmium, chromium, copper, iron, nickel, lead and zinc in bottom
sediments of wet ponds was investigated in south and central Florida (Yousef et al., 1996).
Accumulation rates for certain heavy metals, specifically copper, correlated with ratios of pond
surface area to drainage basin area. The mean copper accumulation rate for the 13 sampling sites
was 1.3 kg/ha yr1.3 kg/ha yr. The content of copper was determined as 23.5 g metal/g dry
sediment in the accumulated top sediments. Copper concentration declined exponentially with
depth. An equation was developed for predicting copper accumulation rates in kg/ha yr as
follows: copper=0.013*(DA/PA)1.08 + 0.35, where DA=total drainage area (ha) and PA=pond
surface area (ha). Copper accumulation rates appear to level off when the PA:DA ratio 2%.

An evaluation of copper, lead and zinc contained in inter-city canal sediments was conducted in
the Netherlands (Bijlsma et al., 1996). The study was designed to characterize the sediment,
which typically had a thickness of 20-30 cm, resting upon the sandy bottom of the canals.
Copper concentration in the sediment top layer ranged between 45 and 281 mg/kg (mean 166
mg/kg) and ranged between 150 and 530 mg/kg (mean 350 mg/kg) in the sediment bottom layer.
Mass balance calculations determined a copper accumulation of 164 kg/yr.

While the previous fresh water studies were of relatively short duration, days to weeks, several
studies of longer duration were conducted in the fresh water environment to establish temporal
sediment trend of metals. The release of trace metals (iron, manganese, lead, copper cadmium
zinc and chromium) from sediments was studied during a two-year period along a 110 km
segment of the Seine River in France surrounding a sewage treatment facility (Garban et al.,
1996). Copper concentrations in the sediment increased noticeably several kilometers
downstream from a sewage treatment facility (8% in 1991 and 87% in 1992) and steadily
decreased from approximately 24 kilometers downstream from the sewage treatment outlet
through the end of the sampling area. Zinc concentrations, while consistently higher than copper,
showed an overlapping concentration profile, suggesting similar aquatic behavior. Copper
concentrations in the sediment were similar from year one (215 mg/kg) to year 2 (216 mg/kg). In
addition, an increase in copper sediment concentration was accompanied by an increase in PAH
and PCB concentrations.

Seasonal variation in river metal contamination was studied in sediments near the discharge from
an abandoned copper mine (Herr & Gray, 1996). Over the course of a one-year period, copper
drainage from the mine averaged 11 kg/day. Significant positive correlations were found
between copper and iron, suggesting co-precipitation of the two metals. Average copper
sediment concentrations were lower in the spring (414.5229 g/gm) than in the summer

COPPER: Environmental Dynamics and Human Exposure Issues


Page 97
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

(587379 g/gm), but gradually continued to increase over time. Mean surface copper decreased
approximately 15% during the study period.

Transport of copper, lead, zinc and cadmium was observed during flood events in an Alpine river
in Germany (Hamm et al., 1996) and it was determined that flooding has a major impact on total
copper transport. During both normal flood and high flood events during the sampling period,
approximately one quarter of annual copper transport occurred. Copper concentrations in flood
waters reached a maximum of 400 g/l at the highest stage of flooding.

Pollution status of the Indus River in Pakistan was examined through the study of fish, sediment
and water heavy metal concentration (Ag, As, Cd, Cr, copper, Fe, Mn, Ni, Pb, Zn, Hg) (Tariq et
al., 1996). This is one of the worlds largest rivers in terms of drainage basin, discharge and
sediment load. Concentration analysis of copper and other heavy metals indicated an increase in
concentration at locations downstream. Sediment concentrations for copper ranged between 2.6
and 33.2 g/g. Fish provided a uniform distribution of metals along the sampling route, whereas
this association could not be established for water or sediment. There was an inverse relationship
between trace metal and nutrient content in the fish. There was a positive correlation between
trace metal concentration in fish and sediment. Maximum macronutrient concentration in
sediment was found where the trace metal concentrations in fish were not at maximum levels. It
was also determined that trace metal concentrations followed a decreasing order as follows:
sediment>fish>water.

A statistical analysis was undertaken in the analysis of river sediments to identify areas of lead,
copper, iron and chromium enrichment caused by anthropogenic activities by examination of
concentration ratios of pairs of metals (Murray, 1996). Sediment contamination was identified by
comparing both trace metal-to-conservative metal concentrations and absolute metal
concentrations in the sediment along the Rouge River in Michigan. Comparison of the ratio of an
element, such as copper, to another element of lower sediment variability, such as iron, can
determine if metal enrichment has occurred. Since these low variability metals are always
detected at higher initial concentrations, they are less affected by anthropogenic activities, while
trace metals are more susceptible to changes from anthropogenic activity. The average copper
concentration along the river as measured from 16 sampling locations was 15316 ppm. The use
of a copper-to-iron ratio reduced the effect of sediment grain size and organic content on metal
concentrations and was seen as a conservative estimate of sediment contamination. Ratios can
also be used to separate metal contamination due to natural and anthropogenic sources.

A statistical approach was also developed to delineate sediment zones according to trace metal
distributions (Poulton et al., 1996). Analysis included ratio matching, cluster analysis and
principal component analysis. The use of ratio matching was employed to reduce the effect of
sediment dilution by other, inert, materials. Sediment samples of similar origin were defined by
ratio matching and cluster analysis. Average total copper in the sediments of Hamilton Harbor on
Lake Ontario was 8936.8 g/gm. Extractable copper for the same sediment was measured to be
24.211 g/gm. Zones were delineated as follows: Group 1 - center of harbor, deep water
undergoing good mixing, fine sediment; Group 2 - located close to shore and had higher sand

COPPER: Environmental Dynamics and Human Exposure Issues


Page 98
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

content and noticeably lower trace metal concentration, high wave action; Group 3 - similar to
Group 2, but high cadmium concentrations, region experiencing longshore drifts; Group 4 -
located in a channel having abnormally high copper and other metal concentrations; Group 5 -
very close to shore with an extremely high sand content (99.5%) and very low metal
contamination, extremely shallow water, large amount of mineral separation, high wave activity.
Copper and trace metal contamination was directly related to the presence of clay and silt and
inversely related to the presence of sand.

It is not possible to provide a clear picture of copper accumulation in sediments found in the
marine, estuarine or freshwater environments based upon the studies examined for the
monograph from the Copper Source Book and other sources of information. Perhaps the most
noticeable aspect of each sediment study reviewed in the document was that copper was seldom
the major focus of a study. Its concentration in the sediment was usually reported without
adequate discussion. Studies of sediment contamination primarily focus on elements, such as
lead, arsenic, mercury and zinc, or examine the essential nutrient composition of sediment in
terms of the requirements of a host of aquatic organisms.

In these studies, a variety of methods were used to sample for heavy metal or trace metal
contamination in the aquatic environment without any noticeable attempt at standardization of
both sampling and analysis. For example, several of the papers focusing on the marine
environment attempted to examine the impact of industrial and municipal waste outfalls on
heavy metal contamination in water and sediment. However, no standardized method or
consistent set of criteria for sediment sampling was used in the vicinity of outfalls. An attempt
should be made to define zonal areas at discrete distances from a wastewater outfall in both
seaward and down-current directions to test copper-specific hypotheses on flux, transport and
deposition in sediment. This would enable researchers to conduct sampling at defined distances
from outfalls and to better compare sediment contamination and contaminated sediment transport
among numerous sampling locations near outfalls. Environmental parameters such as water
current direction and flow rate, salinity and degree of turbidity should also be consistently
reported in the literature when conducting these type of studies. Many of the studies do not
mention the type of sediment under investigation, nor do they provide adequate detail on
sediment particle size fractions.

Core sampling in the aquatic environment needs a set of criteria, like those mentioned above, to
allow for inter-comparison of results from different studies. A review of the literature focusing
on temporal trends of heavy metal (e.g., lead) accumulation in sediments indicates an increase in
contamination beginning in the late 19th century. Then, after the Second World War, a rapid
increase in heavy metal contamination (e.g lead, arsenic, cadmium, mercury) was reported,
reaching a plateau between the 1960s and 1970s when stricter regulations on waste discharges
were implemented. These observations, however, must be carefully interpreted since there does
not appear to be a standardized core sampling methodology to compare data obtained by
different studies. Standardization is necessary to ensure that reported elevations or reductions in
sediment metal concentrations (e.g., copper) are not artifacts of a poorly-developed sampling
strategy. Several of the aforementioned studies only examined sediment in the first 2-5 cm from

COPPER: Environmental Dynamics and Human Exposure Issues


Page 99
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

the surface and this data may be erroneously compared to studies using deeper core samples.
Since turbulent water redistributes contaminated sediment over vast distances, it would be more
appropriate to take deeper samples to determine the concentration of more permanent
(sequestered) copper in the region, as opposed copper found in the superficial upper layer. Even
when deeper core sampling was performed, there was no established standard depth for taking
core samples. Some core sampling is performed between 2 and 10 centimeters below the surface,
whereas other cores are taken to a depth of 50 cm. Appropriate comparison of core profiles can
only be conducted if the profiles are of similar depth. Only when this standardization is
implemented can sediment layering characteristics and depth of copper contamination be fully
understood.

Finally, while many papers state that levels of copper in the sediment do not significantly exceed
background levels for that area or those levels determined in crustal samples, background values
are frequently left unreported for that region. Sufficient research has not been conducted to
determine aquatic background levels of copper in either pristine environments, or areas accepting
urban and industrial effluents.

6.3 Copper Concentrations in Soils and Terrestrial Biota


Copper occurs naturally at levels of ~50 g/g in the earth's crust, which includes soil and parent
rock (Perwak et al., 1980). In agriculturally productive soils, copper ranges from 1-50 g/g,
while in soil derived from mineralized material, copper levels may be much higher (Perwak et
al., 1980). Copper concentrations in soil samples collected throughout the United States yielded a
geometric mean of 18 g/g (Fuhrer, 1986). Samples were taken at a depth of 8 inches to avoid
anthropogenic contamination; 2/3 of the samples contained copper concentrations between 8.0
and 40 g/g. These copper levels are supported by a review of soil copper concentrations that
reported a median concentration of 30 g/g (dry weight) and a range of 2-250 g/g (Davies &
Bennett, 1985). Copper concentrations in soil may be much higher in the vicinity of a source.
Concentrations in the top 5 cm of soil near the boundary of a secondary copper smelter were
2480 585 g/g (Davies & Bennett, 1985). Maximum wetland soil/sediment copper
concentrations were 6912 g/g in the immediate vicinity of a Sudbury, Ontario smelter, but the
concentrations decreased logarithmically with increasing distance from the smelter (Taylor &
Crowder, 1983a). Results suggest that copper in the soil from the study area was primarily
associated with particulate emissions from the smelter.

In a study in which the copper concentrations of 340 soil samples were presented in terms of
land-use types, the average copper concentrations reported were 25 g/g in agricultural land, 50
g/g in suburban/residential land, 100 g/g in mixed industrial/residential areas, and 175 g/g in
industrial/inner urban areas (Haines, 1984). From an analysis of the spatial distribution of the
copper, it was concluded that most of the contamination was a result of airborne deposition from
industrial sources. Soils from Lemhi, Twin Falls, and the Idaho National Engineering Laboratory
in southern Idaho had geometric mean copper concentrations of 13.4-20.4 g/g dry weight (Rope
et al., 1988).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 100
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

Elevated concentrations of copper in soils and dust appears to be localized to areas of industrial
activity, particularly in areas of copper smelting. The effect of land-use on copper concentrations
in top soil was investigated in Richmond-upon-Thames (low industrial activity), and
Wolverhampton (high industrial activity), in the U.K. (Kelly et al., 1996). The concentrations of
copper in Wolverhampton were approximately 2-fold higher than the concentrations in soil from
areas with similar land uses. However, there were order of magnitude differences in residential
soil concentrations within each of the two areas, probably due to the practice of disposing of
fireplace ash in gardens.

The deposition rate of copper in the vicinity of a copper refinery in Prescott, UK before and after
installation of emissions controls was reported in a study of copper concentrations and
ecotoxicity (Turner et al., 1993). Copper deposition appears largely localized to areas in the
immediate vicinity of the plant, where the deposition rate (averaged from the years 1970-1990)
was more than 150-fold greater than urban background in the UK. At a distance of 3 km from the
site, the deposition rate dropped to only two-fold UK urban background. The high rate of copper
deposition resulted in elevated concentrations of copper in soil, and although there was
considerable spatial variability in copper concentrations (281 - 2106 ppm in 1991), largely due to
variation in tree canopies, copper concentrations were generally elevated down to a depth of 1 m.
Copper concentrations in soil near a copper refinery, in which emissions controls were installed
in 1979, were found to decrease over the period 1988 to 1991 from an average concentration of
2363 ppm to 1522 ppm (Turner et al., 1993). The high concentration of copper in soils near the
site resulted in growth reduction in sycamore trees, as evidenced by tree ring widths, compared
to a reference site 3 km away. Subsequent reduction in soil copper concentrations were
accompanied by reduction in extent of growth inhibition at the site.

Several other studies demonstrate the localized effect of industrial activity on copper
concentrations in soil. Copper concentrations in the top soil were studied from the Kastela Bay
region of Croatia (Milos et al., 1993). This is a densely populated 120 sq km coastal region,
containing industrial and urban areas. Copper concentrations ranged from <7.5 to 150 mg/kg.
High copper concentrations are associated with urban and industrial areas. Concentration of
copper and other heavy metals were measured in soil at eight sites in the vicinity of various
industrial complexes in Kiev, Russia (Shotyk et al., 1993). The average concentrations at each
site ranged from 10-36 mg/kg. Concentrations were elevated by 20 mg/kg near a metallurgy
complex and by 15 mg/kg near an oil refinery, and by 10 mg/kg in the middle of the city.

The spatial variability of copper concentrations in the top soil at Gelbe, an industrialized town in
Australia, was studied by extensive random stratified sampling (Markus & McBratney, 1996).
Variation in soil concentration was considerable (2 to 4141 mg/kg), only approximately 15% of
which could be explained on the basis of soil disturbance and distance from roads.

Industrial activity does not appear to have had an effect on the background concentration of
copper in soils. Temporal changes in the concentration of copper between 1935 and 1988 in
agricultural surface soils (0-15 cm) in several different areas of Illinois were studied (Granato et
al., 1994). Copper concentrations ranged from 5-14 mg/kg, and were found to be constant over

COPPER: Environmental Dynamics and Human Exposure Issues


Page 101
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

time. Therefore, although this region of the United States as a whole experienced substantial
growth in industrial activity over the five-decade study period, there was no discernable impact
on soil copper concentration.

The lack of effect of industrial activity on background concentrations of copper in soil is


supported by a study of the extent of spatial variability of copper in soil in the former East
Germany, another heavily industrialized region. Copper concentrations in samples taken from
approximately 9% of East German agricultural top soil were measured (Machelett et al., 1993).
The average and median copper concentrations were 11 and 8.8 mg/kg respectively, 56% of soils
had concentrations less than 10 mg/kg, an additional 33% had concentrations less than 20 mg/kg,
and less than 1% of soils had concentrations greater than 50 mg/kg. Only 0.2% of soil had
concentrations greater than 100 mg/kg, the current limit value for copper.

Furthermore, copper concentration in agricultural soils does not appear to be affected by


phosphorous fertilizer application. Copper is essential for plant growth, and fertilizers generally
contain a small amount of it; there is evidence that soils having less than 1 mg/kg of extractable
copper may inhibit plant growth (Oliver et al., 1996).

The concentration of copper in surface soil is enriched by the application of fertilizer, which
contains between 0.1 and 7 ppm of copper, and although copper concentrations are depleted by
plant uptake (dried wheat contains 4-8 mg/kg of copper), copper soil concentrations could
increase if the rate of uptake is insufficient to offset the effect of fertilizer application.
Concentrations of copper in various types of agricultural surface and sub-surface soils were
measured in Saskatchewan, Canada (Mermut et al., 1996). Copper concentrations were found to
be associated with soil clay and organic content, ranging from approximately 20-60 mg/kg.
Although surface soil concentrations were generally greater than sub-surface soil concentrations,
they were not significantly different.

Total and available copper concentrations were measured in surface and subsurface soils in the
region around the lower Bhavani River in India (Paramasivam & Gopalswamy, 1994). Copper
concentrations ranged from 46 to 156 g/g, and generally decreased with depth. Available
copper concentrations ranged from 2 to 6 g/g.

Copper movement through soil due to infiltration was studied in soil classified as a Podzol (acid
loamy sand with a mor humus), on the eastern ridge of Mont Blanc in Switzerland (Keller &
Domergue, 1996). The fraction of copper in the particulate fraction (< 4.5 um) was relatively
constant among the various soil horizons (17-40%), representing a lower boundary the amount of
mobile copper. Copper concentration in the various soil horizons varied from 1-20 g/kg.

The concentration of copper in soil and grass along a major highway in Alice, South Africa was
measured at various distances up to 20 m perpendicular to the road (Fatoki, 1996). Although
there is heavy traffic along the road, copper concentrations in the soil and grass was low (< 10
g/g) and did not appear to bear any relation to distance from the road. In contrast to this study, a

COPPER: Environmental Dynamics and Human Exposure Issues


Page 102
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

study in Jokionen, Finland (Ylaranta, 1995) found the rate of copper deposition at experimental
plots located 200 m from roads substantially greater than at plots 20 m distant from the road.

Copper and other heavy metals were measured in market garden soils and vegetables in Hong
Kong (Wong, 1996). The range of total copper concentrations in soil was 567 mg/kg, with a
mean total concentration of 18 mg/kg. The ranges and mean of EDTA extractable concentrations
were approximately 50% of the total concentrations. Higher soil concentrations occurred at the
more urbanized and high traffic areas, and the EDTA extractable concentration correlated with
the total concentration. The copper concentrations in vegetables grown in these soils ranged from
0.45 mg/kg to 1.16 mg/kg wet weight, and indicate regulation of copper uptake by these plants.

Copper speciation in soil is an important determinant of its bioavailability (Ritchie & Sposito,
1995). Copper may be available in the soil solution as a solvated ion, and as complexes with
organic or inorganic ligands. It may also be adsorbed onto particles containing metal hydrous
oxides, silicates, carbonates, and organic matter. Metals speciation is a function of several
factors, including pH and organic content. Copper enrichment in soil humic matter was studied
near an open brown coal mine in Latvia (Sebestova et al., 1996), run-off water from which was
acidic (pH of approximately 2.5). Copper enrichment was approximately 1000-fold in some
samples obtained from the site, far exceeding enrichment factors for all of the other heavy metals
studied.

Copper speciation in soil can vary with season and duration of contact with the soil matrix.
Seasonal variation in copper speciation was studied in the top-soil (0-5 cm) of a mixed
deciduous-coniferous forest in Austria (Wenzel et al., 1996). EDTA extractable copper varied by
35% relative to the minimum, and highest mobility was found to occur in August. The high
variability in extractability of copper could not be explained by soil organic content, which
remained relatively constant over the duration of the study.

The amount of copper and other heavy metals (Cd, Co, Ni, Pb, Zn, Fe, and Mn) added to
agricultural soils from fertilizer and pesticide application was studied in a rice farming plot in
Valencia, Spain (Gimeno-Garcia et al., 1996). Total copper concentrations in the upper (0 - 15
cm) and lower (30 - 45 cm) soil horizons were approximately 20 ppm, twice as high as that in
intermediate soil, but the extractable amount decreased monotonically with depth.
Approximately 30% of the copper in the top horizon was extractable by EDTA, whereas only 5%
of the lower horizon was extractable. Moreover, based on the extraction ratio, copper was found
to be more mobile than all the other metals in the study.

The concentration of several heavy metals and copper were measured in the soils, and several
species of plants and animals (vertebrates and invertebrates) of a woodland in Schleswig-
Holstein, Germany relatively unaffected by any local sources of contamination (Scharenberg &
Ebeling, 1996). The average concentration of copper in nine soil horizons up to a depth of 165
cm ranged from 1.8 to 5.8 mg/kg. Although no systematic accumulation of copper was found for
animal species as a function of their position in the food chain, concentrations were generally
higher in animals than in plants. The concentration of copper in litter from beech trees ranged

COPPER: Environmental Dynamics and Human Exposure Issues


Page 103
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

from 5.3 to 10.9 mg/kg, the highest concentrations being found in seeds. The concentration of
copper in homogenized samples from vertebrates was relatively uniform (9.9 - 17.4 mg/kg)
compared with those from invertebrates (18.3 - 192 mg/kg). Beetles and isopods were notable
among invertebrates for having copper concentrations approximately three- to four-fold greater
than the other invertebrate species. However, invertebrates do not appear to be endangered as the
copper concentration in beech litter was approximately ten- to fifteen-fold less than lethal
concentrations for isopods, and the soil concentrations were approximately 2 orders of
magnitude less than the LC50 for earthworms.

6.3.1 Copper at Hazardous Waste Sites


Copper and its compounds were found at 210 of 1177 hazardous waste sites on the National
Priorities List (NPL) of highest priority sites for possible remedial action. No concentration
levels were given. Since copper is found in soil, it should occur at all sites. Additionally, specific
copper compounds, namely, copper cyanide, copper hydroxide, and copper chloride, were each
identified at one site. To ascertain whether elemental concentrations at hazardous waste sites
were elevated above that which would normally be expected in soil, a background-based ranking
(BBR) technique was applied to data for metals in soil at hazardous waste sites in the 1980-1983
Contract Laboratory Program (CLP) Analytical Results Data Base (CARD). The BBR technique
finds the percentage of samples exceeding that normally found in soil at the 95% and 99%
confidence intervals. Of the 1307 samples in CARD, 10.5% and 7.3% had copper concentrations
exceeding the number normally expected in soil at the 95% and 99% confidence intervals,
respectively (Eckel & Langley, 1988).

6.4 Copper Levels in Indicator Biota


The effect of road traffic on heavy metal concentrations in wheat grain straw, rye grass, and
lettuce was studied over a period of two years at Jokionen and Nurmijarvi, Finland (Ylaranta,
1995). Copper deposition at plots 200 m from the road were greater than at plots 58 m and 20 m
distant from the road at both Jokionen and Nurmijarvi. Although the density of traffic at
Nurmijarvi (9500 cars/day) was much greater than that at Jokionen (5500 cars/day), which was
reflected in the approximately six-fold and nine-fold greater than background copper deposition
rate, respectively, copper concentrations in the plants at the two sites were not significantly
different.

Spatial distribution of heavy metals in Germany was studied using moss collected from 593 sites
as a bioindicator of heavy metal deposition. Mosses are good biomonitors of atmospheric
deposition because they obtain much of their nutrients from precipitation. Moreover, since their
leaves are only one cell layer thick, they are in close contact with the ambient atmosphere
(Ruhling, 1995). Copper concentrations in moss ranged from 4.1 g/g to 25.5 g/g, reflecting the
proximity of sources, such as copper smelters and metal working facilities. It is notable that
variation in moss concentrations is considerably less dramatic than variation in soil
concentrations found in other studies (see Markus & McBratney, 1996).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 104
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

Human consumption of trace metals, including copper, was assessed based on concentrations in
drinking water and food in Nagpur, India (Hasan, 1995). Average copper concentrations in
cereals, vegetables, and well and river water were 0.04, 0.3 g/g, 20 g/g and 35 g/g
respectively. Copper concentrations in surface and subsurface soils were approximately 11 g/g
and 8 g/g. The average intake of copper was calculated to be 3.4 mg/week, well below the
WHO recommended limit of 21 mg/week.

The effectiveness of pollution control on copper concentrations in oysters was evaluated in the
Charting coastal area in Taiwan (Lee et al., 1996). In January 1986, green oysters in the
mariculture beds in this area were found to have concentrations as high as 4400 g/g dry weight;
three months later all were dead. Following the green oyster incident, metal reclamation works
on the Erjin Chi River were shut down, and plans made to remove contaminated soil and
sediment from the area. Average copper concentrations in oysters between 1991 and 1993 were
160 to 750 g/g, significantly greater than concentrations of 12 - 90 g/g in control animals. The
maximum measured concentration, 933 g/g, was considerably less than that during the green
oyster incident. The concentration of zinc and copper in oysters appears to be closely related, as
evidenced by the small change in the oyster concentration ratios (approximately 3 - 6) compared
to the range of concentration ratio in the particulate phase (approximately 3 - 12).

The concentration of copper in imported and local shrimp were measured in Saudi Arabia, and
were found to range between 0.16 and 7.96 mg/kg wet weight (Sadiq et al., 1995). This range
represents worldwide variability, since the imported shrimp samples came from several
geographical areas, and were harvested at various seasons.

The biotic and abiotic factors influencing trace metals concentration, including copper, were
examined in freshwater isopods obtained from 28 areas in the Netherlands (Van Hattum et al.,
1996). Copper concentrations ranged from 86 to 176 g/g. Within and between populations,
copper concentrations variability accounted for 33% and 62% of total variability, indicating that
abiotic factors are more important determinants of copper concentration than biotic factors.

The concentration of copper in the tissues of mullet from the Tigris River in Turkey were
investigated (Unlu et al., 1996). Average tissue concentrations ranged from 23 g/g in muscle to
175 g/g in liver.

The effect of wood preservatives applied to docks in South Carolina tidal creeks on sediments
and oysters in the vicinity was investigated (Wendt et al., 1996). Although oysters growing on
the dock pilings had significantly higher concentrations of copper (226 g/g dry weight) than
those growing 10 m distant (134 g/g dry weight), no physiological differences were observed
between the groups. Sediment concentrations immediately adjacent to docks also had copper
concentrations approximately three-fold higher than those from sites greater than 10 m distant.
Four-day field bioassays measuring the survival of mud snails, mud minnows, red drum, and
white shrimp also did not show any difference between the sites immediately adjacent to the
docks and site >10 m distant. The influence of treated dock piling wood appears to be localized

COPPER: Environmental Dynamics and Human Exposure Issues


Page 105
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

to a radius of less than 10 m, and even within this area, elevated concentrations appear to have no
effect on oysters.

The ability of microalgae from the Sea of Japan to monitor heavy metal concentrations in marine
waters was investigated (Zolotukhina et al., 1993). Two microalgae species, chosen as
biomonitors on the basis of laboratory studies, indicated that copper concentrations at three
sampling stations increased over the period 1989-91.

Copper and other trace element accumulation in arctic marine ecosystems was investigated by
examining tissue levels in three species of birds at Hornoya, Norway (Wenzel & Gabrielsen,
1995). The average copper concentration in bird feathers, liver, kidney, and muscle were similar
for all three species, as well as in juvenile birds of one of the species, despite differences in the
feeding habits of these birds. The results indicate that copper concentrations are well regulated in
birds living in the arctic.

Seasonal variations in copper and other heavy metal concentrations in bivalves in the Sea of
Japan, near Vladivostok was investigated (Shulkin & Kavun, 1995). Copper concentrations
increased between February and April, and subsequently decreased as temperatures increased.
Increases in copper concentration were attributed to increased filtration, and the subsequent
decrease was attributed to increased metabolic activity and physiological regulation concomitant
with increased ambient temperature.

Fish
As a part of the National Contaminant Biomonitoring Program of the U.S. Fish and Wildlife
Program, eight species of freshwater fish were collected at 112 stations in the United States in
1978-1979 and 1980-1981 (Lowe et al., 1985). The geometric mean concentrations of copper in
g/g (wet weight) for the two years were 0.86 and 0.68; the 85th percentiles were 1.14 and 0.90
and the ranges were 0.29-38.75 and 0.25-24.10, respectively. The highest concentration, 38.75
and 24.10 g/g, during both collecting periods was in white perch from the Susquehanna River
and the second highest concentration, 19.3 g/g, was found in white perch from the Delaware
River near Trenton, NJ.

Copper residues in muscle of 268 fish specimens were analyzed over a five-year period in
several surface water systems in eastern Tennessee (Blevins & Pancorbo, 1986). The mean
residue levels in the muscle of different species of fish from nine stations ranged from 0.12-0.86
g/g (wet weight). Maximum levels ranged from 0.14-2.2 g/g.

Mean and median copper concentrations of 127 samples of edible fish from Chesapeake Bay and
its tributaries were 1.66 and 0.36 g/g in 1978, and 1.85 and 0.61 g/g in 1979 (Eisenberg &
Topping, 1986). Copper levels were increased in the livers and to a lesser degree, the gonads,
compared with the flesh. The copper content of muscle tissue of several species of fish collected
from metal-contaminated lakes near Sudbury, Ontario, ranged from 0.5-1.4 g/g (dry weight).
No major pattern in variation was evident among species or among the study lakes (Bradley &

COPPER: Environmental Dynamics and Human Exposure Issues


Page 106
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

Morris, 1986). The copper concentration in the livers, however, ranged from 5-185 g/g (dry
weight) and differed significantly among species and among lakes. Unlike muscle tissue, liver
tissue is a good indicator of copper availability, although the data indicate that there are other
factor(s) influencing copper availability and bioaccumulation in these fish.

Aquatic Invertebrates
Copper concentration in the soft tissue of mussels and oysters collected as part of the U.S.
Mussel Watch Program in 1976-78 ranged from 4-10 g/g (dry weight) for mussels and 25-600
g/g for oysters (Goldberg, 1986). Copper concentrations in mussels collected from eleven sites
near Monterey Bay, CA, were 4.63-8.93 g/g (dry weight) (Martin & Castle, 1984). Perwack et
al. (Perwak et al., 1980) reported similar results for mussels (3.9-8.5 g/g) and clams (8.4-171
g/g).

Plants
Although copper concentrations in plants vary widely, they usually range from 1-50 ppm (dry
weight) (Davies & Bennett, 1985; Perwak et al., 1980). Adding lime to the soil to increase pH to
7 or 8 reduces copper availability to plants.

6.4.1 Sewage Sludge


In an EPA-sponsored study to determine metal concentration in sewage sludge, copper
concentrations in primary sludge at seven POTWs were reported to be 3.0-77.4 g/g, with a
median concentration of 20.5 g/g. The plant with the highest copper concentrations received
wastes from plating industries, foundries, and coking plants. In a comprehensive survey of heavy
metals in sewage sludge, 30 sludges from 23 American cities were analyzed (Mumma et al.,
1984). Copper concentration in the sludges ranged from 126-7729 g/g (dry weight), with a
median value of 991 g/g. The proposed limit for copper in sludge spread on agricultural land is
1000 g/g (Mumma et al., 1984). For comparison, the concentration of copper in cow's manure
is ~5 g/g (Mumma et al., 1984). Additional information for copper levels in sludge can be
found in a recent ICA report (Landner et al., 2000).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 107
Chapter 6
COPPER LEVELS IN THE ENVIRONMENT AND HUMAN TISSUES:
OVERVIEW OF SELECTED FIELD STUDIES

COPPER: Environmental Dynamics and Human Exposure Issues


Page 108
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

7 ASSESSING HUMAN EXPOSURE, DOSE AND RISK

7.1 Exposure Potential and Pathways


In discussing exposure to copper, the important question is whether individuals are exposed to
readily available copper, which in general, means free (hydrated) copper(II), and perhaps some
weakly complexed or adsorbed forms of copper. Available data indicate in general that copper in
natural water, sediment, and soil is not in a labile form. Potential for high exposure of the general
population to copper may exist where people consume large amounts of tap water that has picked
up copper from the distribution system. This is most likely to occur where the water is soft and is
not allowed to run and flush out the system. In such cases, copper concentrations frequently
exceeds 1 mg/l, a large fraction of which may be free cupric ion, resulting in exposure by
ingestion and dermal contact.

Based on available data, copper at National Priorities List (NPL) sites is not expected to be
hazardous to people living close to the sites or to clean-up workers. Data suggest that the copper
at these sites would not in general be in a labile form and should not leach into groundwater.

People living near copper smelters and refineries and workers in these and other industries may
be exposed to high levels of copper in dust by inhalation and ingestion. In some industries,
workers may be exposed to fumes or very fine dust that may be more hazardous than more
typical dust.

A key issue in exposure and health risk assessment is the need to evaluate exposure/intake from
all possible sources, routes and pathways. For example, the main route of exposure to copper is
oral, and food and water, the predominant sources. Inhalation exposure from polluted air may
occur in the workplace, especially in mining and agricultural work where copper salts are used as
pesticides. The effect of airborne copper may be of particular importance because of direct
effects on the lung. However, for most practical purposes, the oral route is the only one of health
significance in terms of general population exposures to copper. Food may account for over 90%
of copper intake in adults where water has low copper content (< 0. 1 mg/l). If water copper
content is high, 1-2 mg/l, it may account for close to 50% of total intake. In the case of infants
consuming copper-supplemented artificial formula, the contribution of water may be less than
10%, whereas, if the formula is not fortified with copper, water may contribute over 50% of total
copper intake, especially when water copper content is in the high range of 1-2 mg/l.

7.2 Environmental Exposures

7.2.1 Inhalation
Everyone is exposed to copper in atmospheric dust. Estimates of atmospheric copper
concentrations from representative source categories yielded a maximum annual concentration of
30 mg/m3 (USEPA, 1987; ATSDR, 1990). If a person is assumed to inhale 20 m3 of air/day, this
would amount to an average daily intake of 600 mg of copper. For the reported range of annual

COPPER: Environmental Dynamics and Human Exposure Issues


Page 109
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

atmospheric copper concentrations, 5-200 ng/m3, the average daily intake by inhalation would
range from 0.1-4.0 mg. At a maximum reported ambient air concentration, 100 mg/m3 for a 24-
hour period at a location within one-half mile of a major source, the average daily intake would
rise to 2000 mg. However, these estimates assume that all the copper is attached to particles of
inhalable size, which is usually not the case.

Tobacco
The mean copper content of cigarette tobacco was 24.7 ppm, with a standard deviation of 10.8
ppm (Mussalo-Rauhamaa et al., 1986). However, only 0.2% of this copper passes into
mainstream smoke.

7.2.2 Ingestion
Ingestion exposures are dominated by copper in drinking water and food and are considered in
more detail under a following section. The average daily dietary intake of copper from food is <2
mg/day. Assuming a median copper concentration in drinking water of 75 mg/l, the average daily
copper exposure from consumption of 2 liters of water per day is 0.15 mg; however, many
people may have high levels of copper in their tap water from the water distribution system. If
the system is not permitted to flush out, average intakes from water may in fact be >2 mg/day. It
is less likely that high dermal exposures will result from bathing in this tap water because the
distribution system will flush itself out as the water is drawn. Airborne dust containing copper
may also contribute to ingestion exposures for people living near copper smelters and refineries
and for workers in these and other industries.

7.2.3 Dermal
A less likely situation where exposure to high levels of free copper(II) may occur is from
swimming in water that has been recently treated with a copper-containing algicide. In natural
waters, the level of free copper would be expected to decrease rapidly. Soluble cupric salts are
used extensively in agriculture and in water treatment. Workers engaged in the production and
application of these chemicals, as well as industrial workers, such as those in the plating
industry, may come into dermal contact with these copper-containing compounds.

7.3 Occupational Exposures


A National Occupational Exposure Survey (NOES) conducted by NIOSH from 1981- 1983
estimated that potentially 505,982 workers, including 42,557 women, are occupationally exposed
to copper in the United States. Interpreting the significance of the NOES is particularly difficult
for copper because we are all exposed to copper; furthermore, the survey does not indicate the
level or form of copper to which the worker may be exposed. The NOES estimate is provisional
because all of the data for trade name products that may contain copper have not been analyzed.
Of the potential exposures, 1073 are to pure copper, while in the other cases, the molecular form
of copper was unspecified. Additionally, according to the NOES, 125,045 workers, including

COPPER: Environmental Dynamics and Human Exposure Issues


Page 110
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

38,075 women, are potentially exposed to copper sulfate. The NOES was based on field surveys
of 4490 facilities and was designed as a nationwide survey based on a statistical sample of
virtually all workplace environments in the United States where eight or more persons are
employed in all standard industrial codes (SIC), except mining and agriculture. The exclusion of
mining and agriculture is significant for estimating exposure to copper because of their high
potential for exposure.

Studies of workers that include quantitative measurements of ambient and personal exposure
(i.e., breathing zone) have been performed on U.S. copper smelter employees. In a 1995 study
(reported in Nriagu, 1979b), Wagner described conditions of exposure in fourteen U.S. copper
smelter facilities that were the subject of industrial hygiene surveys conducted by NIOSH.
Measurements of airborne concentrations of As, Pb, Zn, Cd, Mo, copper, and sulfur dioxide
revealed that only the last two occurred consistently at relatively high levels. For copper, the
industry-wide average for both area and personal samples exceeded 1 mg/l for dust and fumes in
several plant locations. The current federal standard for occupational exposure to copper restricts
in-plant exposure to concentration of 1 mg/m3 for dust and 0.1 mg/m3 for fume emissions.
However, it has been found that most of the element typically occurred as a non-respirable
dust, and that this dust consists mainly of a relatively inert copper sulfide.

Urinary levels of copper were determined in the smelter worker studies, but Wagner indicates
that methods of collection and analysis were not well controlled for potentially confounding
factors. These reservations notwithstanding, a mean urinary copper content of 79 g/l
determined from 206 individual samples was well above the 9 to 18 g range suggested as
normal for 24-hr urinary excretion in human beings. Nonetheless, evidence of copper toxicity,
acute or chronic, was not revealed in the smelter survey despite excessive ambient, personal, and
urinary levels of copper. Perhaps this relates to the absence in these facilities of free metallic
copper and copper oxide, the chemical forms most often associated with the so-called metal
fume fever of copper and other types of metal exposures.

It should be noted that arsenic contaminants in copper ore have been considered a likely cause of
the increased risk in pulmonary carcinoma in smelter worker populations and in non-occupa-
tionally exposed residents of at least one copper-mining city.

In general, in epidemiologic studies, as in the U.S. copper smelter studies, health hazards related
to this metal have been found to be of low order. However, over time various questions have
been raised regarding unrecognized, occupationally related, copper-induced illness. Of
considerable importance in this regard was the discovery of serious to fatal pulmonary and
hepatic abnormalities in a group of Portuguese vineyard sprayers exposed to a copper
sulfate-containing fungicide for 3 to 45 years. Such evidence concerning long-term copper
toxicity could imply that it would be prudent to reserve conclusions about the health-hazard
implications of prolonged exposure until critical longitudinal studies of worker health become
available.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 111
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

7.4 Dietary Exposures to Copper: Drinking Water and Food

7.4.1 Copper in Food


Estimates of dietary intake by Americans prior to 1970 were considerably higher than current
intake estimates. This however mostly reflects marked improvements in analytical techniques for
measuring copper and awareness of the importance of avoiding copper contamination of
analytical samples. The usual diet was thought to contain 2 to 5 mg of copper, but studies,
including one study of 132 diet composites, now show that few diets contain over 2 mg per day.
As with all nutrients, copper intake can vary widely, depending on food choices. Diets in
countries where more whole-grain products, legumes, and organ meats are eaten, contain more
copper (Turnlund, 1999).

The amount of copper in the daily diet sufficient for health is in the range of 20-60 g/Cu/kg/day
for most children and adults. The needs of newborn infants are 40-150 g/Cu/kg/day; premature
babies may need 150-400 g/Cu/kg/day and malnourished children may need over 500
g/Cu/kg/day for recovery. The question of how much copper is too much is more difficult to
answer. The upper safe limit of around 200 g/Cu/kg/day is safe for most adults and children,
but formula-fed children may exceed this and the needs of preterm and marasmic infants are
much higher (Ralph & McArdle, 2001).

The richest sources of dietary copper contain from 0.3 to over 2 mg/100 g (50 to >300 nmol/g).
These include shellfish, nuts, seeds (including cocoa powder), legumes, the bran and germ
portions of grains, liver, and organ meats. Most grain products; most products containing
chocolate, fruits and vegetables, such as dried fruits, mushrooms, tomatoes, bananas, grapes, and
potatoes; and most meats have intermediate amounts of copper, from 0.1 to 0.3 mg/100 g (20-50
nmol/g). Other fruits and vegetables, chicken, many fish, and dairy products contain relatively
low concentrations (<0.1 mg/100 g [20 nmol/g)]) of copper. Cow's milk is particularly low in
copper. The major sources of copper in the US diet are meat, nuts, beans/peas, and main dishes.

Because information on the copper content of foods is incomplete and databases often contain
missing values, copper intake is underestimated unless missing values are replaced with imputed
values. A table of the copper content of foods, compiled when much of the available data were
from the 1930s and 1940s, reported consistently higher copper concentrations than tables that
exclude pre-1960 data, suggesting that early values are too high. However, a critical evaluation
of the reliability of post-1960 published values for the copper content of foods demonstrated that
improvement is still needed. The copper content of foods can vary owing to combination of
factors, including analytical method, sampling procedure, recipe, cooking method, and the area
of the country from which foods are collected. Careful analysis of the copper content of
high-copper foods from the FDA total diet study varied by an average CV (coefficient of
variation) of 24%, the mid-range of average variation of other minerals in the same foods.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 112
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

Copper Food Intake for Different Age Groups


Daily intakes of copper and other essential minerals were estimated for eight age-sex groups of
the United States population as part of the FDA's Total Diet Study (Pennington et al., 1986). By
analyzing the mineral content of composite samples of 234 foods obtained in 24 cities from mid-
1982 to mid-1984 and by using previously determined daily intakes of each food, the daily
mineral intake for the age-sex groups was determined. The copper intakes in mg/day of the eight
age-sex groups were: 6- to 11-month-old infant, 0.47; 2-year-old child, 0.58; 14- to 16-year-old
girl, 0.77; 14- to 16-year-old boy, 1.18; 25- to 30-year-old woman, 0.93; 25- to 30-year-old man,
1.24; 60- to 65-year-old woman, 0.86; 60- to 65-year-old man, 1.17. All values were low in
terms of the estimated safe and adequate daily dietary intake of this nutrient. The food item with
the highest copper level was beef/calf liver (61 g/g).

Adequacy of Intake by Various Population Groups


Dietary copper intake in the United States and other countries is usually below the currently
recommended 1.5 mg/day. The new World Health Organization (WHO) estimated minimum
requirements are 0.6 mg/day (9 mol/day) for women and 0.7 mg/day (11 mol/day) for men.
The estimated usual dietary copper intake in the United States, after correcting for missing
values in the USDA database, was 0.96 mg/day (15 mol/day) for women and 1.4 mg/day (22
mol/day) for men. A total diet study estimated intake at 0.92 mg/day (14 mol/day) for women
and 1.2 mg/day (19 mol/day) for men (Turnlund, 1999).

Copper in Mothers Milk


A baseline value for the copper content of mother's milk was determined by screening literature
values. The 28 samples selected had copper concentrations ranging from 197-751 g/l and a
median of 290 g/l (Iyengar & Woittiez, 1988). In a separate study, it was found that the
variability was primarily subject-related, but for individuals, the copper concentration in milk
declined moderately with duration of lactation (Vaughan et al., 1979).

A Review of Recent Studies of Copper in Food


North American wild rice grown in soft, acidic wetlands was analyzed for copper levels, as well
as for other trace metals (Nriagu & Lin, 1995). The intent of the study was to examine the
possible environmental impact of increasing levels of toxic metals due to possible industrial
contaminants.

Wild rice is produced in North America via paddy culture in shallow water along riverbeds
and marshy areas. The ideal conditions for production are acidic substrates high in organic
matter. Concentrations of trace metals were detected via graphite furnace atomic absorption
spectrometer with a Zeeman background corrector. The environmental exposure of the plants to
the metals and the transfer of the metals to the rice grains was not demonstrated by any particular
pathway. It was determined that the wild rice may be a good dietary source of copper.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 113
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

Since copper is an essential element needed by the body for various metabolic processes,
including normal blood cell formation, the presence of copper in food is an important
consideration in dietary studies. Native grown fruits, vegetables and other selected foods in a
semi-arid region of West Africa were compared to imported foods for mineral content including
copper (Smith et al., 1996). The native plants usually exhibited higher levels of copper than the
imported foods. Most of the native species of the various plants found to contain higher copper
levels than the imported species, with some plants containing up to nine times the levels.

Market garden soils and vegetables in Hong Kong were monitored for heavy metal content and
to determine whether these levels were affected by human activity (Wong, 1996). With
increasing urbanization and increased traffic releasing heavy metals into the atmosphere, urban
garden soils accumulate increasing amounts of heavy metals available for plant uptake. EDTA
extractable copper was more prevalent in more highly urbanized areas. Correlations were found
between total acid soluble and EDTA extractable copper and zinc. Low soil pH appeared to
increase metal availability. There was a two- to three-fold increase in sorption capacity of the
soils for each one-unit pH increase. Lower soil organic carbon content reduces sorption of heavy
metals from anthropogenic sources, thus increasing soil metal availability. Another factor
increasing metal availability in soil is low cation exchange capacity. The combination of low pH,
decreased cation exchange capacity, low organic matter, and coarse textural soil increased the
possibility for plant uptake through the roots owing to increased metal availability. All vegetable
samples analyzed contained copper concentrations below maximum permitted concentrations.

The effect of road traffic in southern Finland on heavy metal levels in plants was examined
(Ylaranta, 1995). There was variation from plant to plant, the highest levels of zinc and copper
being found in grain. Since different parts of the plant accumulate the metal in different
concentrations, clean soil pots were used and placed at different distances from the road to
enable researchers to determine what amount of the metal was due exclusively to atmospheric
exposure as opposed to root uptake. At current traffic density at the time of the study (1985-87)
copper concentrations were found to be low.

In Romania, the effect of smelter emissions on soil concentration and plant uptake of heavy
metals was examined (Lacatusu et al., 1996). High soil acidity enhances copper transfer to the
growing plant. Copper pollution was classified as moderate. The edible portions of various
vegetables were analyzed. Significant elevated levels of copper as compared to maximum
allowable concentrations were present in various fresh and, especially, dried vegetables. The
focus of the study was primarily on lead and cadmium.

In Spain, dairy products were analyzed for levels of a variety of heavy metals, including copper.
The highest copper levels were found in milk and increased copper concentrations were found in
foods packaged in glazed ceramic containers.

In vitro sample digestion using simulated gastric juice was used and the digested solutions were
analyzed for copper using GF-AAS. Copper concentrations in milk were in accordance with the
supplement concentrations added to the product. Dairy products packaged in the glazed

COPPER: Environmental Dynamics and Human Exposure Issues


Page 114
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

containers had increased levels of copper, especially acidic products like curd. Cocoa may also
supply increased copper levels to foods containing it.

Analysis of metals levels in food is not sufficient to determine dietary requirement levels.
Bioavailability also has to be taken into consideration. The in vitro digestion procedure makes it
possible to calculate the percentage of the element available for uptake by the digestive tract.
Zinc will have a negative effect on copper bioavailability, as will certain carbohydrates and
ascorbic acid.

Heavy metal levels, including copper, were measured in mullet, a fish that, after spawning,
migrates to the Tigris River in Turkey (Unlu et al., 1996). The Tigris is located near rich copper
ore sites in Southeast Anatolia. Wastes from a copper plant are discharged into the river and
elevated copper levels are found in the water, the sediment, and samples of fish living in the
waters.

Different metal levels were found in different fish tissues; the highest copper concentrations
were found in liver. The liver contains metallothioneins that help store metals that need to be
detoxified. Increased copper levels in mullet were linked to contamination of the Tigris by the
copper plant.

Twenty-three trace element levels, including copper, were measured in wheat, rice, and pulses in
Pakistan (Qureshi et al., 1995). The data collected were to be used as baseline values in order to
compare future values with possible contamination events. Rice was found to have greater
copper content than wheat but the levels of other elements in rice were similar to that of wheat.
Over a six-year period, a decrease in trace element levels was found in wheat flour.

There are numerous measurements available in the 1998 Source Book of copper levels in a
variety of plants. However, there are few systematic analyses available or retrievable placing the
results within the context of food basket surveys designed to assess exposure. APPENDIX A
provides a compilation of data on the levels of copper found in various food products. These are
available for use as assumed concentrations in dietary exposure assessments.

7.4.2 Copper in Drinking Water


Copper concentrations in drinking water vary widely as a result of variations in pH, hardness of
the water supply, and copper picked up in the water distribution system (Davies & Bennett,
1985). Copper concentrations in drinking water range from a few g/l to 10 g/l. A Canadian
national survey of copper and other metals in drinking water was conducted from November
1976 to January 1977 (Meranger et al., 1979). Supplies from 70 municipalities representing 38%
of the Canadian population were included in the survey, including 50 derived from river or lake
water and 20 derived from groundwater. Unfiltered raw, treated, and distributed drinking waters
were analyzed. Whether the water was derived from river, lake, or well water did not
significantly affect the copper concentration in the raw water. Only in a few supplies did copper
levels in raw water exceed 20 g/l, and only one of these was derived from groundwater. The
results in groundwater contrast with those of Page (Page, 1981) in New Jersey, in which over

COPPER: Environmental Dynamics and Human Exposure Issues


Page 115
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

100 wells contained copper levels in excess of 64 g/l. However, that study also included
groundwater that was not a source of drinking water. The copper concentration in Canadian
treated water was generally ~10 g/l (Meranger et al., 1979). In 20% of the samples, the copper
level in distributed water was significantly higher than the treated water; the increase was greater
in areas where the water was soft and corrosive, thus enhancing leaching of copper from the
distribution system. A study of 1000 water samples from random households in Ohio found that
~30% contained copper levels >1 g/l (Strain et al., 1984). The highest copper level in the study
was 18 g/l. In a study of private water wells in four communities in Nova Scotia, Maessen et al.
(Maessen et al., 1985) found that the concentrations of copper increased in water that remained
in the distribution system overnight, indicating that copper was mobilized from the distribution
system. Whereas the level of copper in running water was generally very low, that in the
standing water was variable and exceeded 1.0 g/l in 53% of the homes. Correlation with pH and
nitrate, chloride, and manganese concentration accounted for >99% of the copper picked up from
the distribution system. Similar results were reported for United States cities (Maessen et al.,
1985; Strain et al., 1984; Schock & Neff, 1988). In a study in Seattle, WA, mean copper
concentrations in running and standing water were 0.16 and 0.45 g/l, respectively, and 24% of
the standing water samples exceeded 1.0 g/l (Maessen et al., 1985). The difference in copper
levels between standing and flushed systems became evident at pH 7 and increased with
decreasing pH (Strain et al., 1984).

Lewis (Lewis, 1995b) provides a table with copper concentrations in drinking water from a
number of sources. His evaluation of these data lead to the conclusion that when considering the
possible detrimental effect of copper in water distribution systems, two factors must be
considered:

(a) The chemical condition of the water entering the tubing and conditions within the tubing
play important roles in determining how much copper will be extracted from the tubing.
pH and water hardness are two key factors; soft water or water with a low pH, is more
corrosive than hard water, or water with a high pH. The upper threshold levels given by
the World Health Organization (1993) are a pH of 6.5 and hardness of 60 mg/liter as
CaCO3. Hard water (above approximately 200 mg/liter) has been associated with the
formation of a protective scale within distribution systems.

(b) Detrimental effects of copper occur when there is an excess of copper in a biologically
available state. In most cases, the concentration of total copper in the water is not an
expression of the amount of copper that will have a biological effect. Measurement of
biologically available copper (e.g. Cu2) or its estimation by metal speciation models will,
at present, provide the best estimate of potential impact. (Estimation with metal
speciation models must be made with an adequate hydrographic database.)

The conclusion that hardness is protective is however quite controvesial. Schock (Schock, 2001)
has presented extensive evidence against the conventional wisdom in water treatment, i.e. that
high alkalinity and hard waters form protective coatings of calcium carbonate on pipe surfaces.
He argues that, for example, white deposits on lead pipes have been mistaken for calcium
carbonate for years, though those deposits are almost always hydroxycarbonates, carbonates and

COPPER: Environmental Dynamics and Human Exposure Issues


Page 116
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

hydroxy carbonate/sulfate mixes. Calcium carbonate coatings are true in rare cases (usually
water systems fed by lime softening treatment plants and very high pH). However, in reality,
high concentrations of dissolved carbonate species in water attack copper and promote
cuprosolvency. Only after many years - perhaps 30, 40 or more - can protective coatings of
basic copper carbonate (malachite) eventually evolve. This aging phenomenon is discussed in
a number of articles and reports (Edwards et al., 1996; Ferguson et al., 1996; Lytle & Schock,
1996; Lytle & Schock, 1997; Lytle & Schock, 2000; Schock et al., 2000; Schock et al., 1995).
US compliance data for the Lead and Copper Rule (LCR), show (Schock, 2001) that the two
highest groupings of elevated copper levels occur in the obvious low pH problem water areas
(New England, Pacific Northwest, etc.) but also in the hard groundwater areas of the Midwest
and West, such as Nebraska.

The interplay of pipe age, carbonate concentration (related to alkalinity) and pH are fundamental
in explaining cuprosolvency in oxic systems, but aside from pH, they are rarely examined by
health-effects researchers.

High copper levels have been reported in first-flush drinking water in new copper plumbing
systems (Singh, 1990; Singh & Mavinic, 1990). Corrosion of older copper tubing is also reported
to increase copper concentrations in drinking water and thus impair water quality (Lopez de Sa,
1989). Bacteria and other microorganisms, as well as certain organics, can mediate corrosion of
copper and copper-nickel piping systems (e.g. Bremer & Geesey, 1990; Gianotto et al., 1989;
Wagner et al., 1987). Volume 1 of WHOs Guidelines for Drinking-Water Quality (WHO,
1993) gave a health-based guideline value of 2 mg/liter. The authors commented, however, that
In view of the remaining uncertainties regarding copper toxicity in humans, the guideline value
is considered provisional.

Reservoir sediments have been shown to be a source of metals (including copper) for drinking
water (Schintu et al., 1989) as has groundwater in some industrialized areas (Zahn, 1988). Brass
valves and fittings have also been implicated as a potential source of lead, copper and zinc in
drinking water (Schock & Neff, 1988). A study in Sweden (Nordberg, 1990) noted that acidic
water may dissolve copper from piping systems but that neutralization can reduce the effect and
he stated that In Sweden, a limit for copper in acid water has been set at 3,000 g/l.
Neutralization is recommended when this limit is exceeded. A study in The Netherlands. (Sloof
et al., 1989) reported that, at least for that country, the copper load in drinking water posed no
health hazard in terms of toxicity, but also noted (abstract) that On the other hand, there are
groups of people with an insufficient intake of copper.

Reports of excess copper in drinking water affecting human health are not common. The effects
are generally metallic taste, nausea and, in the extreme case, vomiting (e.g., Spitalny et al.,
1984). An occurrence of diarrhea in children in seven new kindergartens in Sweden was traced to
copper from waterpipes and the use of an ion exchange water softener (Berg, 1987). Bypassing
the water softener and flushing the drinking water lines eliminated the problem. Active removal
of copper from drinking water is also possible (Nelson, 1987). Some drinking water treatment
processes remove trace metals, a factor considered by the U.S. Environmental Protection Agency
when evaluating water treatment processes. Desalination is an example where metal removal

COPPER: Environmental Dynamics and Human Exposure Issues


Page 117
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

often takes place, for example, a study (Peplow & Vernon, 1987) discussed the removal of
copper from reverse osmosis membranes used in seawater desalination. Copper levels measured
at taps in various types of buildings served by a desalination plant in Dhahran (Saudi Arabia)
ranged from less than measurable to more than 2.5 mg/liter (Alam & Sadiq, 1989). Since water
at the desalination plant had less than measurable copper levels, the increased copper must have
leached from the distribution pipes. Copper corrosion in potable water systems can often be
reduced by the addition of a small amount of phosphate (Miller et al., 1988). Another study
(Reiber, 1989) found that orthophosphates significantly reduced corrosion rates on copper
plumbing surfaces.

It has been suggested that Indian childhood cirrhosis (ICC) may be a result of hypersensitivity to
excess copper (Nordberg, 1990). The expression of symptoms has been associated with elevated
levels of copper in drinking water in certain areas (Eife et al., 1987; Madsen et al., 1990;
National Technical Information Service, 1989). The disease was related to hepatic copper
overload in acidic (pH 5.0-6.2) well water that had passed through copper tubing. The copper
level in a particular well was given as 12 g/l, while the level was 3400 g/l after passing
through a copper distribution system (maximum permitted level in Germany was 3000 g/l in
the 1987-1989 period). Contrasting to this are other cases of ICC reported in Australia, Kuwait,
Singapore, and North America in which no linkage to excess copper intake in drinking water (or
otherwise) could be identified (e.g., Adamson et al., 1992; Aljajeh et al., 1994; Lefkowitch et al.,
1982). In addition, an examination of official medical records in the three Massachusetts, USA,
towns having the highest first-draw copper levels at the tap (in excess of 8000 g/l) for medium-
sized towns, determined that, in over 23 years, corresponding to 64,124 child years of exposure
of children under the age of six, there were 135 deaths among these children. However, none of
these deaths were from cirrhosis or any form of liver disease that could be attributed to copper
overload (Scheinberg & Sternlieb, 1994). In discussing this analysis, in conjunction with an
examination of seven reported cases of infant deaths in the U.S. from copper-related liver
disease, where the exposures were much lower than at the Massachusetts towns, they concluded
that genetic susceptibility was critical for copper effects to occur.

A Note On Potable Water Plumbing Systems (from Lewis, 1995b).


Copper tubing is accepted as the standard plumbing material for both potable water, and heating
systems in many countries (Symons, 1986). This is because copper is durable and easy to use and
has excellent thermal properties. Copper is also biostatic and inhibits the growth of organisms in
water systems. Water in distribution systems will contain some copper, from natural sources as
well as from the tubing. The contribution of copper from natural sources will vary, but is usually
only a few g/l (World Health Organization, 1993). Some additional copper can be introduced
into the water from the tubing. As discussed earlier, that amount depends on the chemical nature
of the water and the physical and chemical nature of the tubing or piping as well as the length of
time the water is in contact with the tubing. Details about plumbing materials and drinking water
quality are discussed in a series of articles (Sorg & Bell, 1986).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 118
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

Besides being an essential element, copper is known for its ability to inactivate bacteria (Belay &
Daniels, 1990; Lebedev et al., 1989; Schoenen & Schlomer, 1989; Versteegh et al., 1989).
Copper was used to sterilize water by the Egyptians before 2,000 B.C. and today is effective in
reducing the activity of coliform bacteria under conditions similar to those found in drinking
water (Domek et al., 1987). This can be beneficial in recreational and potable water (Landeen et
al., 1989b; Pyle et al., 1992; Yahya et al., 1989; Yahya & Straub, 1990) for effective control of
coliform bacteria (e.g. Jana & Bhattacharya, 1988). The coliform bacterium Escherichia coli
exhibits reduced oxygen utilization when exposed to copper in carbonate buffer that simulates a
drinking water environment (Domek et al., 1987). Copper from plumbing has been reported to
inhibit growth of Legionella pneumophila (a bacterium which causes the respiratory illness
Legionnaire's disease) in warm water systems of hospitals and hotels in Lower Saxony.
However, other studies (Schulze-Robbecke et al., 1990) found no influence of copper pipes on
Legionella under the conditions found in a hospital hot water system. There is a report (Landeen
et al., 1989a) of reduced activity of Legionella with added electrolytically generated copper and
silver ions (400:40 g/l copper:Ag). Comparable results have also been reported (Lechevallier et
al., 1990). Another study (Nuttall, 1990), in a discussion of methods for controlling bacterial
growth including Legionella, states that copper is probably the only common plumbing material
which has the ability to suppress bacterial growth. Copper in drinking water can also reduce
dental plaque formation and dental caries (Afseth et al., 1986).

Several researchers have compared the ability of tubing of different compositions to inhibit
microbial growth. Adhesion of bacteria is reduced on copper surfaces (Branting et al., 1988)
when compared with many other commercially-available surfaces, which means that the reduced
biofilm makes the tubing less likely to harbor microorganisms. Other studies (Pedersen et al.,
1986) note that polyethylene tubing in municipal drinking water systems exhibited much higher
bacterial attachment than did copper tubing and inert glass surfaces. Little if any microbial
colonization (including Legionella) was found on copper, glass, high grade steel and PTFE water
pipes and hoses when compared with PVC, PE, PA, silicone and rubber tubing (Schoenen &
Wehse, 1988; Schoenen et al., 1988). In contrast, bacterial colonization of copper pipes has also
been reported in the literature (Tuschewitzki, 1990). Another report (McFeters & Singh, 1990)
provides evidence that injured enteropathogenic bacteria can retain virulence. It is important to
recognize that benefical as well as detrimental effects are affected by water/metal chemistry and
the copper tolerance of the pathogen.

Copper in Drinking Water: Review of Recent Field Studies


Trace metal levels were measured in drinking water samples to determine the effect of various
plumbing materials in contributing to the presence of the metals (Viraraghavan et al., 1996).
Factors that were considered in this Canadian study were age and type of plumbing materials,
building height and type of building. Water sources supplying the city were of two different
types: groundwater (aquifer and wells) and surface water from a lake.

In low-rise buildings with copper plumbing that was five years old or less, copper and zinc were
found leaching from both the faucets and the plumbing system. First and second draw samples

COPPER: Environmental Dynamics and Human Exposure Issues


Page 119
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

showed higher copper levels than did the third draw. A similar result was seen for buildings
where the copper plumbing was ten years old or older. In low-rise buildings with lead plumbing,
lower copper levels were found than in the buildings with copper plumbing. Studies of high rise
buildings indicated that copper found in water samples was due to leaching from the water
faucets, and that negligible contributions were due to pipe corrosion. All of the high rise
buildings in the study had copper plumbing. In India, a heavy metal pollution index was used to
estimate heavy metals concentration in drinking water (Mohan et al., 1996). The pollution index
may also be used to compare quality characteristics at different stations.

Copper concentrations in drinking water were studied in Riyadh, Saudi Arabia, where the
sources of drinking water are deep groundwater and desalinated sea water (Al-Saleh, 1996). The
study was concerned with determining whether copper concentrations were in excess of ECC
guidelines of 100 g/l. Copper concentrations ranged from 2-106 g/l, with a mean of 18 g/l.

Concentrations of heavy metals in tap water at various locations in Regina, Canada were
investigated as a function of building height, age, type of plumbing, and water chemistry
(Viraraghavan et al., 1996). Copper concentrations in the first draw exceed those in the second
and third draws at a sampling site, consistent with the known tendency of copper to leach from
plumbing. In several cases the concentration in the first draw exceeded the recommended
maximum concentration level of 1000 g/l. However, in this case there appeared to be no
significant effect of building height, age, type of plumbing, and water chemistry on copper
concentrations.

In a recent Swedish study (Pettersson & Rasmussen, 1999), concentrations of copper in


unflushed drinking water were analyzed for 1,178 children living in Uppsala and Malm,
Sweden, and concentrations and amounts of copper consumed from drinking water were
estimated for 430 of these children, 9-21 months of age. The study children were from Swedish
families, were not enrolled in publicly provided day care, and were not breast-fed more than
three times a day. In the initial population, the 10th percentile for copper concentration in
unflushed drinking water was 0.17 mg/l, the median was 0.72 mg/l, and the 90th percentile was
2.11 mg/l. In the subpopulation of 430 children, the 10th percentile for daily intake of copper
from drinking water was 0.03 mg/l, the median was 0.32 mg/l, and the 90th percentile was 1.07
mg/l. The median daily intake of copper from drinking water was higher in Uppsala, at 0.46 mg,
than in Malm, at 0.26 mg. For groups of children whose families took part in a later prospective
diary study, the copper concentration in consumed water could, to some extent, be predicted
from the concentration of copper in unflushed drinking water. The lowest concentrations of
copper in drinking water were found in households with old water-pipe systems and in those
living in detached houses. A large proportion of the young children satisfied their daily
requirement of copper solely from drinking water. About 10% of the children had a copper
intake above the level recommended by the World Health Organization.

Recently, Allen and coworkers (Impellitteri et al., 2000) have shown that automatic drip coffee
makers can remove up to 85% of both copper and lead in tap water, speculating that coffee
grounds retain heavy metals through surface chelation, forming complexes with organic matter.
Sorption of the metals may also occur on interior surfaces of the coffee maker, paper filter, or

COPPER: Environmental Dynamics and Human Exposure Issues


Page 120
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

glass carafe. These findings can be important to current human exposure assessment estimates of
copper and lead in tap water as could be higher than actual levels for people whose main tap
water intake is through coffee.

Modeling Exposure to Copper in Drinking Water


A model entitled Consumption Habit Exposure Model (CHEM), was developed in Chile (Lagos
et al., 1999) for the calculation of human individual acute and chronic exposure to copper in
drinking water. The model can estimate daily exposure of individuals, as well as the peak
concentration and dose of copper which individuals ingest during a 24-hour period. Evaluation of
the model was performed through application, in a limited number of homes, of the CPS method,
used to measure chronic human consumption of contaminants from drinking water.

There are three main sources of variability in a population exposure study of copper in drinking
water: individual habits variability, chemical variability, and inter-individual variability. It was
established, mainly in evaluating the model, that the first two sources of variability are crucial
for exposure measurement. In some cases these two sources can cancel each other out, whereas,
in other cases, they can add to individual exposure estimation error. The CHEM model is not
sensitive with respect to an individuals habits variability because the WCHS questionnaire asks
for usual behavior. But the CHEM model is very sensitive to chemical variability, i.e. changes in
maximum concentration of copper measured on different days. This suggests that in order to
minimize estimation error of an individuals true exposure, measurements of the chemical
variables of the model, CMAX, CMIN (the minimum and maximum measured concentrations) and
CRAN (a random intermediate concentration), should be made more than once, and preferably
three times.

It was estimated that for people who stayed at home during the day, CMAX and CMIN weighed an
average of 3.8% each, while the rest of copper was ingested at CRAN. For those people who work
or study outside the home, the relative weights of CMAX and CMIN was 3.6 and 3.2%, respectively.

With respect to acute exposure during the winter period, it was found that 4.5% of the sampled
population was exposed to one cup of water or more at the maximum copper concentration
available at the tap. The average age of the sample segment, who arises first in the morning and
drinks stagnated water, was 41.2 years old. The average age for people who return home in the
evening and drink stagnated water, or water with maximum copper concentration, was 63.5
years. In the sample, the probabilities that the different age groups exposed to one cup or more of
water at CMAS during 1 day was 0 for the under 1-year old group: 0.4% for the 1-10-year-old
group; 0.8% for the 10-19-year-old group: 3.3% for the 20-64-year-old group, and 1.2% for the
over 64-year-old group. The results of acute exposure for the summer were similar to those
found for the winter. Finally, in the group of people who get up first, the potentially most
exposed group to high concentrations of copper, i.e. segments s.1.1 and s.1.2, were the over 64-
year-olds, with 53.4% with respect to the total population of this age group. This group is
followed by the 20-64-year-old group, with 41.6%, the 11-19-year-old group with 14.2%, and
the 1-10 year-old group with 7.1%, respectively.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 121
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

From the pespective of an essential element, it was estimated that ingestion of copper from
drinking water by the population of Santiago was on average 9.0% of the WHO recommendation
for minimum total ingestion of copper for adults, assuming that 100% of the copper contained in
drinking water is absorbed.

As mentioned earlier, more recent information on evolving methods and databases for assessing
copper exposures can be found in the forthcoming report A Framework and Data Sources for
the Assessment of Exposures to Copper (Georgopoulos et al., 2002). Details are provided on the
website of the Center for Exposure and Risk Modeling (CERM), in the section devoted to copper
(http://www.CERM.org/copper).

7.5 From Exposure to Dose: Bioavailability of Dietary Copper


The amount of copper in the diet appears to influence bioavailability more than the composition
of the diet or specific dietary components, unless the levels are extremely high or low, or diet
composition is unusual. To assess bioavailabilty and internal dose of copper, it is necessary to
understand its interactions with other components of the diet (nutrients and non-nutrients) as well
as its metabolic processes.

7.5.1 Interactions of Copper with Other Components of the Diet

Interactions with Other Nutrients


Nutrients that are known to affect copper bioavailability when included in the diet of humans or
animals in extreme amounts are iron, zinc, molybdenum, ascorbic acid, and carbohydrates. Also,
high or low levels of dietary copper may affect metabolism of some of these nutrients. The
interactions between dietary copper and other nutrients or dietary components are discussed
extensively in the literature (Turnlund, 1999; Bogden & Klevay, 2000).

The following summary is based on Turnlands study (Turnlund, 1999), where the reader can
find specific references; additional references can be found in the Toxicological Profile for
Copper (ATSDR, 1990).

Iron: Copper and iron may interact in a number of ways. Copper deficiency alters iron
metabolism. Anemia, often accompanied by accumulation of iron in the liver, has been
reported in all species studied, including humans. An excess of copper has produced
anemia in the pig. Excessive iron in the form of inorganic iron salts decreased copper
status and, in time, resulted in clinical signs of copper deficiency in several animal
species.

Zinc: When the diet contains excessive zinc over a sufficient period, the copper status of
animals and humans is impaired; the effect is reversed by copper supplements. One
explanation for this interaction is that high dietary zinc induces intestinal metallothionein
(MT). Copper does not play an important role in the induction of MT, but it has a

COPPER: Environmental Dynamics and Human Exposure Issues


Page 122
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

stronger affinity for MT than does zinc. It displaces zinc in intestinal MT and is trapped.
Copper depletion was observed in humans when supplements of 50 mg (280 mol) or
more of zinc were given for extended periods. This caused concern that dietary zinc
intake slightly above the recommended level might affect copper status, but copper
absorption was not affected by 16.5 mg (250 mol) of zinc per day. High doses of copper
have in some cases reduced the effects of zinc deficiency in animals, but these effects
were inconsistent. In one study, a high-copper diet reduced zinc absorption slightly and
increased zinc excretion in young men, but did not impair zinc status.

Molybdenum: Interactions between copper and molybdenum have been observed


frequently in ruminants. Slight excesses of molybdenum in the presence of sulfide
produce molybdenum toxicity and secondary copper deficiency. A similar response in
rats requires much more molybdenum and is independent of sulfur. A single report of a
high-molybdenum diet in humans increasing urinary copper excretion suggests that a
similar interaction could occur in humans. Molybdenum toxicity in ruminants is
ameliorated with increased dietary copper.

Ascorbic Acid: Ascorbic acid supplements have produced copper deficiency in laboratory
animals and may affect the copper status of humans. Plasma ascorbic acid concentrations
in premature infants are negatively correlated with plasma ceruloplasmin concentrations
and plasma antioxidant activity. Daily ascorbic acid supplements of 1500 mg (8.5 mmol)
given to young men caused ceruloplasmin activity to decline. Copper absorption was not
impaired by 600 mg (3.4 mmol) of ascorbic acid, but ceruloplasmin declined, suggesting
that ceruloplasmin oxidase activity may be impaired by excessive ascorbic acid.

Carbohydrates: Type of dietary carbohydrate affects copper depletion rate and severity in
rats. They are more resistant to copper deficiency when the carbohydrate is cornstarch
than when it is sucrose or fructose. The interaction between carbohydrate source and
copper in humans is not clear. Erythrocyte superoxide dismutase (SOD) levels in humans
in one study were lower with a high-fructose diet than with a high-cornstarch diet, but
copper retention increased. In addition, research in young pigs, whose cardiovascular and
gastrointestinal systems are similar to those in humans, suggests that the interaction
observed in rats may not apply to humans.

Interactions with Drugs and Nonnutrient Components of the Diet


Relatively less is known about interactions of copper with drugs. Penicillamine is used to chelate
endogenous copper in the treatment of Wilson's disease. Its use results in excretion of up to 5
mg/day (79 mmol/day) of copper in the urine. Antacids may interfere with copper absorption
when used in very high amounts. Fiber and phytate in the diet influence the bioavailability of
several minerals, but their effect on copper absorption and metabolism is not clear. Rat studies
have demonstrated both inhibition and enhancement of copper absorption by phytate. Copper use
may be influenced by both phytate and fiber in the diets of humans (Turnlund, 1999).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 123
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

7.5.2 Copper Biokinetics and Metabolism


Copper is absorbed from the gastrointestinal tract as ionic copper or bound to amino acids.
Absorption of the latter apparently involves at least two kinetically distinguishable processes.
The first mechanism transports copper from the mucosal side of the intestine to the serosal side.
Only a small fraction of ingested copper is transported via this mechanism. The second
mechanism of copper absorption involves the delivery of copper to the absorptive surface,
mucosal uptake and binding to metallothionein or another intestinal binding protein (Evans &
LeBlanc, 1976). The copper bound to metallothionein can be slowly released to the blood
(Marceau et al., 1970) or is excreted when the mucosal cell is sloughed off.

In humans, about 60% of an oral dose of 64Cu as copper acetate was absorbed from the
gastrointestinal tract; range of absorption was 15-97% (Weber et al., 1969; Strickland et al.,
1972).

Numerous factors may affect copper absorption. These factors include: (1) competition with
other metals, including zinc and cadmium (Davies & Campbell, 1977; Hall et al., 1979); (2) the
amount of copper in the stomach (Farrer & Mistilis, 1967; Strickland et al., 1972); (3) certain
dietary components; and (4) form of copper. The absorption of copper appears to be inversely
related to the amount of copper in the gastrointestinal tract (Strickland et al., 1972). High dietary
ascorbic acid has been shown to impair copper absorption (Smith & Bidlack, 1980). The
mechanism by which ascorbic acid impairs copper bioavailability is not clear, but it has been
suggested that ascorbic acid alters binding sites on metallothionein (Evans et al., 1970a).
Phytates and fiber in the diet also interfere with copper absorption, most likely by complexing
with copper to form an insoluble complex. In humans, the amount of stored copper does not
appear to influence copper absorption (Strickland et al., 1972).

Copper can pass through dermal barriers when applied with an appropriate vehicle, e.g., salicylic
acid or phenylbutazone (Beveridge et al., 1984; Walker et al., 1977).

Absorbed copper loosely binds to plasma albumin and amino acids in the portal blood and is
taken to the liver (Marceau et al., 1970). There is also evidence that copper binds to another
plasma protein, transcuprein (Weiss & Linder, 1985). In the liver, copper is incorporated into
ceruloplasmin and released into the plasma. Radioactive copper does not accumulate in
extrahepatic organs until after the emergence of ceruloplasmin-64Cu, suggesting that
ceruloplasmin is a copper donor for the tissues. However, the copper in ceruloplasmin is
probably not the sole source of copper in extrahepatic tissue. Some is probably derived from
albumin- and amino acid-bound copper.

Copper metabolism consists mainly of its transfer to and from various organic ligands, most
notably sulfhydryl and imidazole groups on amino acids and proteins. Several specific binding
proteins for copper have been identified that are important in uptake, storage, and release of
copper from tissues.

In the liver and other tissues, copper is stored bound to metallothionein and amino acids and in
association with copper-dependent enzymes. Several studies have shown that copper exposure

COPPER: Environmental Dynamics and Human Exposure Issues


Page 124
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

induces metallothionein synthesis (Mercer et al., 1981; Wake & Mercer, 1985). Increased levels
of metallothionein may be associated with resistance to copper toxicity in pigs (Mehra &
Bremner, 1984). Ceruloplasmin is synthesized in the liver. Copper is incorporated into the
molecule and released from the liver. Copper exposure has also been shown to induce
ceruloplasmin biosynthesis (Evans et al., 1970b; Haywood & Comerford, 1980).

Bile is the major copper excretion pathway. After oral administration of radioactive copper as
copper acetate in healthy humans, 72% was excreted in feces (Bush et al., 1955). Normally, 0.5-
30% of daily copper intake is excreted into the urine (Cartwright & Wintrobe, 1964). Copper in
bile is associated with low molecular weight copper binding components as well as with
macromolecular binding species. Reabsorption of biliary copper is negligible (Farrer & Mistilis,
1967).

A considerable fraction of fecal copper is of endogenous biliary origin. The remainder of fecal
copper is derived from unabsorbed copper and copper from desquamated mucosal cells.

Biliary excretion of copper following intravenous administration does not increase proportionally
with dosage, suggesting that hepatobiliary transport of copper is saturable (Gregus & Klaassen,
1986). Thus, at high copper intakes, urinary copper excretion increases.

Copper homeostasis plays an important role in the prevention of copper toxicity. Copper is
readily absorbed from the stomach and small intestine; after copper requirements are met, several
mechanisms prevent copper overload. Excess copper absorbed into gastrointestinal mucosal cells
is bound to metallothionein. This bound copper is excreted when the cell is sloughed off. Copper
that eludes the intestinal barrier can be stored in the liver or incorporated into bile and excreted
in the feces. Because of the body's efficient means of blocking excess copper absorption, the
most likely pathways for the entry of toxic amounts of copper would be long-term inhalation or
possibly through the skin, both of which would allow copper to pass unimpeded into blood.

Some copper in the diet is absorbed into the body through the intestinal mucosa, transported via
the portal blood to the liver, and incorporated into ceruloplasmin. Ceruloplasmin is released into
the blood and delivers copper to tissues throughout the body. Albumin-bound copper exchanges
with tissue copper, and a number of low-molecular-weight moieties also supply copper to the
tissues. Most endogenous copper is secreted into the gastrointestinal tract, where it combines
with unabsorbed dietary copper and is eliminated from the body. A small amount of copper is
eliminated through other excretory routes. Turnlund (Turnlund, 1999) presents the following
summary overview of the biokinetics of copper:

Absorption: Studies conducted with laboratory animals have begun to provide some basic
information on the mechanism of copper absorption. Copper is absorbed primarily in the
small intestine, with a small amount absorbed in the stomach. Absorption is probably by
a saturable, active transport mechanism at lower levels of dietary copper; at high levels of
dietary copper, passive diffusion plays a role. Absorption may be regulated by the need
for copper, with MT in intestinal cells involved in regulation. The recent development of
methods for using a stable copper isotope to investigate copper metabolism in humans

COPPER: Environmental Dynamics and Human Exposure Issues


Page 125
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

has improved copper absorption measurement reliability. Early estimates of copper


absorption in humans ranged from 15% to 97%. Estimates of copper absorption varied
widely in part because of inadequate methods and because dietary copper level was not
known or controlled. A series of stable isotope studies of copper absorption have since
demonstrated dietary copper level strongly influences absorption. As dietary copper
increases, the fraction absorbed declines and the amount absorbed increases. Absorption
declined from 75% at 0.4 mg (6 mol) copper per day to 12% at 7.5 mg (120 pimol)
copper per day. The amount absorbed increased from 0.3 to 0.9 mg (5 to 15 mol), or
only tripled with twenty times the amount of dietary copper. This suggests that copper
homeostasis is maintained in part by regulation of absorption.

Storage: The adult human body contains only about 50 to 120 mg (0.79-1.9 mmol) of
copper, very little compared with other trace elements, such as iron and zinc. Animal data
suggest that copper is stored in the liver, bound to MT-like proteins. Ruminants and a few
other animal species can store much more copper in the liver than can humans or most
other animal species. The copper content of a 70 kg adult human has been reported to be
72 mg, but more recent results of studies of copper in tissue suggest that it is in the range
50-70 mg (Davies & Bennett, 1985). Wise and Zeisler reported an average copper
concentration of 10 ppm in the human liver in 36 samples. Despite the wide variation in
copper concentrations in the environment, the copper concentration in the liver only
varied by a factor of 2-3.5. Copper may also be held, at least temporarily, bound to
intestinal MT.

Transport: Following absorption, copper is transported bound primarilv to albumin and


to transcuprein and low-molecularweight ligands. The newly absorbed copper disappears
rapidly from the plasma. Most is taken up by the liver; some is taken up by the kidney.
Once in the liver, copper is incorporated into ceruloplasmin within hours. Some is
incorporated into MT in the liver of animals, particularly when copper intake is high; a
role for MT in cellular detoxification has been proposed. A role for copper in the kidney
is not known, but copper is probably filtered by the glomerulus and reabsorbed in the
tubules. Copper bound to ceruloplasmin is released from the liver into the blood and
delivered to cells with specific ceruloplasmin receptors on their surface. Ceruloplasmin
binds to these receptors; the copper is reduced, dissociates from ceruloplasmin, and is
released into the cells. One suggested sequence leading to excretion of copper in bile is
that ceruloplasmin with part of the copper removed may return to the liver, where it is
partially degraded and transferred to the bile accompanied by ceruloplasmin fragments.
Glutathione, although not a cuproenzyme, may play a role in the rapid transfer of
excessive copper to bile.

Excretion: The primary route of copper excretion is via bile into the gastrointestinal tract.
Little of this copper is reabsorbed. It combines with a small amount of copper from
intestinal cells, from pancreatic and intestinal fluids, and unabsorbed dietary copper and
is then eliminated in the feces. Other routes of excretion contribute little to total copper
losses. Healthy humans excrete only 10 to 30 g (0.2-0.5 mol) of copper in the urine,

COPPER: Environmental Dynamics and Human Exposure Issues


Page 126
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

but urinary losses can increase markedly in some conditions, such as renal tubular
defects. Sweat and integumentary losses are usually less than 50 g (0.8 mmol) per day.

7.5.3 A Note on Unusually Susceptible Populations (Based on ATSDR, 1990)


In order to understand and quantify health risks of copper deficiency or excess in the general
population it is also necessary to identify those subpopulation groups that are unusually
susceptible to copper. The following discussion is based on ATSDRs 1990 Toxicological
Profile for Copper:

Individuals with Wilson's disease, a genetic disorder, are unusually susceptible to copper
toxicity because of impaired ability to maintain normal copper homeostasis. Limiting
copper intake through air, water, and food (Schroeder et al., 1966), as well as by special
medical treatment, is essential in treating Wilson's disease.

Infants and children under one year old are unusually susceptible to the toxicity of
copper. Muller-Hocker et al. (Muller-Hocker et al., 1988) reported pronounced
hepatosplenomegaly in two infant siblings exposed to high levels of copper in tap water.
No effects were observed in an older sibling or the parents. Infants and children under
one year old are in a high risk category because they have not yet developed the
homeostatic mechanisms for clearing copper from the body and preventing its entry via
the intestine.

Because of the liver's key role in copper storage, ceruloplasmin synthesis and copper
excretion into bile, metabolic and pathologic dysfunction would probably disrupt copper
homeostasis. Thus, individuals with liver damage may be unusually susceptible to copper
toxicity.

At high copper intakes, urinary copper excretion increases. Individuals with chronic renal
disease may have difficulty in handling high copper intake.

Acute hemolytic anemia has been reported in several dialysis patients exposed to excess
copper in the dialysate (Williams, 1982). Although this population is not unusually
susceptible to copper toxicity, it is at high risk of being exposed to high levels of copper.
Close monitoring of the pH and conductivity of the dialysis fluid will help control high
exposure to copper (Williams, 1982).

Individuals with an inherited deficiency of the enzyme glucose-6-phosphate


dehydrogenase are likely to be susceptible to toxic effects of oxidative stressors such as
copper (Calabrese & Moore, 1979). In humans with this enzyme deficiency, the threshold
for copper exposure may be lower than in those without the condition (Chugh & Sakhuja,
1979).

COPPER: Environmental Dynamics and Human Exposure Issues


Page 127
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

7.6 Modeling Health Risks Due to Copper Exposures


The need to develop specific approaches for risk assessment for exposure to essential elements
has been recognized recently. Traditional methods for health-based risk assessment for exposure
to non-essential elements first determine intake/exposure levels at which no observed adverse
effects (NOAEL) are found, or the lowest level at which adverse effects are observed (LOAEL).
In addition, uncertainty factors (UF) or modifying factors (MF) are used to adjust the NOAEL or
LOAEL and define reference doses (RfD), chronic intake/exposure levels considered safe or of
no significant health risk for humans (Barries & Dourson, 1988; Dourson, 1994). Uncertainty
factors are used in deriving RfDs from experimental toxicity data obtained from animal studies
when data from humans is insufficient to fully account for population variability or special
sensitivity subgroups of the general population, when NOAEL was obtained from studies of
insufficient duration to assure chronic safety, or when the database on which the NOAEL is
based is incomplete; or when experimental data provides a LOAEL, instead of a true NOAEL
(Dourson, 1994; Dourson & Stara, 1983). The usual value for each UF is a ten-fold reduction in
the acceptable exposure level for each of the considerations listed above and may be used alone
or in combination, depending on the specific element being assessed. Modifying factors are
additional uncertainty factors which have a value of 1 or more, but less than 10, and are based on
professional judgment of the overall quality of the data available (Dourson, 1994). Given the
limited human data available, the limitations of animal models, and the uncertainties of
interpretation, the traditional toxicological approach, summarized here, for defining exposure
limits for essential elements, may, in fact, lead to the establishment of limits which, if followed,
promote or even induce deficiency.

A new approach for health risk assessment of essential elements was proposed by Olivares et al.
(Olivares et al., 1999) to take into account risks associated with low intakes, as well as with high
intakes. The relationship between exposure level, intake, and risk for essential elements has a U
or J shape, not perfectly symmetrical, but including risk of deficiency associated to low intakes
and risk of toxicity associated to high intakes.

Following this approach, the steps proposed in applying the new risk assessment model are:

Step 1. Review and Evaluation of Data from Human and Relevant Animal Studies
First it is necessary to review experimental studies in which several intake/exposure levels have
been evaluated in order to estimate requirements and toxic effects. Since the aim is to develop
population guidelines, the estimate should include a measure of variability in order to define the
distribution of essential element requirements and toxicity for an ideal normal population. This
would be defined as acceptable range of exposure/intake (AREI) for a specific element. The
review should include the full range of biological effects, beginning with the most extreme, such
as death, and ending with the most subtle, such as taste.

Death and disease related to acute or chronic essential element exposure/intake can, in fact, occur
and should be critically assessed. Reported deaths from deficit or excess should be studied,
because in some situations, the biological effect of essential element deficit or excess may be due

COPPER: Environmental Dynamics and Human Exposure Issues


Page 128
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

to other intervening factors, such as disease or genetic predisposition. In these cases, it is


particularly difficult to establish a causal relationship, without controlled studies. Eliminating the
element (in the case of toxic effects) or incorporating it (in the case of deficit) in a group of
subjects randomized to control or experimental conditions is necessary to test a causal
relationship hypothesis. This type of data is largely unavailable for chronic exposure to, or acute
toxicity (e.g., suicide via elements such as copper or iron) from essential elements. In the case of
animals, LDs for various percentages have been defined for some elements, but are unavailable
for most. Human data on lethality from deficit or toxicity are limited in that they are not
population-based, but rather are usually linked to single case reports or at best to a cluster of
cases due to unusual circumstances. This is because today, human populations are protected from
death from deficient or toxic exposure/intake levels.

For most essential elements, disease conditions associated to deficit have been well described
(Fe, Cu, Zn, Mn, Se); much less is known about pathologic conditions linked to excess; few have
been well characterized. In both cases, deficit and excess, disease has sometimes been linked to
genetic conditions favoring the occurrence of deficiency or toxicity by modifying requirements,
decreasing or increasing absorption, or altering excretion. In many pathologic conditions specific
data on chronic levels of exposure/intake leading to disease from excess or deficit may be
lacking, but information on the actions required to cure the condition may be indicative of the
safe level of exposure/intake. For example, we know the level of zinc intake required to prevent
growth failure in infants (National Research Council, 1989a), or the level of iron excess which
may induce hemolysis in susceptible individuals (Williams et al., 1975). On the other hand, in
many cases, a metals pathological effect may occur only with a concurrent factor, such as viral
infection, in the case of myocardial degenerative effects induced by selenium deficiency (Keshan
disease) (Cousins, 1996), or copper toxicity in patients undergoing hemodialysis due to chronic
renal failure (Bloomfield et al., 1971). The data from exposure/intake levels to prevent disease
states, if available, represents the best option to define AREI since there is no question that
preventing disease is relevant to the protection of human health.

The next data evaluating level in AREI estimation is assessment of studies evaluating
biologically significant effects of various exposure/intake levels. The difficulty here is
establishing what will be considered biologically significant and what will not. A suitable
definition for biological significance in this context is the capacity of the biological indicator
modified by E/I to predict the occurrence of deficiency or toxicity disease associated with a
particular element. For example, elevated ferritin may serve to predict liver damage induced by
elevated iron intake in susceptible individuals (Beaumont et al., 1979). A functional assay, such
as red cell resistance to peroxide stress may also predict the risk of hemolysis from excess iron
(Farrell et al., 1977). On the other hand, elevation of superoxide dismutase in this same condition
may represent an adaptive response to oxidant stress without pathological implication (Nielsen &
Milne, 1993). Unfortunately most biochemical or functional biomarkers have not been validated
in terms of their ability to anticipate the occurrence of disease. In this context, the most valid are
those that relate to the limits of AREI, for example levels of E/I that indicate excessive or
deficient retention of a specific element. For example, negative Zn balance over time will lead to
a disease state; on the other hand, markedly positive copper balance may serve to indicate a level
of E/I, which, over time, may lead to toxicity. For the purpose of quantifying AREI, a key

COPPER: Environmental Dynamics and Human Exposure Issues


Page 129
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

biomarker may be the change in bioavailability induced by high or low E/I, since this may be the
most sensitive indicator of excess or deficit. For example, high copper E/I is associated with
lower bioavailability, while low copper E/I leads to higher copper bioavailability, as measured
by labeled copper studies, using stable isotopes, which permit the separation of exogenous E/I
from endogenous excreted copper (Turnlund et al., 1989).

Following these concepts, one could develop quantitative estimates of AREI on the basis of
lethal effect, prevention of disease, or assuring a normal range of biomarkers with biological
significance. The most stringent criterion will undoubtedly be the last. In terms of feasibility,
prevention of disease and normalcy of biomarkers are most likely to be used.

Step 2. Assessment of Biokinetics and Interactions


It is important, when defining AREI, to consider the biological processing of the element, which
may affect biological impact in terms of deficit or excess. This includes absorption, transport,
metabolism, excretion and storage of the element and its possible interaction with other elements
or factors which modify them. Consideration of these factors in adjusting AREI levels within a
given context is important since they may modify biological effect significantly. Factors which
affect biokinetics differ for each element, but basically include element interaction, interactions
with other dietary factors, and the effect of environmental agents. For example, excess zinc
interferes with copper absorption (van Campen & Scaifi, 1967); thus, the higher limit of copper
AREI will be greater if there is concomitant exposure to a high level of zinc. Ascorbic acid will
enhance iron absorption (Stekel et al., 1985), thus the lower limit for iron E/I may be smaller, if
diets are high in ascorbic acid. In contrast, high fiber intake lowers iron absorption (Simpson et
al., 1981). Trace metals are not metabolized, thus excess E/I, if not adequately excreted, will
cause accumulation and, once storage capacity is saturated, toxicity. The capacity to adapt to
high E/I for most trace essential metals is dependent on absorption regulation. The capacity to
store trace elements is present for some metals, for example, iron and copper, but not for all, i.e.
zinc. Thus, low-intake induced zinc deficiency may be apparent over a shorter time frame than of
iron or copper. Excretion of most essential trace elements is by the gastrointestinal route, but
homeostasis regulation is not effective since it is easily saturated leading to accumulation and
possible toxicity. There is no efficient way to excrete iron except by blood loss (Bothwell et al.,
1979), while copper is mainly excreted by the biliary route (Linder & Hazegh-Azam, 1996).
Thus excess E/I, once excretion and storage capacity are saturated, may lead to toxicity. Because
storage of trace elements is mainly in the liver, it is potentially subject to toxic manifestations, as
is the case for iron and copper.

Step 3. Procedures for Defining Critical Effects of Both Copper Deficit and Excess
Another key aspect in defining an AREI is the assessment of critical effects of deficiency and
toxicity relevant to human health; the most sensitive indices of excess or deficiency may be
biomarkers without clear functional or health significance. At the other extreme, death associated
with organ damage induced by excess or deficit has clear health significance but is not relevant
as a sensitive indicator of health risk.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 130
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

We will use copper to exemplify the assessment of health risks associated with an element.
Biochemical changes such as red cell SOD activity, as an index of copper deficit (Uauy et al.,
1985), or changes in plasma copper/ceruloplasmin molar ratio, as an index of copper excess are
sensitive but nonsignificant indicators of health risks (Frommer, 1976). Biological indices of
subclinical effects on specific function indicating potential adverse effects of copper deficit or
excess are often used in human studies (Heresi et al., 1985). Examples are decreased white blood
cell phagocytic capacity, in the case of deficit, or increased serum aminotransferase or
transpepticlase hepatic enzymes, in response to excess copper. Clinical effects, such as bone
fractures, in the case of copper deficiency, or liver dysfunction due to fibrosis, in the case of
excess, are clearly significant in terms of health risk (Shaw, 1992; Sternlieb, 1982), but are
difficult to assess in controlled studies since ethical principles in human investigation preclude
precise quantitative definition of these endpoints. Death from bacterial infection associated with
neutropenia, in the case of deficit, or liver failure associated to excess, are clearly not applicable
endpoints to define ranges of AREI, though they are of unquestionable significance.

From the principle of utilizing comparable effects to define excess and deficit, the concept arises
that effects of similar health significance should be selected to define upper and lower ranges of
AREI. In general, the range should be defined by effect response to exposure/intake levels that
prevent subclinical adverse effects. Review of these indices and corresponding studies should be
done by toxicologists and nutritionists familiar with health risk assessment. The combined effort
should yield clearly-defined critical endpoints for upper and lower ranges of AREI, given
available information.

Step 4. Quantitative Evaluation of Critical Effects to Define AREI


Data to support the lower AREI limit come from mineral balance studies and/or clinical studies
indicating amounts needed to prevent signs of deficiency. Data relevant to humans pertaining to
excess are quite limited. The upper end of the exposure range is usually derived from limited
studies of individuals taking mineral supplements, accidental overdose or suicidal attempts, or
exposure in the workplace, or in the environment. It is virtually impossible to find population-
based evidence of toxic effects in adult humans consuming regular food and potable water, since
all public health measures are directed to prevent this. Then, there is, potentially, the rare
genetically-susceptible group that may manifest toxicity at exposure/intake levels observed in
normal diets (Fleet & Mayer, 1996).

Another approach to estimating AREI in quantitative terms is to examine the range of


exposure/intake of healthy populations, assuming that if the population is in good health, the
exposure must be safe. This is particularly helpful if available experimental human information
is limited. A starting point in the definition of AREI can be the customary exposure/intake
observed in healthy populations. If the upper or lower AREI cutoffs are obtained via
extrapolation from animal studies, the customary exposure/intake of healthy populations
should serve to validate the extrapolation. It would be clearly unwise to define AREI outside the
range of customary intakes of healthy populations. Infants are more sensitive to copper deficit
(by a factor of 2.5) and to copper excess (by a factor of 0.8) relative to adults. That is, they need

COPPER: Environmental Dynamics and Human Exposure Issues


Page 131
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

2.5 times as much copper as adults per unit of body weight to prevent deficit and they tolerate
0.8 times less copper than adults per unit of body weight. The human data do not provide firm
data to define a NOAEL, but based on WHO/FAO/AREI expert group recommendations, the no
observed effect level for copper is 200 g per kg body weight. A special case is that of premature
infants, as their need is extremely high (7.5 times greater than that of adults per unit of body
weight), while we know little about their tolerance limit, if the value for adults is considered, the
acceptable range would be particularly narrow.

Olivares et al. (Olivares et al., 1999) demonstrated this approach using data on intake from IAEA
information collected in several developed countries. They demonstrated that there is a
significant proportion of the population, nearly 20%, that may be receiving too little copper,
while excess copper is not a problem, unless food or water is contaminated with copper. The
AREI for infants compared to usual dietary intake revealed a similar problem; a significant
proportion of the infant population is at risk for deficit, while most preterm infants will develop
clinical deficiency unless given formula supplemented with extra copper to concentrations of 1.2
to 2.0 mg/l (Shaw, 1992). Infants are more sensitive to copper deficit and excess than adults per
unit of body weight. Actual intake data demonstrate that, on a population basis, the risk of
deficiency is much greater than excess. The Environmental Health Criteria for Copper meeting
held in Brisbane in 1996 recognized this in establishing its conclusions (Becking, 1996). In fact,
except for contaminated food and drink or genetic defects in copper elimination, toxicity is
extremely uncommon in humans.

Step 5. Qualitative and Quantitative Assessment of Critical Effects to Define AREI for Normal
Populations
A tentative exposure/intake level for preventing both toxicity and deficiency in the general
population can be derived from the data reviewed and adjusted under steps 1-3. The upper and
lower cutoff points for the range of acceptable oral intakes should be defined for population
groups. The lower cutoff point should be sufficient to meet the requirements of most individuals
in the population. This is usually called the recommended dietary allowance or intake (RDA or
RDI). Based on criteria used to define RDAs or RDIs most usually implies 97.5% of the
population: if the mean and standard deviation for requirements are known, this point is defined
by the mean + 2 SD; if the SD is not known, a CV of 15% is customarily used to account for
population requirement variability. Similarly, the upper end point should protect most
individuals from toxicity risk. A statistical definition for most in this context is lacking, but
should be derived based on the mean and distribution of toxic effect dose, using the traditional
toxicology approach. Based on known variability or extrapolation from dose-effect response, an
ED should be defined. Special consideration of upper and lower cutoff points should be made in
defining AREI for physiologic conditions affecting normal populations, such as infancy,
pregnancy, lactation, aging, exercise, and changes in climatic conditions, when pertinent. Lower
cut off points of AREI are usually considered in setting RDAs, but for upper cut off points,
specific values for these categories should be given, if the values are sufficiently different. The
acceptable range of exposure/intake AREI, that is E/I levels sufficient to meet the requirement,
as well as to prevent toxicity risk for most individuals in the population, may be extremely

COPPER: Environmental Dynamics and Human Exposure Issues


Page 132
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

narrow, if traditional uncertainty factors are used in defining the upper value of the acceptable
range. The need to achieve a balanced and comparable approach to assess risk from both deficit
and excess is needed when evaluating essential elements (Bowman & Rishert, 1994).

In the case of copper, the defined AREI needs to meet criteria for both chronic and acute
exposure safety. This is not an issue for food, since copper in food will likely never induce acute
adverse effects. In contrast, soluble copper salts will induce acute vomiting if taken in sufficient
amounts; ionic copper is a powerful emetic agent, as previously indicated. Most values for
drinking water are below 1 mg/l and very few over 2 mg/l. At this level no acute manifestations
of intolerance are found, except for bitter taste in sensitive individuals. Above 5 mg/l an
increasing proportion of subjects evidence nausea and abdominal pain.

In summary, if one were to apply an upper limit of 0.5 mg/kg per day, accepting that the
contribution from diet is usually 0.05, or at most, 0.1 mg per kg per day, the safety margin for
copper intake from water is sufficient to support the present provisional guideline value of 2
mg/l. This analysis considers neither coppers limited bioavailability, nor adaptation to varying
levels of exposure/intake. If these were included, the chronic safety margin would be even
greater.

Step 6. Adjustment of AREI for Special Groups of Public Health Significance


Finally, another principle guiding the definition of an AREI is that the acceptable oral exposure
intake range should be safe for the general population and not be expected to meet the
requirements of, or prevent excess for, special groups. For example, the zinc needs of patients
with chronic diarrhea or of patients on chronic hemodialysis may fall outside the AREI. In both
these cases, intakes risk for the healthy population may be required to meet the special needs
of these patients. Safe zinc exposure/intake levels for hemodialysis patients may lead to
deficiency in healthy subjects, while intakes required by patients with chronic diarrhea may be
excessive for normal subjects. In the case of essential elements it is clearly impossible to assure
that 100% of subjects will be protected from deficit or excess. The range of AREI defined at the
international level is not intended to address disease conditions or genetically-caused alterations
in trace element metabolism determining special sensitivity to excess or deficit. These conditions
should be addressed by national or regional regulating agencies based on the public health
relevance of these special conditions. The AREI is clearly not intended to meet the special needs
of population subgroups with genetic alterations of trace metal metabolism, for example
achrodermatitis enterophatica, a disease of zinc metabolism in which absorption is extremely low
(Moynahan, 1974). In fact, because of an intestinal transport defect, AE patients need two to
three times the usual zinc intake, an E/I level which may interfere with copper absorption and
increase the risk of copper deficit in normal subjects. Similarly, Wilsons disease patients will
demonstrate marginal benefits from eliminating copper from food and water, which may trigger
copper deficiency in most healthy subjects.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 133
Chapter 7
ASSESSING HUMAN EXPOSURE, DOSE AND RISK

COPPER: Environmental Dynamics and Human Exposure Issues


Page 134
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

8 BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

8.1 Biochemical Indices of Copper Status

8.1.1 Biomarkers of Exposure Susceptibility and Effect


Biomarkers are defined as indicators of events in biologic systems or samples. They are
classified as markers of exposure, markers of effect, and markers of susceptibility (National
Research Council, 1989b).

A biomarker of exposure is a xenobiotic substance or its metabolite(s) or the product of


an interaction between a xenobiotic agent and some target molecule or cell that is
measured within a compartment of an organism (National Research Council, 1989a). The
preferred biomarkers of exposure are generally the substance itself or substance-specific
metabolites in readily obtainable body fluid or excreta. However, several factors can
confound the use and interpretation of biomarkers of exposure. The body burden of a
substance may be the result of exposures from more than one source. For example, high
urinary levels of phenol can be the result of exposure to several different aromatic
compounds. The substance being measured may be a metabolite of another xenobiotic.
Depending on the properties of the substance (e.g., biologic half-life) and environmental
conditions (e.g., duration and route of exposure), the substance and all of its metabolites
may have left the body by the time biologic samples can be taken. It may be difficult to
identify individuals exposed to hazardous substances commonly found in body tissues
and fluids, such as the essential mineral nutrients, copper, zinc and selenium.

Biomarkers of effect are defined as any measurable biochemical, physiologic, or other


alteration within an organism that, depending on magnitude, can be recognized as an
established or potential health impairment or disease. This definition encompasses
biochemical or cellular signals of tissue dysfunction, such as increased liver enzyme
activity or pathologic changes in female genital epithelial cells, as well as physiologic
signs of dysfunction, such as increased blood pressure or decreased lung capacity. Note
that these markers are often not substance specific. They also may not be directly
adverse, but can indicate potential health impairment (e.g., DNA adducts).

A biomarker of susceptibility is an indicator of an inherent or acquired limitation of an


organism's ability to respond to the challenge of exposure to a specific xenobiotic. It can
be an intrinsic genetic or other characteristic, or a pre-existing disease that results in an
increase in absorbed dose, biologically effective dose, or target tissue response.

Currently used indices of copper status easily be employed to detect severe copper deficiency.
Serum copper and ceruloplasmin concentrations fall to levels far below the normal range and
respond quickly to copper supplementation. Ceruloplasmin concentration has generally been
considered the most reliable index of copper status, but some consider red cell superoxide
dismutase activity equally or more sensitive. SOD values have not yet been reported in cases of
severe copper deficiency in humans, and a level that would indicate copper deficiency has not

COPPER: Environmental Dynamics and Human Exposure Issues


Page 135
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

been established. The normal ranges of these indices vary between laboratories but are
approximately as follows: 10.0 to 24.6 mol/l (64-156 Ag/dl) for serum copper; 180 to 400 mg/l
(18-40 mg/dL) for ceruloplasmin; and 0.47 0.067 mg/g for erythrocyte SOD.

Although serum copper and ceruloplasmin levels clearly reflect severe deficiency, they may not
be sensitive enough to identify marginal copper status. In addition, ceruloplasmin is an
acute-phase reactant and, as a result, serum copper and ceruloplasmin levels are elevated in a
variety of conditions. Serum copper and ceruloplasmin levels could be within the normal range
or even elevated, masking copper deficiency when one of these conditions occurs at the same
time.

The search for a reliable index of marginal copper status has been unsuccessful, though a number
of possibilities have been suggested. Copper levels in hair, nails, or saliva do not appear to
reliably reflect copper status. Urinary copper varies greatly between individuals and declines
only when dietary copper is very low. Cytochrome C oxidase in red cells, or possibly in platelets
or white cells, may be sensitive to copper status, but more data are needed. Other indices may
potentially provide infor and changes in lysyl oxidase activity.*

Cohen (Cohen, 1979) attempted to establish guidelines for evaluating the patient or worker with
possible chronic exposure to copper. The importance of the history and the physical examination
is emphasized as the initial approach to diagnosis. The patient may bring a suggestive pattern of
complaints to the attention of medical personnel, such as nasal irritation, metallic taste, and
episodes of fever and shaking chills, to name some of the usual symptoms. Of course these
symptoms invariably relate to exposure in an occupational setting, and the most important step in
the detection of a potential occupational hazard is to visit the workplace. It is a usual and
expected industrial hygiene practice to measure ambient levels of suspected toxins; in this case
the concentration of copper fumes, dust, or salt mist, among others, would be determined.

If the source of the exposure is not clear but copper intoxication is suspected, the analysis of
24-hr urinary copper levels provides a simple and reliable screening procedure. However, care
must be taken that a specimen is collected and tested in a copper-free system. To rule out
exogenous urinary contamination, a serum copper level should be performed in all cases where
the urine level is increased.

*
Regarding the issue of characterizing copper status, of particular interest is a workshop initiated by the Fogarty International
Center and the University of Chile, Genetic and Environmental Determinants of Copper Metabolism, on March 18-20, 1996. This
workshop was also cosponsored by the National Institute of Child Health and Human Development, the National Institute of
Diabetes and Digestive and Kidney Diseases, the National Institute of Environmental Health Sciences, and the Environmental
Protection Agency. The workshop brought together over sixty international scientists from academia, industry and governments
who work in the area of copper metabolism, requirements, or toxicity. This group reviewed, from an international perspective,
recent significant advances in understanding of the molecular mechanisms underlying genetic disorders of copper metabolism
and their implications for copper intake and exposure. One session of this workshop focused on the difficult question of how to
adequately measure copper status in humans. The proceedings of this workshop were published as a supplement to the American
Journal of Clinical Nutrition. It should be mentioned that this workshop further identified gaps in basic research related to copper
metabolism and pointed to specific areas where research is needed relative to copper requirements, copper-nutrient interactions,
and specific disease endpoints with regard to dietary copper supplementation.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 136
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

The diagnosis of an allergic contact dermatitis, when copper is the suspected agent, may be
confirmed via dermal patch test. The patch is impregnated with a 5% solution of copper sulfate
(Fisher, 1973). A positive reaction includes the presence of erythema and vesiculation at the test
site. It should be noted, however, that the one case of allergic contact dermatitis to copper
reported by Saltzer and Wilson (Saltzer & Wilson, 1968) resulted in a positive patch test with a
copper sulfate solution as low as 1.25%.

Biomarkers Used to Identify or Quantify Exposure to Copper.


Copper levels can be readily measured in tissues, body fluids, and excreta. Following copper
inhalation and ingestion, increased levels of serum, urine, hair, and hepatic copper have been
reported in humans and animals.

Increased whole blood and serum copper levels have been reported in humans following
ingestion of 1-30 g of copper. Serum and whole blood levels ranging from 239-346 and 383-684
mg/100 mL, respectively, were observed. Serum and whole blood levels of 151.6 and 217
mg/100 mL, respectively, were reported in controls (Chuttani et al., 1965). Plasma serum levels
of greater than 200 mg/100 mL were observed in 16% of factory workers exposed to copper dust
(111-464 Cu/m3) (Suciu et al., 1981). Increased serum copper levels may be only reflective of
recent exposure. It was observed (Chuttani et al., 1965) that serum ionic copper rapidly
diminishes to normal levels following an acute bolus dose.

Increased hair copper levels (705.7 mg/g, controls 8.9 mg/g) have also been reported in workers
exposed to airborne copper (0.64-1.05 mg/m3) (ATSDR, 1990).

Biomarkers Used to Characterize Health Effects of Copper


Harmful health effects of copper occur over a wide range of copper intakes. This section contains
a review of health effects associated with low dietary copper levels and a discussion of the
association of high environmental copper levels with health effects and plasma copper levels.

Low Intakes of Copper. Nutritional copper requirements and health effects associated with
copper deficiency have been reviewed by numerous authors (Mason, 1979). This section is a
summary of these reviews.

Copper deficiency is rarely observed in humans; in fact the existence of covert copper deficiency
among segments of the population is unknown. The limited data available on the human health
effects of low copper intake are derived mostly from case reports of severely malnourished
children, patients maintained by total parameter nutrition without copper, and children with
Menkes' disease. Copper deficiency is characterized by hypochromic anemia, connective tissue
abnormalities, and central nervous system (CNS) disorders. Sudden death associated with
spontaneous rupture of a major blood vessel or of the heart itself has been observed in some
animal species.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 137
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

Copper deficiency manifestations are related to a decrease in several of the copper-containing


metalloenzymes. The most severe biochemical alteration is a decrease in cytochrome oxidase
activity, manifested by poor growth, anemia, and CNS effects. Decreased oxidative metabolism
associated with decreased cytochrome oxidase results in poor growth in infants, weight loss, and
emaciation. Hypochromatic anemia observed during copper deficiency is not distinguishable
from iron deficiency anemia, but is not responsive to iron administration. The exact mechanism
involved in the development of the anemia is not known, but copper is thought to have a role in
iron transportation and utilization. A decrease in protoheme synthesis, a result of decreased
cytochrome oxidase, has also been observed. As with anemia, the CNS effects, primarily the
result of hypomyelination, are associated with low activity levels of cytochrome oxidase.
However, the decreased synthesis of phospholipids observed in copper deficiency may also
contribute to the development of CNS effects. In addition to the decrease in cytochrome oxidase,
a decrease in lysyl oxidase is also observed. Lysyl oxidase is involved in the formation of cross-
links in collagen and elastin. Depending on the species, this impairment results in bone disorders,
a defective cardiovascular system, or abnormal lung structure.

Menkes steely hair syndrome is a severe congenital copper deficiency. It is characterized by


slow growth, progressive cerebral degeneration, convulsions, temperature instability, scorbutic-
like bone changes, and peculiar steel-like hair (Burhanoglu et al., 1996).

Exposure to Excess Levels of Copper. A 1965 study (Chuttani et al., 1965) attempted to correlate
levels of serum and whole blood copper with symptom severity in adults following acute copper
sulfate poisoning. Though the statistical analysis of the correlations between symptoms (i.e.
gastrointestinal effects, jaundice, renal manifestations) and blood copper levels was unclear, a
significant correlation between whole blood copper and symptom severity was found.
Gastrointestinal effects occurred in patients with whole blood copper levels of 287 mg/100 ml.
Whole blood copper levels of 798 mg/100 ml were observed in patients with jaundice, renal
manifestations, or shock, in addition to gastrointestinal symptoms. Single exposure to 30 mg/l
copper or greater in drinking water resulted in gastrointestinal effects (vomiting, diarrhea, and
abdominal pain) in healthy humans (Holleran, 1981; Semple et al., 1960; Wyllie, 1957).
Vomiting and abdominal pain have also been reported in individuals who consumed water
containing 7.8 mg/l of copper for >1 year (Spitalny et al., 1984). Consumption of water
contaminated with high levels of copper (2.2-3.4 mg/l) resulted in severe liver damage in two
infant siblings (Muller-Hocker et al., 1988).

8.2 Copper Biomarkers Used in Recent Human Health Studies


A study was done to determine whether kidney dialysis patients were placed at risk by the
dialysis procedure, for either copper deficiency or copper toxicity. The techniques used involved
studying ceruloplasmin activity, plasma copper, and erythrocyte superoxide dismutase (SOD) in
a single group of dialysis patients (Emenaker et al., 1996). Previous studies had not considered
all three indexes in the same group of dialysis patients. Also unique to this study was the
measurement of the ratio of ceruloplasmin activity to protein and mononuclear cell copper

COPPER: Environmental Dynamics and Human Exposure Issues


Page 138
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

contents. All of the patients in the study had undergone hemodialysis with membranes that were
not copper based.

Examination of the ceruloplasmin related indexes in the dialysis patients found moderate copper
deficiency, but erythrocyte SOD and mononuclear cell copper measurements revealed no such
deficiency. No definitive conclusion could be reached regarding dietary requirements for kidney
dialysis patients.

The role of copper in either causation of, or association with, various pathological states has been
examined superficially in several studies. In Turkey (Dincer & Akar, 1995), metal levels in
maternal hair were examined to see if there was an association between the various metal levels
and neural tube defects in their newborn infants. The study was primarily conducted to examine
zinc deficiency and its correlation to birth defects. Since interactions between copper and zinc
are well documented, copper and magnesium levels in maternal hair were also measured. No
association was found between copper levels in maternal hair and neural tube defects in the
infants.

Another study conducted in Turkey (Onder et al., 1996) in 1995 examined the levels of copper,
zinc, and magnesium in rheumatoid arthritis patients. This study was different from most others
reported in the literature because it included both plasma levels of the metals and intracellular
levels of the metals in both leucocytes and erythrocytes. Two groups of patients were studied;
one group was treated with Tenoxicam and the second was untreated. A control group consisted
of healthy individuals with no evidence of arthritis.

Both of the rheumatoid arthritis patient groups had higher levels of copper in plasma, leucocytes,
and erythrocytes than did the healthy control group. Zinc levels in both patient groups were
lower than in the control group. There seemed to be no effect from the treatment drug
Tenoxicam on copper or zinc levels in the arthritis patients.

An association between high serum copper levels and increased risk for cardiovascular death was
illustrated by a study done in Finland in 1996 (Reunanen et al., 1996). The authors of the study
emphasized that there is no proven causal effect between serum copper and cardiovascular
disease, rather that further studies need to be done to examine the reasons for the variation in
serum copper levels. In fact, animal studies supply data directly contradicting this human study;
namely, low dietary copper and high zinc are associated with hypercholesterolemia and
cardiovascular complications.

In the human study, other risk factors were controlled in an attempt to determine the independent
prognostic value of copper, calcium, magnesium and zinc serum levels on cardiovascular death.
The authors felt it unlikely that the high copper / low zinc serum levels were due to dietary
imbalance. A possible explanation is that the inflammatory response will affect the levels of
various body elements. Elevated copper levels may be due to the inflammatory response set in
motion by some other independent pathological process. Another factor that needs to be
considered is that the serum levels of an element like copper do not necessarily reflect
intracellular levels.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 139
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

The role of copper and other metals in the development of malignant and pre-malignant
conditions has been an area of interest for some time. Oral Submucous Fibrosis (OSF) is a pre-
malignant chronic disease found in many individuals in India and in other parts of Asia. The
disease affects the epithelial lining of the oral cavity. A study was done (Anuradha & Devi,
1995) to examine the role of trace elements, including copper, present in the tissues and blood of
patients with OSF, in the occurrence and development of the disease. There was a significant
decrease in both ceruloplasmin and copper in all patients studied. It was suggested that the
decrease in the patients copper levels subsequently caused a decrease in ceruloplasmin levels.
The significance of the copper decrease is illustrated by the role that copper plays in the etiology
of the disease. OSF is a collagen disorder. In order to synthesize the excess collagen that is
characteristic of the disorder, lysyl oxidase (a copper containing enzyme) must catalyze the
formation of desmosine (the cross linkage group in the collagen macromolecule). Extensive
enzyme activation and utilization would theoretically deplete copper levels in these patients.
Disease progression was directly linked to copper depletion.

Trace metal levels, including copper, were examined in both healthy tissue and malignant tissue
from the same patients in a study done in Pakistan (Tariq et al., 1995). In malignant tissue,
copper levels were lower than in healthy tissue from the same organ. Overall, female cancer
patients had lower blood copper levels than healthy controls. Male cancer patients had higher
blood copper levels than did healthy controls. No clearly definable overall trend in trace element
concentration was observed.

Another study that examined copper (as well as zinc and magnesium) concentration in both
healthy and malignant tissue indicated a 20% lower copper concentration in malignant tissue
than in healthy tissue (Isbir et al., 1995). Serum copper concentration in cancer patients was also
elevated significantly (31 - 40%). Malignancies included carcinoma of the breast, colon, and
stomach. The authors of the study proposed that copper movement out of and into various tissues
may be due to an increase in cytokine activity, rather than being a feature directly related to
development of the malignancy.

The role of many environmental and biological factors in the development of breast cancer has
been examined. In a study done in Finland (Dabek et al., 1996), dietary copper intake and
subsequent blood copper levels were measured in several groups of women with early breast
cancer. Both total copper (P-copper) and copper ultrafilterable from plasma (P-edu-copper) were
analyzed by Proton Induced X-ray Emission analysis. P-edu-copper appears to more accurately
reflect copper status than does ceruloplasmin level or P-copper because the P-edu-copper reflects
both the major copper enzyme in plasma and dietary intake. This type of measurement may aid
future studies attempting to elucidate coppers role in the occurrence and development of cancer,
especially the dynamics of copper movement into and out of tissues as the disease progresses.
Furthermore, this study concluded that copper metabolism and movement through blood and
tissue may be different in cancer patients than in healthy individuals.

From the literature listed in the 1998 Copper Source Book (Harrison, 1998), it is apparent that
studies of copper specifically designed to test hypotheses for human exposure and disease are
limited in number. Further, there does not appear to be any baseline normal population

COPPER: Environmental Dynamics and Human Exposure Issues


Page 140
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

database available to rigorously compare with data collected in diseased populations. The
available databases cannot be used to make general statements about copper and specific disease
in human subjects. However, they do point out the need to consider the development of
reasonable hypotheses about copper in both healthy and diseased populations before one can
adequately establish exposure-related relationships for this essential element.

8.3 Copper Biomarkers and Exposure Markers in Populations


Several studies have examined whether the concentration of trace metals, including copper,
varies in individuals with different types of diet and lifestyle. Several enzymes depend on copper
for proper function. Some of these are involved in the reduction of free radicals generated during
metabolism (superoxide dismutase). Others are involved in bone growth and strength, immune
function, infant growth, iron transport, blood cell maturation, circulatory system health and
proper neural development and function. Copper status is currently typically monitored by
measurement of serum copper and ceruloplasm concentrations. Severe copper deficiency is
detected more easily by these measurements than is moderate deficiency. A possible alternative
index for evaluating moderate or marginal copper deficiency is measurement of erythrocyte
superoxide dismutase and platelet cytochrome c activity (Olivares & Uauy, 1996b).
Measurement of copper in hair was not considered a reliable index of deficiency since copper
levels are reduced only after prolonged and extreme copper depletion. Environmental deposition
rather than nutritional uptake may also be a factor in hair levels.

Individuals with both higher than normal and lower than normal copper intake are at risk for
disease development. Different biochemical mechanisms would be involved for each of the
extremes. Certain trace elements were measured in the blood of vegetarian nonsmokers
(Krajcovicova Kudlackova et al., 1995). Copper levels were higher in vegetarians than in
nonvegetarians. Copper is utilized by various anti-oxidation enzyme systems as are other trace
elements. There may be reduced amounts of copper in vegetarians if grain intake is high because
grains are phytates that may inhibit copper absorption from the intestine.

A nine-year study was done to determine dietary intakes of eleven nutritional elements for eight
age-sex groups (Pennington & Schoen, 1996). Copper intake for all groups was below the
standard set by the National Academy of Sciences. Dietary intake alone is not sufficient to
trigger changes in health nutrition policy. These values also must be correlated with any
biochemical and/or physiological changes accompanying lowered intake. Copper, along with
magnesium and phosphorus, was not considered a public health problem.

A study in Belgium analyzed the daily dietary intake of various trace metals, including copper
(Robberecht et al., 1996). The technique of duplicate portion sampling was utilized. Another
study done in Germany (Blagoi et al., 1982) also used duplicate food sampling as a technique,
along with the keeping of dietary logs.

Copper intake in a population of children in Duisbeurg, Germany was studied using both
duplicate food sampling and diet records (Laryea et al., 1995). Reference to copper and other
trace element values in foods is not widely available in food tables. The purpose of the study was

COPPER: Environmental Dynamics and Human Exposure Issues


Page 141
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

to measure the actual amount of copper (and other trace elements) in food and to link the data
with consumption patterns. Consumption patterns indicated that there was inadequate intake of
copper in the childrens diet and that new copper data are needed to supplement existing food
tables.

Trace element dietary intake was estimated by coupling data from published tables of trace
element food content with food intake dietary records (Shimbo et al., 1996). The food duplicate
method was used. Almost all elements intake was via diet, with inhalation almost negligible.
Copper intake was found to be higher in lacto-ovo-vegetarian female adolescents than in a
similar population comprised of omnivorous or vegetarian female adolescents (Donovan &
Gibson, 1996).

Copper, as well as other trace elements, plays a major role in proper metabolic function (Airede,
1993). During fetal development, the last trimester of pregnancy is the time that 70% of the total
body store of both copper and zinc is accumulated. Premature infants may not be born with
adequate copper levels, and this would influence many enzymatic functions, particularly the
antioxidant enzymes like superoxide dismutase. Premature infants, at the same time, are exposed
to higher than normal levels of oxygen if the lungs are not fully developed. The combination of
inadequate copper / inadequate antioxidant activity / supplemental oxygen may set the stage for
oxidative tissue damage if trace element supplements are not given.

AAS was used to measure trace element levels in both serum and urine of mature infants (Onag
& Taneli, 1995). The infants were either breast fed or formula fed and were six and twelve
weeks old. The study found that copper excretion levels in breast fed infants were higher than in
formula fed infants at six weeks of age. Serum levels at the same age were not appreciably
different. At twelve weeks, the difference in urine copper excretion levels disappeared with
kidney maturation.

8.3.1 Copper Levels in Human Hair


Lead has no known nutritional value and its toxic effects are well documented. Copper is
considered an essential element needed for the proper functioning of a myriad of enzyme
pathways. Levels of copper in human scalp hair were proposed and used in various studies as an
indicator of dietary status in an occupationally unexposed individual.

Three population groups occupationally unexposed to lead were studied to investigate whether
baseline values could be established for both lead and copper levels in human scalp hair (Sen &
Chaudhuri, 1996). The three groups were geographically distinct (Southern, Southwestern, and
Northern Calcutta), but all were adult males. There were differences in copper levels among the
three groups that could not be explained by occupational or environmental exposure. Baseline
values could not be established due to these unexplained differences. It was recommended that
further studies be done focusing on diet and nutritional status of the subjects.

Copper and zinc levels were measured in human hair (Bertazzo et al., 1996). The subjects were
approximately equally divided among males and females and the ages ranged from 1 to 92 years

COPPER: Environmental Dynamics and Human Exposure Issues


Page 142
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

old. The study was conducted to determine whether pigmentation had an effect on copper and
zinc levels in human hair. There was found to be no appreciable difference in levels of copper
due to sex. Female age influenced copper levels with decreases being detected in ages over 60
years. The decrease in males in the same age group was not as significant. There was an effect
due to pigmentation. Male white hair contained less copper than any other color group. Males
and females with the same hair color had comparable copper levels. In females, the lowest
copper levels were in white hair. A greater number of samples is needed to clarify the effects of
hair color on copper levels.

A study conducted during the period 1982 to 1991 in northeastern Finnish Lapland (Mussalo-
Rauhamaa et al., 1996) found no trend towards either an increase or a decrease in copper levels
in both hair and serum. Waste emitted into the atmosphere from industry located in the
Murmansk region in Russia contains heavy metals and sulphur compounds. This study attempted
to follow the effects of the exposure on the inhabitants of northeastern Finland.

Measurements of trace metals in scalp hair of both professional drivers and university teachers in
Hong Kong was done to try and determine whether environmental factors influenced trace
metals levels (Man et al., 1996). The target groups were both in the age range of 35 - 45 years
old. The drivers were occupationally exposed to outdoor pollution, as well as to pollution
contributed by the vehicle. The teachers were primarily exposed to indoor environments. A
higher metabolic rate was presumed for the drivers and a lower metabolic rate was presumed for
the teachers. No significant difference in copper levels was found between the two groups.

Data for normal trace element ranges of copper in childrens hair is scarce.

A study (Perrone et al., 1996) attempted to establish trace element levels of copper (as well as
other trace elements) in healthy children by using Proton-Induced X-Ray Emission (PIXE).
Several different age groups in Southern Italy were studied: a) full term neonates b) 6-11 year
old children and c) 11-16 year old adolescents. All age groups contained approximately equal
numbers of males and females. Copper levels rise from birth to approximately 8 years of age and
then decrease. Female adolescents have significantly higher copper levels than do the male
adolescents.

8.3.2 Copper Levels in Blood and Serum


Guidelines were presented to establish proper protocols for handling and collection of human
blood and urine and subsequent analysis of these samples for the presence and levels of trace
elements. Copper levels in blood and urine are affected by physiological or pathological
conditions. No special analytical precautions need to be considered and no major methodological
problems are to be expected. The concentration of copper in human serum or plasma ranges from
0.8 to 1.4 mg/l. The mean urinary concentration ranges from 15 to 36 g/24 h.

A more sensitive method for detection of direct-reacting copper in blood plasma was
investigated. The method (Buckley et al., 1996) was based on stable isotope dilution analysis.
Direct reacting copper includes ionic Cu+2 plus all other copper species in plasma that will

COPPER: Environmental Dynamics and Human Exposure Issues


Page 143
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

readily exchange with ionic copper at room temperature. Determining direct-reacting copper
levels may aid in establishment of kinetic models of copper metabolism.

Five physiological factors were evaluated (Fimmel et al., 1994) for their effects on serum copper
concentrations. Age, sex, diurnal rhythm, fasting, and carrier protein were all considered in a
population of subjects in Berlin ranging from 70 to 103 years old. No age correlation to copper
levels was found for either sex, nor was any effect observed in circadian rhythm for individuals
fasting or non-fasting. The effect of physical activity in 50 healthy individuals living in a
polluted environment in Madrid, Spain was studied in relation to plasma trace metals
concentration (Rodriguez Tuya et al., 1996). In professional sportsmen undergoing anaerobic
activity, an increase in plasma copper level was found. No deficiency of copper was found in
athletes at the beginning of the season.

A review of the literature dealing with measurements of trace elements in blood in the Danish
population indicated that copper levels appeared to be higher in females taking oral
contraceptives (Poulsen et al., 1994). The values could not be adequately documented, however,
and could not be considered proper reference values for the general Danish population. The
literature review revealed that more studies need to be done to determine serum copper levels in
well-defined Belgian populations (Cornelis et al., 1994). The same holds true for copper levels in
urine and in erythrocytes. The information that is available deals with random, undefined groups.

The above two review studies reinforce a common theme among the sections reviewed in this
monograph for both environmental and biofluid copper concentrations. There does not exist an
adequate database for comparison of human tissue copper levels in populations potentially at risk
for high or low copper levels to a baseline for the general population. The reasons for this are
probably multifaceted, but evidence from long-term dietary surveys of trace elements shows that
there need be little cause for concern about public health. However, this does not mean that we
should ignore copper. Some of the review studies relating to environmental levels, an exposure
study conducted in Chile (Lagos et al., 1999) on copper in drinking water and the hypothesis
discussed in the NRC report Copper in Drinking Water (National Research Council, 2000)
indicate the potential for problems in individuals or in specific populations.

Of course, copper, although ubiquitous, does not raise the same level of concern for health
effects within the scientific and regulatory community that lead does. However, to address local
situations effectively, there needs to be an effort to provide measured distributions, within a
statistically representative population, for validated biological markers of exposure. These could
assist in identifying the levels of concern relevant to populations potentially at risk for low or
high levels of copper in the body. An example of what exists but needs to be expanded are the
measurements of serum copper concentration in representative members of the U.S. population
taken during the second National Health and Nutrition Survey (NHANES II) conducted between
1976 and 1980 (Schwartz & Weiss, 1990). The average concentration of copper in serum of
9074 adults aged 30 years or more was 126.3 28.6 mg/100 ml. Unfortunately, copper values
were not collected in NHANES III, nor were they included in NHANES 1999 (formerly known
as NHANES IV). These kinds of baseline data are essential for comparisons in future studies of
population risk.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 144
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

Table 9: Tissue and Body Copper Levels in Healthy Adults and Adults with
Wilson's Disease*
Copper (g/g Fresh Tissue)
Tissue/Body Part
Healthy Wilson's Disease
Liver 7.8 99.2
Kidney 2.8 36.2
Heart 2.8 3.2
Spleen 1.5 NR
Lung 1.5 NR
Muscle 1.2 1.2
Stomach 2.3 4.9
Intestine (large) 2.1 1.6
Long bone 2.9 31.0
Brain 5.4 54.9
Nail 15.6 16.4
Skin 0.8 1.1
Adrenal 1.8 2.4
Pancreas 2.4 3.1
Testis 1.1 NR
Ovary 1.7 1.3
Cornea 3.8 35.1
Cartilage 0.55 1.8
Hair 23.4 15.8
Cerebrospinal fluid 0.13 0.15
Bile 2.6 0.7

*Source: Stokinger, 1981. NR = not reported.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 145
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

Table 10: Copper Content of Human Tissues and Body Fluids*


Tissue Mean Content (g/g dry weight)
Normal Wilson's Disease
Adrenal 7.4 17.64
Aorta 6.7 -
Bone 4.2 -
Brain 23.9 -
Caudate Nucleus - 212
Cerebellum - 261
Frontal Lobe Cortex - 118
Globus Pallidus - 254.5
Putamen - 313.5
Cornea - 92.9
Erythrocytes (/100 ml packed red blood cells) 89.1 -
Hair 23.1 -
Heart 16.5 12.7
Kidney 14.9 96.2
Leukocytes (/109 cells) 0.9 -
Liver 25.5 584.2
Lung 9.5 15.5
Muscle 5.4 25.9
Nails 18.1 -
Ovary 8.1 5.2
Pancreas 7.4 4.2
Placenta 13.5 -
Prostate 6.5 -
Skin 2 5.2
Spleen 6.8 5.6
Stomach and Intestines 12.6 22.9
Thymus 6.7 -
Thyroid 6.1 -
Uterus 8.4 -
Aqueous Humor 12.4 -
Bile (Common Duct) 1050 173
Cerebrospinal Fluid 27.8 -
Gastric juice 28.1 -
Pancreatic juice 28.4 -
Plasma, Wilson's disease 50 -
Saliva 31.7 -
Serum
Female 120 -
Male 109 -
Newborn 36 -

*
Source: Scheinberg, 1979 ; Sternlieb & Scheinberg, 1977

COPPER: Environmental Dynamics and Human Exposure Issues


Page 146
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

Tissue Mean Content (g/g dry weight)


Normal Wilson's Disease
Sweat
Female 148 -
Male 55 -
Synovial fluid 21 -
Urine (24 hr) 18 -

COPPER: Environmental Dynamics and Human Exposure Issues


Page 147
Chapter 8
BIOMARKERS OF COPPER EXPOSURE AND EFFECTS

COPPER: Environmental Dynamics and Human Exposure Issues


Page 148
Chapter 9
CONCLUSIONS AND RECOMMENDATIONS FOR FUTURE RESEARCH

9 CONCLUSIONS AND RECOMMENDATIONS FOR FUTURE RESEARCH


Our team considered an extensive array of publications from the list developed by the University
of British Columbia (UBC) for ICA (Copper Sourcebook, 1998), as well as a number of studies
conducted in the 1970's to the 1990's and various publications previously not recorded by UBC.
Critical analysis of the selected items indicate a need for detailed review and evaluation prior to
database entry in order to prioritize these for their utility in a variety of analyses and
assessments. Further, historical context needs to be established for future evaluations of data. It
should be noted that the available studies were not long-term investigations; in fact, most were
short-term studies and the best compilations of data and their evaluation were done more than
twenty years ago. Our analysis has led to the following conclusions:

1. Of those studies listed in the 1998 Source Book, or of many others examined for this
monograph, very few of copper in environmental media or in human tissues and fluids
were designed to study copper specifically. The ICA should consider doing future studies
on environmental media and human tissues and design research based on copper-derived
hypotheses.

2. Evaluation and prioritization of studies using some simple objective criteria regarding the
reporting of data is necessary before including them in a reference database. This would
significantly increase the value of such a database.

3. A most cost effective approach for cataloging every literature reference that might be
related to copper would be to formulate specific questions, e.g. regarding population
exposure to copper in drinking water, spatial and temporal trends of copper in
atmospheric particles, etc. and then to perform a focused information search to identify
and evaluate such data.

4. The ICA should consider coupling the Relational Database Management System for
Copper Distribution and Exposure Studies RDMS-CEDES, developed and used to
complete this monograph, with current approaches used to catalog manuscripts and
reports for the Source Book. The utility of RDMS-CEDES should be tested by applying
the tiered approach to manuscript evaluation and summarization for the UBC acquisitions
in 2000. This will show how, by identifying variables, the selections can be cataloged
according to a variety of indicators. This can lead to more efficient ways to select
manuscripts and data for priority analyses of environmental media and human
exposure/dose.

5. The evaluations described in the Monograph demonstrate that most of the components of
the conceptual model needed to characterize copper in environmental systems lack
critical information. Copper management requires thorough evaluation of which aspects
of copper distribution in the environment and in human populations require further study
in order to improve data for use in the conceptual model. For analyses of potential or
actual exposure and dose, this is important to ensure sufficient information for comparing
copper levels with regulations or guidelines, for maintaining acceptable levels of copper
exposure, and for quantifying reductions in unacceptable copper exposures.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 149
Chapter 9
CONCLUSIONS AND RECOMMENDATIONS FOR FUTURE RESEARCH

6. The environmental flux processes identified in the conceptual framework provide a basis
for examining the movement of copper in various pathways which can eventually affect
human exposure. The ICA should consider developing a model for copper that can be
used for future and current data sets to examine the importance of specific pathways, or
to quantify uncertainties in the database. The model could then be used to define
uncertainties that can be addressed in future studies for specific regions or populations.

7. The ICA should determine if programs exist that can be augmented with additional
resources to test copper-related hypotheses. Further, the ICA should identify long-term
monitoring programs that need to be augmented to ensure that copper distributions in
various media and copper exposure distributions in various populations are available for
use in models to predict future changes and trends over time.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 150
APPENDIX A
A COMPILATION OF COPPER DATA IN FOOD, DRINKING WATER, AND AIR

10 APPENDIX A: A COMPILATION OF COPPER DATA IN FOOD, DRINKING


WATER, AND AIR
(for Human Exposure Assessments)

Table 11: Concentrations of Copper in Air*


Concentrationa (ng/m3) Reference
Date/Sample Location Comments
(mean) [median]
1977, United States [133], 43390, 115699 4648 samples, Evans et al., 1984
Urban (207.5), 3296max National Survey
1978, United States [138], 43090, 97599 3614 samples, Evans et al., 1984
Urban (200.8), 4625max National Survey
1979, United States [96], 36390, 51999 2507 samples, Evans et al., 1984
Urban (259.3), 1627max National Survey
1977, United States [120], 45090, 106599 709 samples, Evans et al., 1984
Nonurban (193.2), 18,706max National Survey
1978, United States [179], 60790, 139699 458 samples, Evans et al., 1984
Nonurban (265.7), 1396max National Survey
1977, United States [76], 32290, 64599 235 samples, Evans et al., 1984
Nonurban (141.7), 4003max National Survey
Urban 20-200, [50] Representative values Davies & Bennett, 1985
Rural 5-50, [20] Representative values Davies & Bennett, 1985
Remote 0.29-12 Values from Schroeder et al., 1987
literature survey
Rural 3-280 Values from Schroeder et al., 1987
literature survey
Urban Canada 17-500
Urban U.S.A. 3-5140
Urban Europe 13-2760
Urban Other 2.0-6810
1979, Smokey Mountain (1.6) Above canopy, crustal Davidson et al., 1985
Remote National Park enrichment factor 31
1980, Olympic National 3.3-6.7, (5.6) Crustal enrichment Davidson et al., 1985
Remote Park factor 76
1981, 1982, Camden, NJ 16.0-18.0 Seasonal variations Lioy et al., 1987
Summer 100.0b max noted
Urban
1981, 1982, Elizabeth, NJ 21.0-29.0, 120.0 max Seasonal variations Lioy et al., 1987
Summer noted
Urban
1981, 1982, Newark, NJ 25.0-33.0, 131.0 max Seasonal variations Lioy et al., 1987
Summer noted
Urban
1981, 1982, Ringwood, NJ 13.0-63.0, 77.0 max Seasonal variations Lioy et al., 1987
Summer noted
Rural
1982, 1983, Camden, NJ 17.0-21.0, 231.0 Seasonal variations Lioy et al., 1987
Winter noted
Urban
1982, 1983, Elizabeth, NJ 28.0-36.0, 493 0 max Seasonal variations Lioy et al., 1987
Winter noted
Urban

*
Adapted from ATSDR, 1990

COPPER: Environmental Dynamics and Human Exposure Issues


Page 151
APPENDIX A
A COMPILATION OF COPPER DATA IN FOOD, DRINKING WATER, AND AIR

1982, 1983, Newark, NJ 21.0-27.0, 380.0 max Seasonal variations Lioy et al., 1987
Winter noted
Urban
1982, 1983, Ringwood, NJ 6.0-18.0, 29.0 max Seasonal variations Lioy et al., 1987
Winter noted
Rural

COPPER: Environmental Dynamics and Human Exposure Issues


Page 152
APPENDIX A
A COMPILATION OF COPPER DATA IN FOOD, DRINKING WATER, AND AIR

Table 12: Concentrations of Copper in Water


Sample Concentration (ppb)
Location Comments Reference
Type/Source Range (mean) [median]
Drinking water
Private Wells Nova Scotia, 40-200 at tap, running water Maessen et al., 1985
4 communities
Private Wells Nova Scotia, 130-2450, 53% of at tap, standing water Maessen et al., 1985
4 communities samples >1000 ppb
Not specified Seattle, WA (160) running water Maessen et al., 1985
Not specified Seattle, WA (450), 24% of standing water Maessen et al., 1985
samples >1000 ppb
River water Canada (National 5-530 [5] raw water Meranger et al., 1979
Survey)
River water Canada (National 5-100 [5] treated water Meranger et al., 1979
Survey)
River water Canada (National 5-220 [20] distributed water Meranger et al., 1979
Survey)
Lake Water Canada (National 5-80 [5] raw water Meranger et al., 1979
Survey)
Lake Water Canada (National 5-100 [5] treated water Meranger et al., 1979
Survey)
Lake Water Canada (National 5-560 [40] distributed water Meranger et al., 1979
Survey)
Well water Canada (National 5-110 [5] raw water Meranger et al., 1979
Survey)
Well water Canada (National 5-70 [5] treated water Meranger et al., 1979
Survey)
Well water Canada (National 10-260 [75] distributed water Meranger et al., 1979
Survey)
Groundwater
Representative New Jersey [5.0] 1063 samples,
sample 90th percentile 64.0 ppb,
maximum 2783 ppb,
groundwater may or
may not be used
for drinking water
Surface water
U.S. Geological United States (4.2) [4.0] 53,862 occurrences Eckel & Jacob, 1988
Survey stations
Representative New Jersey [3.0] 590 samples,
sample 90th percentile 9.0 ppb,
maximum 261 ppb
Surface, marine E. Arctic Ocean (0.126) 26 locations, Mart et al., 1984
0.5-1 m depth
Surface, marine Atlantic Ocean 0.0572-0.210 20 sites, Mart et al., 1984
2 cruises, 0-1 m
Pond Massachusetts <10-105 low in summer, high Kimball, 1973
Lakes Canada 1-8 (2) in winter, Reed & Henningson, 19
acid sensitive lakes 1984

COPPER: Environmental Dynamics and Human Exposure Issues


Page 153
APPENDIX A
A COMPILATION OF COPPER DATA IN FOOD, DRINKING WATER, AND AIR

Table 13: Copper Content of Selected Foods*



Food Copper Content [mg/100g]
Milk and Milk Products
Low-fat milk, 2% fat 0.003 (0.003)
Yogurt, plain, low-fat 0.004 (0.005)
Evaporated milk 0.009 (0.011)
Cottage cheese 4% milkfat 0.016 (0.014)
Cheddar cheese 0.040 (0.024)

Meat. Poultry. Fish and Eggs


Beef, ground 0.080 (0.011)
Beef, chuck, roasted 0.098 (0.015)
Pork, ham, cured, cooked 0.095 (0.025)
Lamb chop, fried 0.171 (0.034)
Chicken, roasted 0.069 (0.026)
Liver (beef/calf), fried 6.089 (1.448)
(Cod/Haddock) fillet, cooked 0.031 (0.014)
Tuna, canned in oil, drained 0.051 (0.012)
Eggs, soft boiled 0.066 (0.010)

Legumes
Pinto beans, cooked 0.263 (0.045)
Lima beans, cooked 0.205 (0.048)
Green peas, frozen, boiled 0.103 (0.017)

Nuts and Nut products


Peanuts, dry roasted, salted 0.675 (0.095)
Pecans, unsalted 1.245 (0.190)

Grains and Grain Products


Rice, white, enriched, cooked 0.079 (0.024)
Oatmeal, cooked 0.099 (0.020)
White bread, enriched 0.126 (0.014)
Corn bread 0.060 (0.008)
Whole wheat bread 0.250 (0.036)
Rye bread 0.192 (0.030)
Egg noodles, enriched, cooked 0.080 (0.037)
Shredded wheat cereal 0.476 (0.023)

Fruits
Apple 0.026 (0.015)
Orange 0.042 (0.014)
Banana 0.143 (0.037)
Peach 0.055 (0.018)
Grapes 0.089 (0.048)
Raisins 0.315 (0.045)

Fruit Juices and Drinks


Orange juice, frozen, reconstituted 0.022 (0.010)
Apple juice 0.010 (0.008)

* Source: (Pennington et al., 1986)


Mean (standard deviation)

COPPER: Environmental Dynamics and Human Exposure Issues


Page 154
APPENDIX A
A COMPILATION OF COPPER DATA IN FOOD, DRINKING WATER, AND AIR


Food Copper Content [mg/100g]
Lemonade, frozen, reconstituted 0.001 (0.002)

Vegetables and Vegetable Products


Spinach, boiled 0.083 (0.036)
Broccoli, boiled 0.028 (0.013)
Tomato 0.054 (0.020)
Green pepper 0.076 (0.032)
Carrots 0.049 (0.023)
Potato, boiled without peel 0.068 (0.033)
Potato, baked with peel 0.115 (0.051)

Mixed Dishes
Beef and vegetable stew 0.076 (0.021)
Pizza, cheese 0.123 (0.019)
Beef and beans 0.167 (0.011)
Hamburger sandwich, quarter pound, garnished 0.089 (0.015)

Soups
Chicken noodle soup 0.010 (0.008)
Tomato soup 0.035 (0.010)
Vegetable beef soup 0.16 (0.009)

Condiments, Fats, and Sweeteners


Corn oil 0.009 (0.010)
Mayonnaise 0.012 (0.014)
Grape jelly 0.019 (0.009)
Catsup 0.221 (0.069)

Desserts
Ice cream, chocolate 0.149 (0.057)
Cookies, chocolate chip 0.228 (0.055)
Candy, plain milk chocolate 0.453 (0.043)

Beverages
Carbonated soda, cola 0.002 (0.002)
Coffee 0.001 (0.002)
Tea 0.006 (0.004)
Beer 0 005 (0.003)

COPPER: Environmental Dynamics and Human Exposure Issues


Page 155
APPENDIX A
A COMPILATION OF COPPER DATA IN FOOD, DRINKING WATER, AND AIR

Table 14: Copper Content of Selected Foods per 100 Grams*


Cu Cu
Food Name Food Name
(mg) (mg)
Dairy products Meat/Fish
Egg, whole, hard boiled, large 0.01 Pot roast, arm, beef, cooked 0.16
Cheese, cottage, uncreamed 0.03 Hamburger patty, beef/lean, broiled 0.07
Cream, coffee, table, light 0.01 Steak, sirloin, lean, broiled 0.15
Cream, sour, cultured 0.02 Chicken, leg, no skin, roasted 0.08
Milk, buttermilk, fluid 0.01 Chicken, breast, no skin, roasted 0.05
Milk, whole, 3.3% fat, fluid 0.01 Lamb, all cuts, lean/fat, cooked 0.12
Milk, nonfat/skim, fluid 0.01 Turkey, dark meat, no skin, roasted 0.16
Milk, whole, low sodium 0.01 Turkey, light meat, no skin, roasted 0.04
Veal, all cuts, lean, cooked 0.12
Fats Bluefish, cooked, dry heat 0.07
Butter, regular 0.02 Flatfish, cooked, dry heat 0.03
Vegetable oil, corn 0.00 Cod, cooked, dry heat 0.04
Vegetable oil, olive 0.00 Halibut, cooked, dry heat 0.04
Shortening, soybean/cotton seed 0.00 Shrimp, mixed species, cooked 0.19
Margarine, regular, hard, unsalted 0.00 Tuna, can/oil, drained 0.07
Mayonnaise, soy, commercial 0.01 Tuna, can/water, low sodium 0.05

Cereals Sweets
Bran flakes, Kellogg's 0.40 Honey, strained/extracted 0.04
Corn flakes, Kellogg's 0.10 Ice milk, vanilla, hard, 4% fat 0.01
Cream of Rice, cooked 0.03 Ice cream, vanilla, hard, 11% fat 0.02
Cream of Wheat, instant 0.04 Ice cream, vanilla, rich, 16% fat 0.02
Farina, cooked, enriched 0.01 Jams/ preserves, regular 0.10
Oatmeal, cooked 0.06 Sherbet, orange, 2% fat 0.03
Wheat, puffed, plain 0.61 Sugar, brown, pressed down 0.30
Wheat, shredded, biscuit 0.50 Sugar, white granulated 0.04
Rice Krispies, Kellogg's 0.20
Juices
Breads, cookies, crackers Apple juice, can and bottle 0.02
Bread, white, soft 0.13 Apricot nectar, can 0.07
Bread, whole-wheat, soft 0.28 Cranberry juice cocktail, bottle 0.02
Crackers, graham, plain 0.20 Grape juice, can and bottle 0.03
Crackers, low sodium/whole-wheat 0.44 Grapefruit juice, can, unsweetened 0.04
Crackers, saltines 0.20 Lemon juice, can and bottle 0.04
Muffin, English, plain 0.13 Orange juice, from frozen 0.04
concentrate
Bread, Italian, enriched 0.19 Pear nectar, can 0.07
Roll, hard, enriched 0.16 Pineapple juice, can 0.09
Roll, hamburger/hot dog 0.11 Prune juice, can 0.07
Cookies, vanilla wafer, lower fat 0.10 Tomato juice, can 0.10
Tomato juice, can, low sodium 0.10

*
Data excerpted from Shils et al., 1999, from tables originally compiled by the Nutrition Coordinating Center, University of Minnesota.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 156
APPENDIX A
A COMPILATION OF COPPER DATA IN FOOD, DRINKING WATER, AND AIR

Cu Cu
Food Name Food Name
(mg) (mg)
Vegetables, Pasta, Rice Fruits
Asparagus, can, spears 0.10 Apples, raw, unpeeled 0.04
Asparagus, frozen, cooked 0.17 Apples, raw, peeled 0.03
Beans, snap, green, can 0.04 Applesauce, can, unsweetened 0.03
Beans, snap, green, boiled, drained 0.06 Apricots, can, light syrup 0.08
Beans, snap, wax, boiled, drained 0.10 Bananas, raw, peeled 0.10
Beets, can, whole 0.06 Blueberries, raw 0.06
Beets, boiled, drained 0.07 Cherries, sweet, can/juice pack 0.07
Broccoli, boiled, drained 0.04 Grapefruit, raw, all varieties 0.05
Cabbage, common, boiled, drained 0.01 Oranges, raw, all varieties 0.04
Carrots, can, sliced, drained 0.10 Peaches, raw, whole 0.07
Carrots, boiled, drained 0.13 Peaches, can, light syrup 0.05
Carrot, raw, whole, scraped 0.05 Pears, raw, unpeeled 0.11
Cauliflower, boiled, drained 0.03 Pineapple, can/juice pack 0.09
Celery, raw, stalk 0.03 Strawberries, raw, whole 0.05
Corn, sweet, can, drained 0.06
Corn, sweet, boiled, drained 0.04
Cucumber, raw, sliced, with peel 0.03
Peas, green, can, drained 0.08
Peas, green, boiled, drained 0.17
Potato, baked, with skin 0.30
Potato, boiled, peeled before cooked 0.17
Tomato, raw, red, ripe 0.07
Tomato, red, can, stewed 0.11
Tomato, can, no salt added 0.11
Noodles, egg, enriched, cooked 0.09
Rice, white, enriched, parboiled, 0.09
cooked

COPPER: Environmental Dynamics and Human Exposure Issues


Page 157
APPENDIX A
A COMPILATION OF COPPER DATA IN FOOD, DRINKING WATER, AND AIR

COPPER: Environmental Dynamics and Human Exposure Issues


Page 158
APPENDIX B
A COMPILATION OF COPPER DATA SOURCES

11 APPENDIX B: A COMPILATION OF COPPER DATA SOURCES


USGS Water Quality Monitoring Network (WQN) MT2 Data
Databases Conc. Geographical
Data Group Searched Media (units) Information Frequency Period
Region 1 New 35 Surface ug/l Long/Lat quarterly 73-95
England Water 1 cross-sxn NASQAN
int. sample 62-95 HBN
Region 2 Mid- 39 Surface ug/l Long/Lat quarterly1 73-95
Atlantic Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 3 South 82 Surface ug/l Long/Lat quarterly1 73-95
Atlantic Gulf Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 4 Great Lakes 63 Surface ug/l Long/Lat quarterly1 73-95
Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 5 Ohio 50 Surface ug/l Long/Lat quarterly1 73-95
Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 6 Tennessee 12 Surface ug/l Long/Lat quarterly1 73-95
Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 7 Upper 36 Surface ug/l Long/Lat quarterly1 73-95
Mississippi Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 8 Lower 42 Surface ug/l Long/Lat quarterly1 73-95
Mississippi Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 9 Souris-Red- 13 Surface ug/l Long/Lat quarterly1 73-95
Rainy Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 10 Missouri 83 Surface ug/l Long/Lat quarterly1 73-95
Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 11 Arkansas 50 Surface ug/l Long/Lat quarterly1 73-95
White-Red Region Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 12 Texas 33 Surface ug/l Long/Lat quarterly1 73-95
Gulf Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 13 Rio 20 Surface ug/l Long/Lat quarterly1 73-95
Grande Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 14 Upper 17 Surface ug/l Long/Lat quarterly1 73-95
Colorado Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 15 Lower 28 Surface ug/l Long/Lat quarterly1 73-95
Colorado Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 16 Great 21 Surface ug/l Long/Lat quarterly1 73-95
Basin Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 17 Pacific 59 Surface ug/l Long/Lat quarterly1 73-95
Northwest Water cross-sxn NASQAN
int. sample 62-95 HBN

COPPER: Environmental Dynamics and Human Exposure Issues


Page 159
APPENDIX B
A COMPILATION OF COPPER DATA SOURCES

USGS Water Quality Monitoring Network (WQN) MT2 Data


Databases Conc. Geographical
Data Group Searched Media (units) Information Frequency Period
Region 18 California 30 Surface ug/l Long/Lat quarterly1 73-95
Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 19 Alaska 13 Surface ug/l Long/Lat quarterly1 73-95
Water cross-sxn NASQAN
int. sample
62-95 HBN
Region 20 Hawaii 9 Surface ug/l Long/Lat quarterly1 73-95
Water cross-sxn NASQAN
int. sample 62-95 HBN
Region 21 Carribean 6 Surface ug/l Long/Lat quarterly1 73-95
Water cross-sxn NASQAN
int. sample 62-95 HBN
Note: data consist of surface water dissolved and trace metals. Each file contains one record
for each date and time sampled. Water quality stream flow data were available for 679
locations within the United States for 63 Hydrologic Benchmark Network (HBN) stations
from 1962-1995 and 618 National Stream Quality Accounting Network (NASQAN) stations
from 1973-1995. Sediment data are also available.

Quarterly denotes one integrated coss-sectional sample taken every four months at the same time and location of
previous samples. A sample was collected in both horizontal and vertical planes of the water column.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 160
APPENDIX B
A COMPILATION OF COPPER DATA SOURCES

USGS National Geochemical Atlas Data


Databases Conc. Geographical
Data Group Searched Media (units) Information Frequency Period
USGS/NGA 1 Soils ppm Long/Lat N/A 70s-80s
Note: NGA contains data from 260,000 samples collected between the late seventies
and early eighties. Data include stream, lake, pond and spring sediment and soil
samples.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 161
APPENDIX B
A COMPILATION OF COPPER DATA SOURCES

NOAA Ocean Resources Conservation and Assessment (ORCA) Data


Trends and Status, Mussel Watch and Benthic Surveillance
Databases Conc. Geographical
Data Group Searched Media (units) Information Frequency Period
Bioeffects Biscayne Bay Sediment mg/kg Station ID NI2 96
Boston Harbor Sediment ug/g 93
Hudson- Sediment mg/kg 91
Raritan
Sediment ug/g 92
Sediment ug/g NI2
Los Angeles
L.I. Sound
CADS 6 Water NI2 County ID NI2 ---
Shellfish
Maine 7 Water lbs Station ID quarterly ---
Mexico 8 Water lbs Long/Lat quarterly ---
NsandT Sites 3 Fish tissue NI2 Long/Lat monthly 86-96
Fish liver
Sediment Long/Lat 84-92

NI2 Long/Lat NI2 84-95


Tampa Bay 2 Sediment ppm Long/Lat NI2
Note: sediment data files were also available for Mussel Watch data for 1986-1991
and benthic surveillance data for 1984-1988.
2
NI - not identified as yet

COPPER: Environmental Dynamics and Human Exposure Issues


Page 162
APPENDIX B
A COMPILATION OF COPPER DATA SOURCES

EPA Environmental Monitoring Assessment Program (EMAP)


Databases Conc. Geographical
Data Group Searched Media (units) Information Frequency Period
Carolinian 2 Fish tissue ug/g Station ID NI2 95-97
Province
Estuaries 40 Sediment ug/g Station ID NI2 91-94
Louisianian Fish tissue ug/g 91-94
Province
Estuaries Virginian 38 Sediment ug/g Station ID NI2 90-93
Province Fish tissue ug/g 91
Region 6 10 Sediment ug/g Station ID NI2 93-94
Galveston Bay Fish tissue 93
Region 8 Colorado 2 Sediment ug/g Station ID NI2 94-95
Streams Water ug/l 94-95
Region 3 DE/MD 21 Sediment ug/g Station ID NI2 93
Coastal Bays
Great Lakes 20 Search process on-going
Region 1 Maine 13 Search process on-going
Region 4 South 4 Search process on-going
Florida
Region 10 9 Search process on-going
Surface Waters 28 Search process on-going
2
NI - not identified as yet

COPPER: Environmental Dynamics and Human Exposure Issues


Page 163
APPENDIX B
A COMPILATION OF COPPER DATA SOURCES

Other EPA Databases


Databases Conc. Geographical
Data Group Searched Media (units) Information Frequency Period
Toxics Release 1 Air Emis in Long/Lat Yearly 87-96
Inventory Land Tons/yr Avgs.
Water
NHEXAS 1 Water ug/l County ID Once 95-97
Food ug/kg Daily
Bev ug/kg Daily
Note: data set includes copper concentrations measured in drinking water, food and
beverages from 2154 samples collected between 1995 and 1997.
AIRS 32 TSP ug/m3 Long/Lat every 6th 84-99
3
PM10 ug/m day - 24hr.
avg.
Note: the database contains one-day average (every sixth day) air monitoring data
from 1984-1999 for copper as total suspended particulates and PM10.
SDWIS* 1a Surface N/A Station ID Yearly 74-99
Water complian Avgs.
ce data
*Note: information within this database includes the name of the public water
system information about the type of area served by the water system (e.g.,
households, schools, restaurants, gas stations, or rest areas), number of people
served by the water system, operating season (year-round or seasonal), who
regulates the water system (typically, states regulate systems within their
jurisdictions; EPA currently regulates Tribal systems and systems in Wyoming),
when (or if) a water system has violated any national drinking water standard what
(if any) follow-up actions, including enforcement actions, have been taken to make
sure the water system returns to compliance following a violation.
a
Data in SDWIS have been requested from USEPA.
STORET** 1b Surface N/A Station ID and Yearly 65-99
Water Long/Lat
**Note: a repository for water quality, biological, and physical data. STORET
contains two data systems, legacy data which contain historical water quality data
from the early part of the 20th century through 1998 and modernized STORET
data collected since 1999. Both systems contain raw biological, chemical , and
physical data on surface and ground water encompassing all fifty states, territories,
and jurisdictions of the U.S., along with portions of Canada and Mexico.
b
Data in STORET have been requested from USEPA.
2
NI - not identified as yet

COPPER: Environmental Dynamics and Human Exposure Issues


Page 164
APPENDIX B
A COMPILATION OF COPPER DATA SOURCES

CDC Databases
Databases Conc. Geographical
Data Group Searched Media (units) Information Frequency Period
NHANES II* 1 Blood ug/dl County ID NI2 76-80
Food mg
Beverages
NHANES III* 1 Food ug/dl County ID Daily 88-94
Beverages mg
*Note: the data collected through the NHANES II project (1976-1980) contain
copper concentrations in food, beverages and blood serum from over 27,000
participants. Data from NHANES III (1988-1994) however only contain food and
beverage copper concentrations for 33,994 individuals.
2
NI - not identified as yet

COPPER: Environmental Dynamics and Human Exposure Issues


Page 165
APPENDIX B
A COMPILATION OF COPPER DATA SOURCES

Selected European Databases


Databases Conc. Geographical
Data Group Searched Media (units) Information Frequency Period
Geochemical Sediment g/ton 91-95
Atlases of
Hungary, Slovakia,
Sweden, etc. *
SWAD 7 Belgium ug/L Station ID monthly 90-97
(Surface WAter Flanders
Database) total
dissolve

Sweden ug/L Long/Lat. bi- 92-97


total monthly
bi-
German ug/L Region ID monthly 96
total
bi-
Neth. total ug/L Region ID monthly 93-97
dissolve

Nor. monthly 94-98


Ireland ug/L Region ID
total
dissolve

United yearly
King. ug/L Region ID 88-95
total
dissolve

Spain monthly
dissolve ug/L Region ID 85-94
Note: SWAD contains a compilation of environmental concentrations of zinc and
other heavy metal concentrations in European surface waters. In addition to
information on metal concentrations, data on physico-chemical characteristics of
these surface waters were also collected. The actual geographical coordinates for
most of the monitored data were not included in SWAD; however it is expected that
geographical information was included within the original (source) data files that
were used to develop SWAD.

*
At press time, these atlases were not yet available, but the information will be included in a future edition.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 166
APPENDIX C
ACRONYMS, ABBREVIATIONS AND GLOSSARY OF TERMS

12 APPENDIX C: ACRONYMS, ABBREVIATIONS AND GLOSSARY OF TERMS

Acronyms and Abbreviations


AAS: Atomic Absorption Spectroscopy
AE: Achrodermatitis Enterophatica
AIRS: Aerometric Information Retrieval System
AREI: Acceptable Range of Exposure/Intake
ASTM: American Society for Testing and Materials
ATEOS: Airborne toxic element and organic substances
ATSDR: Agency of Toxic Substances and Disease Registry
BBR: Background-Based Ranking
BCF: Bioconcentration Factor
CalDHCD: California Department of Housing and Community
Development
CARD: Contract Laboratory Program Analytical Results Data Base
CDC: Centers for Disease Control
CERCLA: Comprehensive Environmental Response, Compensation, and
Liability Act
CHEM: Consumption Habit Exposure Model
CLP: Contract Laboratory Program
CNS: Central Nervous System
COD: Chemical Oxygen Demand
C-O-N-S-P ratio: Carbon-Oxygen-Nitrogen-Sulfur-Phosphorous ratio
CPS Composite Proportional Method
CVPC: Cross-Linked Polyvinyl Chloride
CWA: Clean Water Act
CWS: Community Water Systems
Da: Daltons, a measure of atomic mass
DTPA: Diethylenetriamine Pentaacetic Acid
E/I: Exposure/Intake
EC-SOD: Extracellular Superoxide Dismutase
ED: Effective Dose
EDTA: Ethylene Diamine Tetraacetic Acid
EIR: Environmental Impact Report
EMAP: Environmental Monitoring Assessment Program

COPPER: Environmental Dynamics and Human Exposure Issues


Page 167
APPENDIX C
ACRONYMS, ABBREVIATIONS AND GLOSSARY OF TERMS

EPR: Electron Paramagnetic Resonance


EU: European Union
FAO: Food and Agriculture Organization
FIFRA: Federal Insecticide, Fungicide, and Rodenticide Act
FNB: Food and Nutrition Board
FORTRAN: Computer language for scientific programming
IAEA: International Atomic Energy Agency
ICA: International Copper Association
ICC: Indian Childhood Cirrhosis
LCRMR: Lead and Copper Rule Minor Revisions
LD: Lethal Dose
LDL: Low Density Lipoprotein
LOAEL: Lowest dose at which adverse health effects are observed
MATLAB: Interpretive computer language for scientific programming
MCLG: Maximum Contaminant Level Goal
MF: Modifying Factors
MINTEQA2: Computer program to predict chemical speciation at
equilibrium
MT: Metallothionein
MT2: As in USGS Water Quality Monitoring Network (WQN) MT2
Data
NHANES II/III: National Health and Nutrition Examination Survey; Phase II
and III
NHEXAS: National Human Exposure Assessment Survey
NIOSH: National Institute for Occupational and Safety and Health
NOAEL: Highest dose as which there are no observed adverse effects
NOES: National Occupational Exposure Survey
NPDWR: National Primary Drinking Water Regulations
NPL: National Priorities List
NRC: National Research Council
NTNCWS: Non-Transient Non-Community Water Systems
ORCA: Ocean Resources Conservation and Assessment
OSF: Oral Submucous Fibrosis
PHGs: Public Health Goals
PIXE: Proton-Induced X-Ray Emission

COPPER: Environmental Dynamics and Human Exposure Issues


Page 168
APPENDIX C
ACRONYMS, ABBREVIATIONS AND GLOSSARY OF TERMS

POTW: Publicly Owned Treatment Works


PPL: Priority Pollutants List
RCRA: Resource Conservation and Recovery Act
RDA: Recommended Daily Allowance
RDI: Recommended Dietary Intake
RDMS-CEDES: Relational Database Management System-Copper
Environmental Distribution Exposure Studies
RfD: Reference Dose (an estimate of dose that is likely to be without
appreciable risk of adverse effects in humans)
RQ: Reportable Quantities
SD: Standard Deviation
SIC: Standard Industrial Codes
SOD: Superoxide Dismutase
SPM: Suspended Particulate Matter
STELLA: Interpretive computer language for scientific programming
STORET: A USEPA repository for water quality, biological, and physical
data
SWAD: European Surface Water Database
SX-EW: Solvent Extraction-Electrowinning
TRI: Toxics Release Inventory
UF: Uncertainty Factors
USEPA SDWIS: Safe Drinking Water Information System; a USEPA national
database storing information about the nations drinking water
US NOAA: United States National Oceanic and Atmospheric
Administration
USDA: United States Department of Agriculture
USEPA: United States Environmental Protection Agency
USGS: United States Geological Survey
WCHS: Water Consumption Habit Survey
WHO: World Health Organization
WQN: Water Quality Monitoring Network

COPPER: Environmental Dynamics and Human Exposure Issues


Page 169
APPENDIX C
ACRONYMS, ABBREVIATIONS AND GLOSSARY OF TERMS

Glossary of Terms
abiotic: pertaining to non-living matter
activated sludge: sludge processed to enhance adherence of chemicals to
particles in sludge
amino acid: an amphoteric organic acid containing the amino group NH2;
especially: any of the alpha-amino acids that are the chief
components of proteins and are synthesized by living cells or
are obtained as essential components of the diet
anthropogenic: resulting from human activity
benthic: organisms living at the bottom of a body of water
bioavailable: amount of a substance that is available for uptake into an
organism
biogenic: resulting from the activity of living organisms
bioindicator: a biological indicator of exposure, or effect
biomass: mass of living organisms
biomonitors: organism that could serve as sentinels of exposure
biota: living organisms
ceruloplasmin: a blue copper-binding serum alpha globulin with enzymatic
activity
crustal: pertaining to the top layer of the earth
cuproenzyme: copper containing enzyme
cytochrome: a family of enzymes
detritus: organic matter that was once living
diagenesis: process of rock formation from sediments
dismutase: a metal-containing enzyme that reduces potentially harmful free
radicals of oxygen formed during normal metabolic cell
processes to oxygen and hydrogen peroxide
enteropathogenic: tending to produce disease in the intestinal tract
enzyme: any of numerous complex proteins that are produced by living
cells and catalyze specific biochemical reactions at body
temperatures
exogenous: introduced from or produced outside the organism or system
extrahepatic: external to the liver
ferritin: iron-containing protein

COPPER: Environmental Dynamics and Human Exposure Issues


Page 170
APPENDIX C
ACRONYMS, ABBREVIATIONS AND GLOSSARY OF TERMS

glutathione: a peptide that contains one amino-acid residue each of glutamic


acid, cysteine, and glycine, that occurs widely in plant and
animal tissues, and that plays an important role in biological
oxidation-reduction processes and as a coenzyme
hemolytic anemia: anemia caused by excessive destruction (as in chemical
poisoning, infection, or sickle-cell anemia) of red blood cells
hepatic: of, relating to, affecting, associated with, supplying, or draining
the liver
humic: organic matter in soil
hydrosphere: the aqueous vapor of the atmosphere
hypercholesterolemia: the presence of excess cholesterol in the blood
imidazole: a white crystalline heterocyclic base C3H4N2 that is an
antimetabolite related to histidine
leachate: a solution or product obtained by leaching
leaching: removing (nutritive or harmful elements) from soil by
percolation
macronutrient: a chemical element of which relatively large quantities are
essential to the growth and health of a plant
mariculture: the cultivation of marine organisms in their natural
environment
mean residence time: the average time a substance spends in a system
metalloenzyme: an enzyme which is a metalloprotein
metallothionein: enzymes involved in transporting metals in organisms
micronutrient: a chemical element present in minute quantities; one used by
organisms and held essential to their physiology
ombrogenic: from rain water
pedosphere: the earth's soil layer
peptide: any compound in which two or more amino-acids are linked
together by peptide bonds
podzol: type of acid loamy sand
protein: complex substance that consists of amino-acid residues joined
by peptide bonds
solvated: bound to solvent molecules
tailings: residue from mining operations
trophic: of or relating to nutrition
xenobiotic: a chemical foreign to an organism

COPPER: Environmental Dynamics and Human Exposure Issues


Page 171
APPENDIX C
ACRONYMS, ABBREVIATIONS AND GLOSSARY OF TERMS

COPPER: Environmental Dynamics and Human Exposure Issues


Page 172
REFERENCES

13 REFERENCES
1. Aalbers, T.G., Houtman, J.P.W., and Makkink, B. 1987. Trace-element concentrations in
human autopsy tissue. Clin. Chem. 33:2057-2064. ICA no. A-6272064.

2. Adamson, M., Reiner, B., Olson, J.L., Goodman, Z., Plotnick, L., Bernardini, I., and Gahl,
W.A. 1992. Indian childhood cirrhosis in an American child. Gastroenterology 102:1771-1777.
ICA no. A-1367.

3. Adriano, D.C. 1986. Copper. In Trace Elements in the Terrestrial Environment. NY: Springer-
Verlag.

4. Afseth, J., Rolla, G., Helgeland, K., and Opperman, R.V. 1986. Aspects of Cu2+ and Zn2+ in
mouth rinses with regards to dental health. Acta Pharmacol. Toxicol. Suppl. 59:300-304. ICA no.
A-609.

5. Airede, A.K.I. 1993. Copper, zinc and superoxide dismutase activities in premature infants: A
review. East Afr. Med. J. 70 (7):441-444. ICA no. A-2199.

6. AJCN. 1998. Genetic and Environmental Determinants of Copper Metabolism. Proceedings of


an international conference. Bethesda, Maryland. March 18-20, 1996. American Journal of
Clinical Nutrition (Supplement) 67 (5):951-1102.

7. Alam, I.A., and Sadiq, M. 1989. Metal contamination of drinking water from corrosion of
distribution pipes. Environ. Pollut. 57:167-178. ICA no. A-964.

8. Alderdice, D.F., and McLean, W.E. 1982. A review of the potential influence of heavy metals
on salmonid fishes in the Campbell River, Vancouver Island, British Columbia. Can. Tech. Rep.
Fish Aquat. Sci. 1104:1-60. ICA no. A-209.

9. Aldini, R., Bigliardi, D., and Zanotti, A. 1987. Heavy metals in canned foods. Riv. Ital.
Sostanze Grasse 64:49-52. ICA no. A-821.

10. Aljajeh, I.A., Mughal, S., Al-Tahou, B., Ajrawl, T., Esmail, E.A., and Nayak, N.C. 1994.
Indian childhood cirrhosis-like liver disease in an Arab child; A brief report. Virchows Archiv.
424:225-227. ICA no. A-1669.

11. Allen, H.E., and Hansen, D.J. 1996. The importance of trace metal speciation to water quality
criteria. Water Environ. Res. 68 (1):42-54. ICA no. A-2227.

12. Al-Rajhi, M.A., Seaward, M.R.D., and Al-Aamer, A.S. 1996. Metal levels in indoor and
outdoor dust in Riyadh, Saudi Arabia. Environ. Int. 22 (3):315-324. ICA no. A-2245.

13. Al-Saleh, I.A. 1996. Trace elements in drinking water coolers collected from primary
schools, Riyadh, Saudi Arabia. Sci. Total Environ. 181 (3):215-221. ICA no. A-2270.

14. Amsden, M.P., Sweetin, R.M., and Treilhard, D.G. 1978. Selection and design of Texas-Gulf
Canada's copper smelter and refinery. J. Metals 30:16-26. ICA no. A-044.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 173
REFERENCES

15. Anderson, C.J., Connett, J.M., Schwarz, S.W., Rocque, P.A., Guo, L.W., Philpott, G.W.,
Zinn, K.R., Meares, C.F., and Welch, M.J. 1992. Copper-64-labelled antibodies for PET
imaging. J. Nucl. Med. 33:1685-1691. ICA no. A-1463.

16. Anderson, J.R., Aggett, F.J., Buseck, P.R., Germani, M.S., and Shattuck, T.W. 1988.
Chemistry of individual aerosol particles from Chandler, Arizona, and arid urban environment.
Environ. Sci. Technol. 22:811-818. ICA no. A-855.

17. Andersson, L., Sulkowski, E., and Porath, J. 1991. Immobilized metal ion affinity
chromatography of serum albumins. Bioseparation 2:15-22. ICA no. A-1468.

18. Angelidis, M.O. 1995. The impact of urban effluents on the coastal marine environment of
Mediterranean islands. Water Sci. Technol. 32 (9-10):85-94. ICA no. A-2229.

19. Anuradha, C.D., and Devi, C.S.S. 1995. Studies on the hematological profile and trace
elements in oral submucous fibrosis. J. Clin. Biochem. Nutr. 19 (1):9-17. ICA no. A-2384.

20. Araya, M., Mcgoldrick, M.C., Klevay, L.M., Strain, J.J., Robson, P., Nielsen, F.H., Olivares,
M., Pizarro, F., Baker, S.R., and Poirier, K.A. 2001. Determination of an Acute No-Observed-
Adverse-Effect-Level (NOAEL) for Copper in Water. Regulatory Toxicology and Pharmacology
34:137-145.

21. ATSDR. 1990. Copper: Public Health Statement. Atlanta, GA, USA: Div. Toxicol., U. S.
Agency Toxic Subst. Dis. Reg. ICA no. U-460.

22. Aulenbach, D.B., Meyer, M.A., Beckwith, E., Joshi, S., Vasudevan, C., and Clesceri, N.L.
1987. Removal of heavy metals in POTW. Environ. Prog. 6 (2):91-98. ICA no. A-1922B.

23. Azcue, J.M., Rosa, F., and Mudroch, A. 1996. Distribution of major and trace elements in
sediments and pore water of Lake Erie. J. Great Lake Research 22 (2):389-402. ICA no. A-2225.

24. Badri, M.A., and Aston, S.R. 1985. Heavy metal occurrence and geochemical fractionation:
The relationships of catchment soils to associated estuarine sediments. Environ. Pollut. Ser. B
(10):61-75. ICA no. B-978.

25. Barrie, L.A., Lindberg, S.E., Chan, W.H., Ross, H.B., Arimoto, R., and Church, T.M. 1987.
On the concentration of trace metals in precipitation. Atmos. Environ. 21:1133-1135. ICA no. B-
1449.

26. Barries, D.G., and Dourson, M.L. 1988. Reference dose (RtD): description and use in health
risk assessments. Toxicol. Pharmacol. 8:471-486.

27. Beaumont, C., Simon, M., Fauchet, R., Hesppel, J.P., Brissot, P., Genetet, B., and Bourel, M.
1979. Serum ferritin as a possible market of the hemochromatosis allele. N. Engl. J. Med.
301:169-174.

28. Becking, G. 1996. Copper: essentiality and toxicity. IPCS News 10:4-5.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 174
REFERENCES

29. Belay, N., and Daniels, L. 1990. Elemental metals as electron sources for biological methane
formation from carbon dioxide. Antonie Van Leeuwenhoek 57:1-7. ICA no. B-1984.

30. Berg, R. 1987. Information and privately-owned faucets for the prevention of copper
contamination in drinking water. Lakartidningen 84:273-274. ICA no. B-2403.

31. Bertazzo, A., Costa, C., Biasiolo, M., Allegri, G., Cirrincione, G., and Presti, G. 1996.
Determination of copper and zinc levels in human hair: Influence of sex, age, and hair
pigmentation. Biol. Trace Element Res. 52 (1):37-53. ICA no. B-3718.

32. Beveridge, S.J., Boettcher, B., Walker, W.R., and Whitehouse, M.W. 1984. Biodistribution
of 64Cu in rats after topical application of two lipophilic anti-inflammatory Cu(II) formulations.
Agents Actions 14:291-295. ICA no. B-700.

33. Bewers, J.M., and Yeats, P.A. 1977. Oceanic residence times of trace metals. Nature
268:595-598. ICA no. B-279.

34. Bijlsma, M., Galione, A.L.S., Kelderman, P., Alaerts, G.J., and Clarisse, I.A. 1996.
Assessment of heavy metal pollution in inner-city canal sediments. Water Sci. Technol. 33
(6):231-237. ICA no. B-3777.

35. Blagoi, Y.P., Kornilova, S.V., and Sokhan, V.I. 1982. Change in the intrinsic viscosity of
DNA in interaction with Cu2+ Mn2+ ions. Mol. Biol. 16:170-175. ICA no. B-394.

36. Blevins, R.D., and Pancorbo, O.C. 1986. Metal concentrations in muscle of fish from aquatic
systems in East Tennessee, U. S. A. Water Air Soil Pollut. 29:361-371. ICA no. B-3213B.

37. Bloomfield, J., Dixon, S., and McCredie, D.A. 1971. Potential hepatotoxicity of copper in
recurrent hemodialvsis. Arch Int Med 128:555-560.

38. Bogden, J.D., and Klevay, L.M. 2000. Clinical Nutrition of the Essential Trace Elements and
Minerals. Totawa, NJ: Humana Press.

39. Boon, P.I. 1994. Discrimination of algal and bacterial alkaline phosphatases with a
differential-inhibition technique. Aust J Mar Freshwater Res 45 (1):83-107. ICA no. B-3189.

40. Bothwell, T.H., Chariton, R.W., Cook, J.D., and Finch, C.A. 1979. Iron Metabolism In Man.
Oxford: Blackwell.

41. Bowen, H.J.M., ed. 1966. Trace Elements in Biochemistry. London, England: Academic
Press. ICA no. B-040.

42. Bowman, B.A., and Rishert, J.F. 1994. Comparison of the methodological approaches used
in the derivation of recommended dietary allowances and oral reference doses for nutritionally
essential elements. In Risk Assessment Of Essential Elements, edited by Mertz, W., Abernathy,
C. O. and Olin, S. S. Washington, DC: ILSI Press.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 175
REFERENCES

43. Boyle, E.A. 1979. Copper in natural waters. In Copper in the Environment, edited by Nriagu,
J. O. New York, NY, USA: John Wiley & Sons. ICA no. B-180.

44. Bradley, R.W., and Morris, J.R. 1986. Heavy metals in fish from a series of metal-
contaminated lakes near Sudbury, Ontario. Water Air Soil Pollut. 27:341-354. ICA no. B-1190.

45. Branting, C., Oden, A., Wiatr-Adamczak, E., Linder, L.E., and Sund, M.-L. 1988. Adhesion
of streptococcus rattus and streptococcus mutans to metal surfaces. Scand. J. Dent. Res. 96:218-
225. ICA no. B-1479.

46. Bremer, P.J., and Geesey, G.G. 1990. Involvement of biofilm microorganisms in copper
corrosion. In Copper '90: Refining, Fabrication, Markets, edited by Institute of Metals. London,
England: The Institute of Metals. ICA no. B-2098.

47. Brown, D.A., Gossett, R.W., Hershelman, P., Schaefer, H.A., Jenkins, K.D., and Perkins,
E.M. 1983a. Bioaccumulation and detoxification of contaminants in marine organisms from
southern California coastal waters. In Waste Disposal in the Oceans: Minimizing Impact,
Maximizing Benefits, edited by Soule, D. F. and Walsh, D. Boulder, CO, USA: Westview Press.
ICA no. B-712.

48. Brown, K.W., Thomas, J.C., and Slowey, J.F. 1983b. The movement of metals applied to
soils in sewage effluent. Water Air Soil Pollut 19:43-54. ICA no. B-585.

49. Buckley, D.E., Smith, J.N., and Winters, G.V. 1995. Accumulation of contaminant metals in
marine sediments of Halifax Harbour, Nova Scotia: Environmental factors and historical trends.
Appl. Geochem. 10 (2):175-195. ICA no. B-3983.

50. Buckley, W.T., Vanderpool, R.A., Godfrey, D.V., and Johnson, P.E. 1996. Determination,
stable isotope enrichment and kinetics of direct-reacting copper in blood plasma. J. Nutr.
Biochem. 7 (9):488-494. ICA no. B-3893.

51. Budd, J., Kerfoot, W.C., Pilant, A., and Jipping, L.M. 1999. The Keweenaw current and ice
rafting: Use of satellite imagery to investigate copper-rich particle dispersal. J. Great Lakes
Research 25 (4):642-662.

52. Bugenyi, F.W.B., and Lutalo Bosa, A.J. 1990. Likely effects of salinity on acute copper
toxicity to the fisheries of the Lake George-Edward basin. Hydrobiologia 208:39-44. ICA no. B-
2259.

53. Burhanoglu, M.S., Tutuncuoglu, S., Coker, C., Tekgul, H., and Ozgur, T. 1996.
Hypozincaemia in febrile convulsion. Eur. J. Pediatr. 155 (6):498-501. ICA no. B-3828.

54. Bush, J.A., Mahoney, J.P., Markowitz, H., Gubler, C.J., Cartwright, G.E., and Wintrobe,
M.M. 1955. Studies on copper metabolism XVI Radioactive copper studies in normal subjects
and in patients with hepatolenticular degeneration. J Clin Invest 34:1766-1778. ICA no. B-
3214B.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 176
REFERENCES

55. Cairns, J., Jr. 1990a. Global climate/ecosystem monitoring: Inland aquatics. Spec. Sci.
Technol. 13:71-80. ICA no. C-1465.

56. Cairns, J., Jr. 1990b. The genesis of biomonitoring in aquatic ecosystems. Environ. Prof.
12:169-176. ICA no. C-1466.

57. Calabrese, E.J., and Moore, G.S. 1979. Can elevated levels of copper in drinking water
precipitate acute hemolysis in G-6-PD deficient individuals? Med. Hypotheses 5:493-498. ICA
no. C-2310B.

58. Callahan, M.A., Slimak, H.W., Gabel, N.W., May, I., Fowler, C., Freed, R., Jennings, P., and
et al. 1979. Water-Related Environmental Fate of 129 Priority Pollutants. In Introduction and
Technical Background, Metals and Inorganics, Pesticides and PCBs, Final Report. Washington,
DC, USA: Off Water Plan Stand U.S. Environmental Protection Agency. ICA no. C-2444B.

59. Campbell, P.G.C., Lewis, A.G., Chapman, P.M., Crowder, A.A., Fletcher, W.K., Imber, B.,
Luoma, S.N., Stokes, P.M., and Winfrey, M. 1988. Biologically Available Metals in Sediments.
Ottawa, Ontario, Canada: NRCC/CNRC Publ. ICA no. C-984.

60. Cartwright, G.E., and Wintrobe, M.M. 1964. Copper metabolism in normal subjects. Amer J
Clin Nutr 14:224-232. ICA no. C-2311B.

61. Chan, W.H., Tang, A.J.S., Chung, D.H.S., and Lusis, M.A. 1986. Concentration and
deposition of trace metals in Ontario - 1982. Water Air Soil Pollut 29:373-389. ICA no. C-851.

62. Cheevaporn, V., Jacinto, G.S., and San Diego-McGlone, M.L. 1995. Heavy metal fluxes in
Bang Pakong River Estuary, Thailand: Sedimentary vs. diffusive fluxes. Mar. Pollut. Bull. 31 (4-
12):290-294. ICA no. C-2790.

63. Chugh, K.S., and Sakhuja, V. 1979. Acute copper intoxication. Int. J. Artificial Organs 2
(4):181-182. ICA no. C-2446B.

64. Chuttani, H.K., Gupta, P.S., Gulati, S., and Gupta, D.N. 1965. Acute copper sulfate
poisoning. Amer. J. Med. 39:849-854. ICA no. C-2316B.

65. Clegg, S.L., and Sarmiento, J.L. 1989. The hydrolytic scavenging of metal ions by marine
particulate matter. Prog Oceanogr 23:1-21. ICA no. C-1252.

66. Coale, K.H., and Bruland, K.W. 1988. Copper complexation in the Northeast Pacific. Limnol
Oceanogr 33:1084-1101. ICA no. C-1021.

67. Cohen, S.R. 1979. Environmental and Occupational Exposure to Copper. In Copper in the
Environment - Part II: Health Effects, edited by Nriagu, J. O. New York, NY, USA: John Wiley
& Sons.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 177
REFERENCES

68. Cole, R.H., Frederick, R.E., Healy, R.P., and Rolan, R.G. 1984. Preliminary findings of the
priority pollutant monitoring project of the nationwide urban runoff program. J. Water Pollut.
Control Fed. 56 (7):898-908. ICA no. C-2451B.

69. Cornelis, R., Sabbioni, E., and Van der Venne, M.T. 1994. Trace element reference values in
tissues from inhabitants of the European Community. VII. Review of trace elements in blood,
serum and urine of the Belgian population and critical evaluation of their possible use as
reference values. Sci. Total Environ. (158):191-226. ICA no. C-2892.

70. Cousins, R.J. 1996. Zinc. In Present Knowledge In Nutrition, edited by Ziegler, E. E. and
Filer Jr., L. J. Washington, DC: ILSI Press.

71. Dabek, J.T., Hyvonen-Debek, M., Kupila-Rantala, T., Harkonen, M., and Adlercreutz, H.
1996. Early breast cancer, diet and plasma copper fractions. Ann. Clin. Lab. Sci. 26 (3):215-226.
ICA no. D-2148.

72. Davidson, C.I., Goold, W.D., Mathison, T.P., Wiersma, G.B., Brown, K.W., and Reilly, M.T.
1985. Airborne trace elements in Great Smoky Mountains, Olympic and Glacier National Parks.
Environ. Sci. Technol. 19:27-34. ICA no. D-353.

73. Davies, D.J.A., and Bennett, B.G. 1985. Exposure of man to environmental copper: An
exposure contaminant assessment. Sci. Total Environ. 46:215-228. ICA no. D-493.

74. Davies, N.T., and Campbell, J.K. 1977. The effect of cadmium on intestinal copper
absorption and binding in the rat. Life Sci. 20:955-960. ICA no. D-1715B.

75. Davies-Colley, R.J., Nelson, P.O., and Williamson, K.J. 1984. Copper and cadmium uptake
by estuarine sedimentary phases. Environ. Sci. Technol. 18:491-499. ICA no. D-331.

76. Davies-Colley, R.J., Nelson, P.O., and Williamson, K.J. 1985. Sulfide control of cadmium
and copper concentrations in anaerobic estuarine sediments. Mar. Chem. 16:173-186. ICA no. D-
440.

77. Davis, G.K., and Mertz, W. 1987. Copper. In Trace Elements in Human and Animal
Nutrition, edited by Mertz, W. San Diego, CA, USA: Acad Press, Inc. ICA no. D-775.

78. Davis, W.L., Goodman, D.B.P., Kipnis, M., and Matthews, J.L. 1988a. Immunolocalization
of copper-zinc superoxide dismutase in rat growth plate cartilage. Anat Rec 220:28A. ICA no. D-
784.

79. Davis, W.L., Goodman, D.B.P., Shibata, K., and Kipnis, M. 1988b. The immunolocalization
of copper-zinc superoxide dismutase (SOD) in the epithelial cells of the toad urinary bladder.
Anat Rec 29A. ICA no. D-783.

80. Dincer, N., and Akar, N. 1995. Maternal hair zinc, copper and magnesium concentration in
neural tube defects in Turkey. Trace Elem. Med. 12 (4):184-185. ICA no. D-2129.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 178
REFERENCES

81. DiToro, D., Allen, H.E., Bergman, H.L., Meyer, J.S., and Santore, R.C. 2000. The Biotic
Ligand Model, Copper in the Environment and Health. New York, NY: International Copper
Association.

82. Domek, M.J., Robbins, J.E., Anderson, M.E., and McFeters, G.A. 1987. Metabolism of
escherichia coli injured by copper. Can. J. Microbiol. 33:57-62. ICA no. D-656.

83. Donovan, U.M., and Gibson, R.S. 1996. Dietary intakes of adolescent females consuming
vegetarian, semi-vegetarian, and omnivorous diets. J. Adolescent Health 18 (4):292-300. ICA no.
D-2085.

84. Dourson, M.L. 1994. Methods for establishing oral reference doses. In Risk Assessment Of
Essential Elements, edited by Mertz, W., Abernathy, C. O. and Olin, S. S. Washington, DC: ILSI
Press.

85. Dourson, M.L., and Stara, J.F. 1983. Regulatory history and experimental support of
uncertainty (safety) factors. Regul Toxicol Pharmacol 3:224-238.

86. Dressler, R.L., Strom, G.L., Tzilkowski, W.M., and Sopper, W.E. 1986. Heavy metals in
cottontail rabbits on mined lands treated with sewage sludge. J Environ Qual 15:278-281. ICA
no. D-564.

87. Eckel, W.P., and Jacob, T.A. 1988. Ambient levels of 24 dissolved metals in U.S. surface
and ground waters. Abstr. Pap. Amer. Chem. Soc. 196 (ENVR 152). ICA no. E-446.

88. Eckel, W.P., and Langley, W.D. 1988. A background-based ranking technique for
assessment of elemental enrichment in soils at hazardous waste sites. Paper read at Superfund '88
Proceedings of the 9th National Conference, 1988, at Silver Spring, MD, USA. ICA no. E-799B.

89. Ecological Planning and Toxicology Inc. 1998. Literature Summary: Effects of Copper in
Terrestrial Systems. Corvallis, Oregon.

90. Edwards, M., Meyer, T.E., and Schock, M.R. 1996. Alkalinity, pH and Copper Corrosion
By-Product Release. Journal of the AWWA 88 (3):81.

91. Effler, S.W., Litten, S., Field, S.D., Tong-Ngork, T., Hale, F., Meyer, M., and Quirk, M.
1980. Whole lake responses to low level copper sulfate treatment. Water Res 14:1489-1499. ICA
no. E-080.

92. Eife, R., Bender-Goetze, C., Mueller-Hoecker, J., Wiebecke, B., Thanner, F., Schneider, K.,
Schmoelz, A., Arleth, S., Kellner, M., and et al. 1987. Impaired natural killer cell (NK) activity
and interferon (IFN) deficiency, features of Indian childhood cirrhosis, and hepatic copper
overload induced by chronic copper poisoning (via tap water) in a lower Bavarian family. Eur. J.
Pediatr. 146:331. ICA no. E-343.

93. Eisenberg, M., and Topping, J.J. 1986. Trace metal residues in finfish from Maryland waters,
1978-1979. J. Environ. Sci. Health 21 (B):87-102. ICA no. E-306.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 179
REFERENCES

94. Eisler, R. 2000. Copper. In Handbook of Chemical Risk Assessment: Health Hazards to
Humans, Plants and Animals. Boca Raton: Lewis Publishers.

95. Elliott, H.A., Liberati, M.R., and Huang, C.P. 1986. Competitive adsorption of heavy metals
by soils. J Environ Qual 15:214-219. ICA no. E-309.

96. Elliott, J.E., Scheuhammer, A.M., Leighton, F.A., and Pearce, P.A. 1992. Heavy metal and
metallothionein concentrations in Atlantic Canadian seabirds. Arch Environ Contam Toxicol
22:63-73. ICA no. E-623.

97. Emenaker, N.J., Disilvestro, R.A., Nahman, N.S., and Percival, S. 1996. Copper related
blood indexes in kidney dialysis patients. Am. J. Clinic. Nutr. 64 (5):757-760. ICA no. E-943.

98. Ergin, M., Kazan, B., and Ediger, V. 1996. Source and depositional controls on heavy metal
distribution in marine sediments of the Gulf of Iskenderun, eastern Mediterranean. Mar. Geol.
133 (3-4):223-239. ICA no. E-898.

99. Evans, G.W., and LeBlanc, F.N. 1976. Copper-binding protein in rat intestine: Amino acid
composition and function. Nutr Rept Int 14 (3):281-288. ICA no. E-800B.

100. Evans, G.W., Majors, P.F., and Cornatzer, W.E. 1970a. Ascorbic acid interaction with
metallothionein. Biochem Biophys Res Commun 41:1244-1247. ICA no. E-772B.

101. Evans, G.W., Majors, P.F., and Cornatzer, W.E. 1970b. Induction of ceruloplasmin
synthesis by copper. Biochem Biophys Res Commun 41:1120-1125. ICA no. E-773B.

102. Evans, J., Harris, O., and van Deth, A.G. 1984. Congenital hepatic fibrosis associated with
mallory bodies and copper retention. Aust. N.Z. J. Med. 14:500-503. ICA no. E-244.

103. Farrell, P.M., Bieri, J.G., Fratatoni, J.F., Wood, R.E., and Sant'Agnese, P.A.D. 1977. The
occurence and effects of human vitamin E deficiency. J. Clin. Invest. 60:233-241.

104. Farrer, P., and Mistilis, S.P. 1967. Absorption of exogenous and indogenous biliary copper
in the rat. Nature 213 (Jan 21):291-292. ICA no. F-1401B.

105. Fatoki, O.S. 1996. Trace zinc and copper concentration in roadside surface soils and
vegetation: Measurement of local atmospheric pollution in Alice, South Africa. Environ. Int. 22
(6):759-762. ICA no. F-1631.

106. Ferguson, J.L., von Franqu, O., and Schock, M.R. 1996. Internal Corrosion of Water
Distribution Systems. In Corrosion of Copper in Potable Water Systems. Denver, Colorado:
AWWA Research Foundation/DVGW-TZW.

107. Fimmel, S., Borchelt, A., Kage, A., and Kottgen, E. 1994. Trace elements and carrier
proteins in the aged. Arch. Gerontol Geriatr. (Suppl. 4):67-74. ICA no. F-1638.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 180
REFERENCES

108. Fisher, A.A. 1973. Contact Dermatitis, 2nd edition. Philadelphia, PA, USA: Lea and
Febiger.

109. Fleet, J.C., and Mayer, J. 1996. Discovery of the hemochromatosis gene will require
rethinking the regulation of iron metabolism. Nutr Rev 54:258-292.

110. Frommer, D.J. 1976. Direct measurement of serum noncaeruloplasmin copper in liver
disease. Clin. Chim. Acta 68:303-307.

111. Fuhrer, G.J. 1986. Extractable Cadmium, Mercury, Copper, Lead, and Zinc in the Lower
Columbia River Estuary, Oregon. Portland, OR, USA: U S Geol Surv, Prepared in Cooperation
with the U S Army Corps of Engineers Water-Resources Investigations Report. ICA no. F-712.

112. Garban, B., Ollivon, D., Carru, A.M., and Chesterikoff, A. 1996. Origin, retention and
release of trace metals from sediments of the River Seine. Water Air Soil Pollut. 87 (1-4):363-
381. ICA no. G-2318.

113. Gardner, M.J., and Ravenscroft, J.E. 1991. The range of copper-complexing ligands in the
Tweed estuary. Chem Speciation Bioavail 3:22-29. ICA no. G-1556.

114. Gaylarde, C.C. 1989. Microbial corrosion of metals. Environ. Eng. 2:30-32. ICA no. G-
1257.

115. Georgopoulos, P.G., Roy, A., Lioy, M.J., Opiekun, R.E., and Lioy, P.J. 2001.
Environmental copper: Its dynamics and human exposure issues. Journal of Toxicology and
Environmental Health Part B, 4:341-394.

116. Georgopoulos, P.G., Tan, H.C., Wang, S.W., Vyas, V.M., Georgopoulos, I.G., Yang, Y.C.,
and Lioy, P.J. 2002. A Framework and Data Sources for the Assessment of Exposures to Copper:
Technical Report Prepared for the International Copper Association (Draft) (available at
http://www.CERM.org/copper).

117. Gerritse, R.G., and van Driel, W. 1984. The relationship between adsorption of trace metals,
organic matter and pH in temperate soils. J Environ Qual 13:197-204. ICA no. G-400.

118. Gianotto, A.K., Wichlacz, P.L., Jolley, J.G., Hankins, M.R., and Geesey, G.G. 1989.
Biocorrosion of copper by biopolymers as examined in situ, in real time FT-IR-CIR-ATR in
conjunction with pre and post XPS-AES (Photoelectron Spectroscopy-Auger Electron
Spectroscopy). Paper read at Annual Meeting and Exhibit of the Society of Mining Engineers,
1989, at Washington, DC, USA. ICA no. G-1178.

119. Gibbs, R.J. 1977. Transport Phases of Transition Metals in the Amazon and Yukon Rivers.
Geol. Soc. Am. Bull. 88:829-843.

120. Giesy, J.P., Alberts, J.J., and Evans, D.W. 1986. Conditional stability constants and binding
capacities for copper(II) by dissolved organic carbon isolated from surface waters of the
Southeastern United States. Environ Toxicol Chem 5:139-154. ICA no. G-779.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 181
REFERENCES

121. Giesy, J.P., Newell, A., and Leversee, G.J. 1983. Copper speciation in soft, acid, humic
waters: Effects on copper bioaccumulaiton by and toxicity to Simocephalus serrulatus
(Daphnidae). Sci Total Environ 28:23-36. ICA no. G-300.

122. Gimeno-Garcia, E., Andreu, V., and Boluda, R. 1996. Heavy metals incidence in the
application of inorganic fertilizers and pesticides to rice farming soils. Environ. Pollut. 92 (1):19-
25. ICA no. G-2432.

123. Goldberg, E.D. 1965. Minor elements in seawater. In Chemical Oceanography, edited by
Riley, J. P. and Skirrow, G. New York, NY, USA: Acad Press. ICA no. G-024.

124. Goldberg, E.D. 1986. The Mussel Watch concept. Environ Monit Assess 7:91-103. ICA no.
G-2049B.

125. Granato, T.C., Kristoff, L., Pietz, R.I., Leu-Hing, C., and Cothern, C.R. 1994. Changes in
trace metals and radionuclides in Illinois soils since 1935. Paper read at Trace Substances,
Environment and Health: Annual Conference of the Society of Environmental Geochemistry and
Health, 1994, at Northwood, England, UK. ICA no. G-2509.

126. Gregus, Z., and Klaassen, C.D. 1986. Disposition of metals in rats: A comparative study of
fecal, urinary and biliary excretion and tissue distribution of eighteen metals. Toxicol Appl
Pharmacol 85:24-38. ICA no. G-659.

127. Gupta, G., and Karuppiah, M. 1996. Heavy metals in sediments of two Chesapeake Bay
tributaries - Wicomico and Pocomoke Rivers. J. Hazard. Mater. 50 (1a):15-29. ICA no. G-2383.

128. Haines, R.C. 1984. Environmental contamination: Surveys of heavy metals in urban soils
and hazard assessment. Trace Subst Environ; Health 18:450-460. ICA no. H-2299B.

129. Hall, A.C., Young, B.W., and Bremner, I. 1979. Intestinal metallothionein and the mutual
antagonism between copper and zinc in the rat. J Inorg Biochem 11:57-66. ICA no. H-2300B.

130. Hamm, A., Glassmann, M., and Liepelt, A. 1996. Transport of particulate matter in an
alpine river (River Salzach) and its importance for river ecology. Ergeb. Limnol. (47):507-513.
ICA no. H-2863.

131. Harrison, B.J. 1998. Copper Information Sourcebook. Vancouver, BC, Canada: University
of British Columbia.

132. Harrison, F.L. 1984a. Chemicals in Effluent Waters from Nuclear Power Stations: The
Distribution, Fate and Effects of Copper. Livermore, CA, USA: Lawrence Livermore Nat Lab.
ICA no. H-563.

133. Harrison, F.L. 1984b. Review of the Impact of Copper Released into Freshwater
Environments. Washington, D.C., USA: Nuclear Regul Comm. ICA no. H-451.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 182
REFERENCES

134. Harrison, F.L. 1985. Effect of physicochemical form on copper availability to aquatic
organisms (Special Technical Publication). Paper read at Aquatic Toxicology and Hazard
Assessment: Seventh Symposium ASTM, 1985, at Philadelphia, PA, USA. ICA no. H-613.

135. Harrison, F.L., Bishop, D.J., Emerson, R.R., and Rice, D.W., Jr,. 1980. Concentration and
Speciation of Copper in Waters Collected Near the San Onofre and Diablo Canyon Nuclear
Power Stations. Washington, D.C., USA: U.S. Nuclear Regul Comm. ICA no. H-235M.

136. Harrison, F.L., and Lam, J.R. 1982. Concentrations of Copper-Binding Proteins in Livers of
Bluegills from the Cooling Lake at the H. Washington, D.C., USA: B Robinson Nuclear Power
Station Nuclear Regul Comm. ICA no. H-245.

137. Hasan, M.Z. 1995. Trace element intake through food and water. Paper read at Trace and
th
Toxic Elements in Nutrition and Health: Proceedings of the 4 International Confernece on
Health and Diseases: Effects of Essential and Toxic Trace Elements, 1995, at New Delhi, India.
ICA no. H-2798.

138. Haywood, S., and Comerford, B. 1980. The effect of excess dietary copper on plasma
enzyme activity and on the copper content of the blood of the male rat. J Comp Pathol 90:233-
238. ICA no. H-2389B.

139. He, M., Want, A., and Tang, H. 1997. Spatial and temporal patterns of acidity and heavy
metals in predicting the potential for ecological impact on the Le An river polluted by acid mine
drainage. Sci. Total Environ. 206:67-77.

140. Heit, M., and Klusek, C.S. 1985. Trace element concentrations in the dorsal muscle of white
suckers and brown bullheads from two acidic adirondack lakes. Water Air Soil Pollut 25:87-96.
ICA no. H-634.

141. Helz, G.R., Huggett, R.J., and Hill, J.M. 1975. Behavior of Mn, Fe, Cu, Zn, Cd, and Pb
discharged from a wastewater treatment plant into an estuarine environment. Water Res 9:631-
636. ICA no. H-064.

142. Henriksen, A., and Wright, R.F. 1978. Concentrations of heavy metals in small Norwegian
lakes. Water Res 12:101-112. ICA no. H-040.

143. Heresi, G., Castillo-Duran, C., Munoz, C., Arevalo, M., and Schlesinger, L. 1985.
Phagocytosis and immunoglobulin levels in hypocupremic infants. Nutr Res (New York) 5:1327-
1334. ICA no. H-774.

144. Hermann, R., and Neumann-Mahlkau, P. 1985. The mobility of zinc, cadmium, copper,
lead, iron and arsenic in ground water as a function of redox potential and pH. Sci Total Environ
48:1-12. ICA no. H-603.

145. Hernandez, L.M., Gonzalez, M.J., Rico, M.C., Fernandez, M.A., and Baluja, G. 1985.
Presence and biomagnification of organochlorine pollutants and heavy metals in mammals of

COPPER: Environmental Dynamics and Human Exposure Issues


Page 183
REFERENCES

Donana National Park (Spain), 1982-1983. J Environ Sci Health 20 (B):633-650. ICA no. H-
2390B.

146. Herr, C., and Gray, N.F. 1996. Seasonal variation of metal contamination of riverine
sediments below a copper and sulphur mine in South-East Ireland. Water Sci. Technol. 33
(6):255-261. ICA no. H-2712.

147. Higuchi, S., Higashi, A., Nakamura, T., and Matsuda, I. 1988. Nutritional copper deficiency
in severely handicapped patients on a low copper enteral diet for a prolonged period: Estimation
of the required dose of dietary copper. J Pediatr Gastroenterol Nutr 7:583-587. ICA no. H-1054.

148. Hlavay, J., Polyak, K., Bodog, I., Molnar, A., and Meszaros, E. 1996. Distribution of trace
elements in filter-collected aerosol samples. Fresenius J. Anal. Chem. 354 (2):227-232. ICA no.
H-2870.

149. Holleran, R.S. 1981. Copper sulfate overdose. J. Emergency Nursing 7:136-137. ICA no. H-
2392B.

150. Hong, S.-M., Candelone, J.P., Patterson, C.C., and Boutron, C.F. 1996a. History of ancient
copper smelting pollution during Roman and medieval times recorded in Greenland ice. Science
272 (a):246-249. ICA no. H-2727.

151. Hong, S.-M., Candelone, J.P., Soutif, M., and Boutron, C.F. 1996b. A reconstruction of
changes in copper production and copper emissions to the atmosphere during the past 7000
years. Sci. Total Environ. 188 (2-3):183-193. ICA no. H-2828.

152. Horslen, S.P., Tanner, M.S., Lyon, T.D.B., Fell, G.S., and Lowry, M.F. 1994. Copper
associated childhood cirrhosis. Gut 35 (10):1497-1500. ICA no. H-2296.

153. Howell, J.M., and Gawthorne, J.M. 1987. Copper in Animals and Man. Boca Raton, FL:
CRC Press.

154. Impellitteri, C.A., Allen, H.E., Lagos, G., and McLaughlin, M.J. 2000. Removal of Soluble
Cu and Pb by the Automatic Drip Coffee Brewing Process: Application of Risk Assessment.
Human and Ecological Risk Assessment 6 (2):313-322.

155. Isbir, T., Tamer, L., Erkisi, M., Kekec, Y., Doran, F., Varinli, S., and Taylor, A. 1995.
Copper, zinc and magnesium in serum and tissues from patients with carcinoma of breast,
stomach and colon. Trace Elem. Electrolytes 12 (3):113-115. ICA no. I-559.

156. Ismail, S.S. 1996. Distribution of trace elements in Egyptian ground and Nile water. J.
Trace Microprobe Tech. 14 (1):243-253. ICA no. I-588.

157. Iyengar, V., and Woittiez, J. 1988. Trace elements in human clinical specimens: Evaluation
of literature data to identify reference values. Clin. Chem. 34:474-481. ICA no. I-186.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 184
REFERENCES

158. Jana, S., and Bhattacharya, D.N. 1988. Effect of heavy metals on growth population of a
fecal coliform bacterium Escherichia coli in aquatic environment. Water Air Soil Pollut. 38:251-
254. ICA no. J-407.

159. Jeong, J., Urban, N.R., and Green, S. 2000. Release of Copper from Mine Tailings on the
Keweenaw Peninsula. J. Great Lakes Research 25 (4):721-734.

160. Johns, C.E. 1993. An assessment of contamination of riparian wetlands by metals from past
mining and smelting activities in the headwaters region of the Clark Fork River, MT, U.S.A.
Paper read at Heavy Metals in the Environment, International Conference, 1993, at Edinburgh,
UK. ICA no. J-1065.

161. Johnson, C.A., Sigg, L., and Zobrist, J. 1987. Case studies on the chemical composition of
fogwater: The influence of local gaseous emissions. Atmos Environ 21:2365-2374. ICA no. J-
925B.

162. Joint Group of Experts on the Scientific Aspects of Marine Pollution (GESAMP). 1989.
The Atmospheric Input of Trace Species to the World Ocean. In Reports and Studies of the
GESAMP. Geneva, Switzerland: World Meteorol Org. ICA no. J-537.

163. Joseph, G. 1999. Copper: Its Trade, Manufacture, Use, and Environmental Status. Edited
by Kundig, K. J. A. Materials Park, OH: ASTM International.

164. Ke, D.-S., and Jiang, Y.-J. 1995. Contents and distribution of heavy metals in the offshore
area along Guangdong Province. Collected Oceanic Works 18 (1):61-69. ICA no. K-2982.

165. Keller, C., and Domergue, F.L. 1996. Soluble and particulate transfers of Cu, Cd, Al, Fe
and some major elements in gravitational waters of a Podzol. Geoderma 71 (3-4):263-274. ICA
no. K-2933.

166. Kelly, J., Thornton, I., and Simpson, P.R. 1996. Urban geochemistry: A study of the
influence of anthropogenic activity on the heavy metal content of soils in traditionally industrial
and nonindustrial areas of Britain. Appl. Geochem. 11 (1-2):363-370. ICA no. K-2983.

167. Kennish, M.S. 1997. Practical Handbook of Estuarine and Marine Pollution. Boca Raton,
FL: CRC Press.

168. Kerfoot, W.C., Harting, S., Rossmann, R., and Robbins, J.A. 1999. Anthropogenic copper
inventories and mercury profiles from Lake Superior: Evidence for mining impacts. J. Great
Lakes Research 25 (4):663-682.

169. Kerfoot, W.C., and Nriagu, J.O. 1999. Copper mining, copper cycling and mercury in the
Lake Superior ecosystem: An introduction. J. Great Lakes Research 25 (4):594-598.

170. Kerfoot, W.C., and Robbins, J.A. 1999. Nearshore regions of Lake Superior: Multi-element
signatures of mining discharges and a test of Pb-210 deposition under conditions of variable
sediment mass flux. J. Great Lakes Research 25 (4):697-720.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 185
REFERENCES

171. Kester, D.R., O'Connor, T.P., and Bynne, R.H. 1975. Solution chemistry, solubility, and
adsorption equilibria of iron, cobalt, and copper in marine systems. Thalassia Jugoslavica
11:121-134. ICA no. K-059M.

172. Kies, C., ed. 1989. Copper Bioavailability and Metabolism. Vol. 258, Advances in
Experimental Medicine and Biology. New York, NY, USA: Plenum Press. ICA no. K-1396.

173. Kimball, K.D. 1973. Seasonal fluctuations of ionic copper in Knights Pond, Massachusetts.
Limnol. Oceanogr. 18:169-172. ICA no. K-2510B.

174. King, L.D. 1988. Retention of metals by several soils of the Southeastern United States. J
Environ Qual 17:239-246. ICA no. K-1134.

175. Knobeloch, L., Ziarnik, M., Howard, J., Theis, B., Farmer, D., Anderson, H., and Proctor,
M. 1994. Gastrointestinal upsets associated with ingestion of copper contaminated water.
Environ. Health Perspect. 102 (11):958-961. ICA no. K-2730.

176. Kolak, J.J., Long, D.T., Kerfoot, W.C., Beals, T.M., and Eisenreich, S.J. 1999. Nearshore
versus offshore copper loading in Lake Superior sediments: Implications for transport and
cycling. J. Great Lakes Research 25 (4):611-624.

177. Krajcovicova Kudlackova, M., Simoncic, R., Babinska, K., Bederova, A., Brtkova, A.,
Magalova, T., and Grancicova, E. 1995. Selected vitamins and trace elements in blood of
vegetarians. Ann. Nutr. Metab. 39 (6):334-339. ICA no. K-2954.

178. Lacatusu, R., Rauta, C., Carstea, S., and Ghelase, I. 1996. Soil-plant-man relationships in
heavy metal polluted areas in Romania. Appl. Geochem. 11 (1-2):105-107. ICA no. L-2556.

179. Lagos, G. 2001. Corrosion of Copper Plumbing Tubes and the Release of Copper By-
Products to Drinking Water - A Literature Summary, Copper in the Environment and Health.
New York, NY: International Copper Association.

180. Lagos, G.E., Maggi, L.C., Peters, D., and Revco, F. 1999. Model for estimation of human
exposure to copper in drinking water. Sci. Total Environ. 239:49-70.

181. Landeen, L.K., Yahya, M.T., and Gerba, C.P. 1989a. Disinfection of Legionella
pneumophila using copper and silver ions and free chlorine. Abstr. Annu. Meet. Amer. Soc.
Microbiol. 89:292. ICA no. L-1047.

182. Landeen, L.K., Yahya, M.T., and Gerba, C.P. 1989b. Efficacy of copper and silver ions and
reduced levels of free chlorine in inactivation of Legionella pneumophila. Appl Environ
Microbiol 55:3045-3050. ICA no. L-064.

183. Landner, L., and Lindestrom, L. 1999. Copper in Society and in the Environment: An
Account of the Facts on Fluxes, Amounts and Effects of Copper in Sweden. Second Revised ed.
Sweden: Swedish Environmental Research Group.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 186
REFERENCES

184. Landner, L., Walterson, E., and Hellstrand, S. 2000. Copper in Sewage Sludge and Soil: A
Literature Review and Critical Discussion of Disposal of Copper-Containing Sludges to
Agricultural Land, Copper in the Environment and Health. New York, NY: International Copper
Association.

185. Larsen, P.F., Zdanowicz, V., and Johnson, A.C. 1983. Trace metal distributions in the
surficial sediments of Penobscot Bay, Maine. Bull. Environ. Contam. Toxicol. 31:566-573. ICA
no. L-325.

186. Laryea, M.D., Schnittert, B., Kerstring, K., Wilhelm, M., and Lombeck, I. 1995.
Macronutrient, copper, and zinc intakes of young German children as determined by duplicate
food samples and diet records. Ann. Nutr. Metabol. 39 (5):271-278. ICA no. L-2531.

187. Lebedev, V.S., Deinega, E.Y., Savluk, O.S., and Fedorov, Y.I. 1989. The role of
transmembrane potential in Cu2+-induced breakdown of escherichia coli cytoplasmic membrane.
Biol. Memb. 6:1313-1316. ICA no. L-1325.

188. Lechevallier, M.W., Lowry, C.D., and Lee, R.G. 1990. Disinfecting biofilms in a model
distribution system. Amer. Water Works Assoc. J. 82:87-99. ICA no. L-1281.

189. Lee, C.-L., Chen, H.Y., and Chuang, M.Y. 1996. Use of oyster, Crassostrea gigas, and
ambient water to assess metal pollution status of the charting coastal area, Taiwan, after the 1986
green oyster incident. Chemosphere 33 (12a):2505-2532. ICA no. L-2579.

190. Lefkowitch, J.H., Honig, C.L., King, M.E., and Hagstrom, J.W.C. 1982. Hepatic copper
overload and features of Indian childhood cirrhosis in an American sibship. New England J.
Med. 307:271-277. ICA no. L-355.

191. Leonard, E.N., Ankley, G.T., and Hoke, R.A. 1996. Evaluation of metals in marine and
freshwater surficial, sediments from the environmental monitoring and assessment program
relative to proposed sediment quality criteria for metals. Environ. Toxicol. Chem. 15 (12):2221-
2232. ICA no. L-2551.

192. Leone, A., and Mercer, J.F.B., eds. 1999. Copper Transport and its Disorders: Molecular
and Cellular Aspects. Vol. 448. New York, NY: Kluwer Academic/Plenum Publishers.

193. Lewis, A.G. 1995a. The Biological Importance of Copper: A Literature Review (Final
Report). New York, NY: International Copper Association. ICA no. L-2064.

194. Lewis, A.G. 1995b. Copper in Water and Aquatic Environments. New York, NY:
International Copper Association. ICA no. L-2553.

195. Linder, M.C. 1991a. Biochemistry of Copper. Vol. 10, Biochemistry of the Elements. New
York, NY, U.S.A.: Plenum Press. ICA no. L-1634.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 187
REFERENCES

196. Linder, M.C. 1991b. Nutrition and metabolism of the trace elements. In Nutritional
Biochemistry and Metabolism: With Clinical Applications, edited by Linder, M. C. Amsterdam,
The Netherlands: Elsevier Sci. Publ. B.V. ICA no. L-1716.

197. Linder, M.C., and Hazegh-Azam, M. 1996. Copper biochemistry and molecular biology.
Amer. J. Clin. Nutr. 63 (5):797S-811S. ICA no. L-2609.

198. Line, D.E., Arnold, J.A., Jennings, G.D., and Wu, J. 1996. Water quality of stormwater
runoff from ten industrial sites. Water Resour. Bull. 32 (4):807-816. ICA no. L-2657.

199. Lioy, P.J., Daisey, J.M., Morandi, M.T., Harkov, R.D., Greenberg, A., Bozzelli, J.,
Kebbekus, B., Louis, J., and McGeorge, L.J. 1987. The airborne toxic element and organic
substances (ATEOS) study design. In Toxic Air Pollution: A Comprehensive Study of Non-
Criteria Air Pollutants, edited by Lioy, P. J. and Daisey, J. M. Chelsea, MI, USA: Lewis Publ.
Inc. ICA no. L-2168B.

200. Lodenius, M., and Braunschweiler, H. 1986. Volatilization of heavy metals from a refuse
dump. Sci Total Environ 57:253-255. ICA no. L-742.

201. Long, D.T., and Angino, E.E. 1977. Chemical speciation of cadmium, copper, lead, and
zinc in mixed fresh water sea water and brine solutions. Geochim. Cosmochim. Acta 41:1183-
1191. ICA no. L-038.

202. Lopez de Sa, A. 1989. Some considerations about the influence of materials used for water
pipes on water quality. Alimentaria 26:65-67. ICA no. L-1329.

203. Lowe, T.P., May, T.W., Brumbaugh, W.G., and Kane, D.A. 1985. National contaminant
biomonitoring program: Concentrations of seven elements in freshwater fish, 1978-1981. Arch.
Environ. Contam. Toxicol. 14:363-388. ICA no. L-164.

204. Lower, M.W. 1987. Comprehensive Cooling Water Study: Volume 3; Radionuclide and
Heavy Metal Transport, Savannah River Plant: Final Report. Washington, DC, USA: U.S. Dept
Energy. ICA no. L-986.

205. Luoma, S.N. 1983. Bioavailability of trace metals to aquatic organisms. A review Sci Total
Environ 28:1-22. ICA no. L-285.

206. Lytle, D.A., and Schock, M.R. 1996. Stagnation Time, Composition, pH and
Orthophosphate Effects on Metal Leaching from Brass. Washington, DC: Office of Research and
Development.

207. Lytle, D.A., and Schock, M.R. 1997. An Investigation of the Impact of Alloy Composition
and pH on the Corrosion of Brass in Drinking Water. Advances in Environmental Research 1
(2):213-233.

208. Lytle, D.A., and Schock, M.R. 2000. Impact of Stagnation Time on Metal Dissolution from
Plumbing Materials. J. Water SRT-Aqua 49 (5):243.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 188
REFERENCES

209. Machelett, B., Grun, M., and Bergmann, H. 1993. Heavy metal content of arable soils in
East Germany. Paper read at Heavy Metals in the Environment. International Conference, 1993,
at Edinburgh, UK. ICA no. M-3867.

210. Madany, I.M., Akhter, M.S., and Raveendran, E. 1996. Spatial and temporal patterns in
heavy metals concentration in street dust in a hot desert region. Int. J. Environ. Health Res. 6
(2):93-101. ICA no. M-3825.

211. Madsen, H., Poulsen, L., and Grandjean, P. 1990. High copper per content in drinking water
and the risks involved. Ugeskr. Lg. 152:1806-1809. ICA no. M-1994.

212. Maessen, O., Freedman, B., and McCurdy, R. 1985. Metal mobilization in home well water
systems in Nova Scotia. J. Amer. Water Works Assoc. 77:73-80. ICA no. M-892.

213. Man, A.C.K., Zheng, Y.H., and Mak, P.K. 1996. Trace elements in scalp hair of
professional drivers and university teachers in Hong Kong. Biol. Trace Elem. Res. 53 (1-3):241-
247. ICA no. M-3725.

214. Mansilla-Rivera, I., and Nriagu, J.O. 1999. Copper chemistry in freshwater ecosystems: an
overview. J. Great Lakes Research 25 (4):599-610.

215. Marceau, N., Aspin, N., and Sass-Kortsak, A. 1970. Absorption of copper 64 from
gastrointestinal tract of the rat. Amer J Physiol 218 (2):377-383. ICA no. M-3364B.

216. Markus, J.A., and McBratney, A.B. 1996. An urban soil study: Heavy metals in Glebe,
Australia. Aust. J. Soil Res. 34 (3):453-465. ICA no. M-3903.

217. Mart, L., Nurnberg, H.W., and Dyrssen, D. 1984. Trace metal levels in the eastern Arctic
Ocean. Sci. Total Environ. 39:1-14. ICA no. M-754.

218. Martin, J.-M., and Windom, H.L. 1991. Present and future roles of ocean margins in
regulating marine biogeochemical cycles of trace elements. In Ocean Margin Processes in
Global Change, Dahlem Workshop, edited by Mantoura, R. F. C., Martin, J.-M. and Wollast, R.
Chichester, NY, USA: John Wiley & Sons. ICA no. M-2277.

219. Martin, M., and Castle, W. 1984. Petrowatch: Petroleum hydrocarbons, synthetic organic
compounds and heavy metals in mussels from the Monterey Bay area of Central California. Mar
Pollut Bull 15:259-266. ICA no. M-3365B.

220. Martin, M., Stephenson, M.D., and Martin, J.H. 1977. Copper toxicity experiments in
relation to abalone deaths observed in a power plant's cooling waters. Calif Fish Game 63:95-
100. ICA no. M-088.

221. Mason, K.E. 1979. A conspectus of research on copper metabolism and requirements of
man. J. Nutr. 109 (11):1979-2066. ICA no. M-3367B.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 189
REFERENCES

222. Mateu, J., Forteza, R., Cerda, V., and Colom Altes, M. 1996. Particle size distribution and
long-range transport of metals in atmospheric aerosols from the Alfabia station (Majorca, Spain).
J. Environ. Sci. Health 31A (1):31-54. ICA no. M-3912.

223. McFeters, G.A., and Singh, A. 1990. Effects of aquatic environmental stress upon enteric
bacterial pathogens. J. Appl. Bacteriol. 69:VII-VIII. ICA no. M-2297.

224. McIlroy, L.M., Depinto, J.V., Young, T.C., and Martin, S.C. 1986. Partitioning of heavy
metals to suspended solids of the Flint River, Michigan. Environ Toxicol Chem 5:609-624. ICA
no. M-1061.

225. McManus, J.P., and Prandle, D. 1996. Determination of source concentrations of dissolved
and praticulate trace metals in the southern North Sea. Mar. Pollut. Bull. 32 (6):504-512. ICA
no. M-3755.

226. Mehra, R.K., and Bremner, I. 1984. Species differences in the occurrence of copper-
methallothionein in the particulate fractions of the liver of copper-loaded animals. Biochem J
219:539-546. ICA no. M-672.

227. Meranger, J.C., Subramanian, K.S., and Chalifoux, C. 1979. A national survey for
cadmium, chromium, copper, lead, zinc, calcium and magnesium in Canadian drinking water
supplies. Environ. Sci. Technol. 13:707-711. ICA no. M-3368B.

228. Mercer, J.F.B., Lazdins, I., Stevenson, T., Camakaris, J., and Danks, D.M. 1981. Copper
induction of translatable metallothionein messenger RNA. Biosci Rept 1:793-800. ICA no. M-
3369B.

229. Mermut, A.R., Jain, J.C., Song, L., Kerrick, R., Kozak, L., and Jana, S. 1996. Trace element
concentrations of selected soils and fertilizers in Saskatchewan, Canada. J. Environ. Qual. 25
(4):845-853. ICA no. M-3923.

230. Mesuere, K., and Fish, W. 1989. Behavior of runoff-derived metals in a detention pond
system. Water Air Soil Pollut 47:125-138. ICA no. M-2028.

231. Meyer, J.S., Santore, R.C., Bobbitt, J.P., DeBrey, L.D., Boese, C.J., Paquin, P.R., Allen,
H.E., Bergman, H.L., and DiToro, D.M. 1998. Binding of nickel and copper to fish gills predicts
toxicity when water hardness varies, but free ion activity does not. Environ Sci Technol 33:913-
916.

232. Mihelcic, G., Surija, B., Juracic, M., Barisic, D., and Branica, M. 1996. History of the
accumulation of trace metals in sediments of the saline Rogoznica Lake (Croatia). Sci. Total
Environ. 182 (1-3):105-115. ICA no. M-3929.

233. Miller, D.R., Byrd, J.E., and Perona, M.F. 1987. The source of Pb, Cu and Zn in fogwater.
Water Air Soil Pollut 32:329-340. ICA no. M-1285.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 190
REFERENCES

234. Miller, F.N., Lominadze, D.G., and Schuschke, D.A. 1995. Copper: A dietary nutrient
essential for normal microvascular function. Paper read at Microcirculatory Approach to Asian
Traditional Medicine: Strategy for the Scientific Evaluation. Satellite Symposium of the 2nd
Asian Congress for Microcirculation (ACM '95), August 17, 1995, at Beijing, China. ICA no. M-
3936.

235. Miller, M.W., Benjamin, M.B., Reiber, S.H., and Ferguson, J.F. 1988. Phosphate inhibition
of copper corrosion in potable water systems. Amer. Water Works Assoc. J. 80:84. ICA no. M-
1586.

236. Mills, G.L., and Quinn, J.G. 1984. Dissolved copper and copper-organic complexes in the
Narragansett Bay Estuary. Mar. Chem. 15:151-172. ICA no. M-602.

237. Millward, G.E., Allen, J.I., Morris, A.W., and Turner, A. 1996. Distribution and fluxes of
non-detrital particulate Fe, Mn, Cu, Zn in the Humber Coastal Zone, UK. Continental Shelf Res.
16 (7):967-993. ICA no. M-3757.

238. Milos, B., Knezic, S., Pavasovic, S., Odzak, N., Simac, L., and Orlov, M. 1993. The
distibution of heavy metals in soils from Adriatic coastal zone: Kastela Bay, Croatia. Paper read
at Heavy Metals in the Environment. International Conference, 1993, at Edinburgh, UK. ICA no.
M-3795.

239. Minear, R.A., Ball, R.O., and Church, R.L. 1981a. Data Base for Influent Heavy Metals in
Publicly Owned Treatment Works. Cincinnati, OH, USA: Municipal Environ Res Lab. ICA no.
M-301M.

240. Minear, R.A., Cantrell, M.W., Church, R.L., Hannah, S.A., and Ball, R.O. 1981b. Heavy
Metals in Municipal Wastewater Treatment Plant Influents: An Analysis of the Data Available
from Treatment Plants. Paper read at Conference on Combined Municipal/Industrial Wastewater
Treatment, 1981, at Cincinnati, OH, USA. ICA no. M-510M.

241. Moffett, J.W., and Zika, R.G. 1987a. Photochemistry of copper complexes in sea water.
Paper read at Photochemistry of Environmental Aquatic Systems, 1987, at Washington, DC,
USA. ICA no. M-1530.

242. Moffett, J.W., and Zika, R.G. 1987b. Reaction kinetics of hydrogen peroxide with copper
and iron in seawater. Environ Sci Technol 21:804-810. ICA no. M-1294.

243. Moffett, J.W., and Zika, R.G. 1987c. Solvent extraction of copper acetylacetonate in studies
of copper(II) speciation in seawater. Mar Chem 21:301-313. ICA no. M-1293.

244. Mohan, S.V., Nithila, P., and Reddy, S.J. 1996. Estimation of heavy metals in drinking
water and development of heavy metal pollution index. J. Environ. Sci. Health 31A (2):283-289.
ICA no. M-3951.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 191
REFERENCES

245. Morales, J.A., Pirela, D., and Duran, J. 1996. Determination of the levels of Na, K, Ca, Mg,
Fe, Zn and Cu in aerosols of the western Venezuelan savannah region. Sci. Total Environ. 180
(2):155-164. ICA no. M-3957.

246. Moynahan, E.J. 1974. Acrodermatitis enterophatica: a lethal inherited human zincdeficiency
disorder. Lancet:399-340.

247. Muller, F.L.L. 1996. Interactions of copper, lead and cadmium with the dissolved, collotidal
and particulate components of estuarine and coastal waters. Mar. Chem. 52 (3-4):245-268. ICA
no. M-3796.

248. Mller, T., Feichtinger, H., Berger, H., and Mller, W. 1996. Endemic Tyrolean infantile
cirrhosis: an egocentric disorder. Lancet 347 (9005):877-880.

249. Muller-Hocker, J., Meyer, U., Wiebecke, B., Hubner, G., Eife, R., Kellner, M., and
Schramel, P. 1988. Copper storage disease of the liver and chronic dietary copper intoxication in
two further German infants mimicking Indian childhood cirrhosis. Pathol. Res. Pract. 183:39-45.
ICA no. M-1380.

250. Multhaup, G.S., and Hermann, D. 2000. Copper in the Pathogenesis of Neurodegenerative
Disorders: A Literarature Summary, Copper in the Environment and Health. New York, NY:
International Copper Association.

251. Mumma, R.O., Raupach, D.C., Waldman, J.P., Tong, S.S.C., Jacobs, M.L., Babish, J.G.,
Hotchkiss, J.H., Wszolek, P.C., Gutenman, W.H., Bache, C.A., and Lisk, D.J. 1984. National
survey of elements and other constituents in municipal sewage sludges. Arch Environ Contam
Toxicol 13:75-83. ICA no. M-752.

252. Murray, K.S. 1996. Statistical comparisons of heavy-metal concentrations in river


sediments. Environ. Geol. 27 (1):54-58. ICA no. M-3845.

253. Mussalo-Rauhamaa, H., Kantola, M., Seppanen, K., Soininen, L., and Koivusalo, M. 1996.
Trends in the concentrations of mercury, copper, zinc and selenium in inhabitants of north-
eastern Finnish Lapland in 1982-1991. A pilot study. Arctic Med. Res. 55 (2):83-91. ICA no. M-
3976.

254. Mussalo-Rauhamaa, H., Salmela, S.S., Leppnen, A., and Pyysalo, H. 1986. Cigarettes as a
source of some trace and heavy metals and pesticides in man. Arch Environ, Health 4 (1):49-55.
ICA no. M-3374B.

255. National Academy of Sciences. 1975. Assessing Potential Ocean Pollutants. Washington
D.C., USA.

256. National Institute of Medicine. 2001. Dietary Reference Intakes for Vitamin A, Vitamin K,
Arsenic, Boron, Chromium, Copper, Iodine, Iron, Manganese, Molybdenum, Nickel, Silicon,
Vanadium, and Zinc (Chapter 7 - Copper). Washington D.C.: Institute of Medicine, Food and
Nutrition Board.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 192
REFERENCES

257. National Research Council. 1989a. Recommended Dietary Allowances (10th edition).
Washington, D.C., USA: National Academy Press.

258. National Research Council. 1989b. Biological Markers of Reproductive Toxicology.


Washington, DC: National Academy Press.

259. National Research Council. 2000. Copper in Drinking Water. Washington, D.C.: National
Academy Press.

260. National Technical Information Service. 1989. Wastewater Treatment. In June 1970 -
November 1987 (Citations from the Compendex Database). Springfield, VA, USA: National
Technical Information Service. ICA no. N-537.

261. Neal, C., Smith, C.J., Jeffery, H.A., Jarvie, H.P., and Robson, A.J. 1996. Trace element
concentrations in the major rivers entering the Humber estuary, NE England. J. Hydrol.
(Amsterdam) 182 (1-4):37-64. ICA no. N-1291.

262. Nelson, S.G., Jr. 1987. Ion-selective Water Filter. United States: Off. Gaz. U.S. Pat.
Trademark Off. Pat. ICA no. N-521.

263. Nielsen, F.H., and Milne, D.B. 1993. Effect of high dietary fructose on copper homeostasis
and status indicators in men during copper deprivation. Paper read at Trace Elements in Man and
Animals TEMA-8 Proceedings of the Eighth International Symposium, 1993, at Gersdorf,
Germany. ICA no. M-3144.

264. Nolte, J. 1988. Pollution source analysis of river water and sewage sludge. Environ.
Technol. Lett. 9:857-868. ICA no. N-520.

265. Nolting, R.F., Van Dalen, M., and Helder, W. 1996. Distribution of trace and major
elements in sediment and pore waters of the Lena delta and Laptev Sea. Mar. Chem. 53 (3-
4):285-299. ICA no. N-1319.

266. Nordberg, G.F. 1990. Human health effects of metals in drinking water: Relationship to
cultural acidification. Environ. Toxicol. Chem. 9:887-894. ICA no. N-701.

267. Nowak, B. 1993. Levels of heavy metals in the biological tests (hair, teeth) as an indicator
of the environment pollution. Paper read at Heavy Metals in the Environment. International
Conference, 1993, at Edinburgh, UK. ICA no. N-1315.

268. Nriagu, J.O. 1979c. The global copper cycle. In Copper in the Environment, Part I
Ecological Cycling, edited by Nriagu, J. O. New York, NY, USA: John Wiley & Sons. ICA no.
N-050.

269. Nriagu, J.O. 1979d. Copper in the atmosphere and precipitation. In Copper in the
Environment, Part I Ecological Cycling, edited by Nriagu, J. O. New York, NY, USA: John
Wiley & Sons. ICA no. N-050.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 193
REFERENCES

270. Nriagu, J.O. 1988. Trace Metals in Lakes. Paper read at Science of the Total Environment,
International Conference, 1989, at McMaster University, Hamilton, Ontario, Canada. ICA no. N-
680.

271. Nriagu, J.O. 1989. Human perturbation of the global cycles of the trace elements. Abstr Pap
Amer Chem Soc 198 (ENVR 45). ICA no. N-604.

272. Nriagu, J.O., ed. 1979a. Copper in the Environment, Part I: Ecological Cycling. New York,
NY, USA: John Wiley & Sons. ICA no. N-049.

273. Nriagu, J.O., ed. 1979b. Copper in the Environment, Part II: Health Effects. New York,
NY, USA: John Wiley & Sons. ICA no. N-045.

274. Nriagu, J.O., and Coker, R.D. 1980. Trace metals in humic acid and fulvic acid from Lake
Ontario, Canada sediments. Environ Sci Technol 14:443-446. ICA no. N-071.

275. Nriagu, J.O., Lawson, G., Wong, H.K.T., and Cheam, V. 1996. Dissolved trace metals in
Lakes Superior, Erie and Ontario. Environ. Sci. Technol. 30 (1):178-187. ICA no. N-1379.

276. Nriagu, J.O., and Lin, T.S. 1995. Trace metals in wild rice sold in the United States. Sci.
Total Environ. 172 (2-3):223-228. ICA no. N-1378.

277. Nriagu, J.O., and Pacyna, J.M. 1988. Quantitative assessment of worldwide contamination
of air, water and soils by trace metals. Nature (London) 333 (6169):134-139. ICA no. N-459.

278. Nuttall, J.L. 1990. The use of copper water systems in controlling bacterial growth
including Legionella pneumophila. In Copper '90: Refining, Fabrication, Markets, edited by
Institute of Metals. London, England: The Institute of Metals. ICA no. N-618.

279. O'Dell, B.L., and Sunde, R.A. 1997. Health of nutritionally essential mineral elements. In
Copper, edited by Harris, E. D. New York, NY: Marcel Dekker.

280. Odhiambo, B.K., Macdonald, R.W., O'Brien, M.C., Harper, J.R., and Yunker, M.B. 1996.
Transport and fate of mine tailings in a coastal fjord of British Columbia as inferred from the
sediment record. Sci. Total Environ. 191 (1-2):77-94. ICA no. O-894.

281. Oganesyan, R.O., and Babayan, G.G. 1987. Heavy metals in the surface waters of the Sevan
Lake Basin, Armenia. Biol Zh Arm 40:1004-1007. ICA no. O-300.

282. Olivares, M., and Uauy, R. 1996a. Copper as an essential nutrient. Amer. J. Clin. Nutr. 63
(5a):791S-796S. ICA no. O-904.

283. Olivares, M., and Uauy, R. 1996b. Limits of metabolic tolerance to copper and biological
basis for present recommendations and regulations. Amer. J. Clin. Nutr. 63 (b):846S-852S. ICA
no. O-903.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 194
REFERENCES

284. Olivares, M., Uauy, R., Icaza, G., and Gonzalez, M. 1999. Models to evaluate health risks
derived from copper exposure/intake in humans. In Copper Transport and its Disorders, edited
by Leone and Mercer. New York, NY: Kluwer Academic/Plenum Publishers.

285. Oliver, M.A., Webster, R., and McGrath, S.P. 1996. Disjunctive kriging for environmental
management. Environmetrics 7 (3):333-357. ICA no. O-905.

286. Onag, A., and Taneli, B. 1995. Trace elements in infants fed human milk or formula. Paper
read at Trace and Toxic Elements in Nutrition and Health: Proceedings of the Fourth
International Conference on Health and Diseases: Effects of Essential and Toxic Trace Elements.
ICA no. O-912.

287. Onder, E., Guler, M., Deger, O., Orem, A., and Tosun, M. 1996. Bloodplasma, erythrocyte
and leukocyte zinc, copper, and magnesium of patients with rheumatoid arthritis. Trace Elem.
Med. 13 (2):85-87. ICA no. O-929.

288. Owen, C.A. 1982a. Biochemical Aspects of Copper, Copper Proteins, Ceruloplasmin and
Copper; Protein Binding Copper in Biology and Medicine Series. Park Ridge, NJ, USA: Noyes
Publications. ICA no. O-146.

289. Owen, C.A. 1982b. Biological Aspects of Copper, Occurence, Assay and Interrelationships;
Copper in Biology and Medicine Series. Park Ridge, NJ, USA: Noyes Publications. ICA no. O-
148.

290. Page, G.W. 1981. Comparison of groundwater and surface contamination by toxic
substances. Environ. Sci. Technol. 15:1475-1481. ICA no. P-1838B.

291. Pais, I., and Benton Jones, J., Jr. 1997. The Handbook of Trace Elements. Boca Raton, FL:
St. Lucie Press.

292. Paramasivam, P., and Gopalswamy, A. 1994. Distribution of micronutirents in Lower


Bhavani Project Command Area soil profiles. Madras Agric. J. 81 (10):545-547. ICA no. P-
2175.

293. Parametrix Inc., and EPT. 2001. Acclimation and Adaptation of Terrestrial Organisms to
Metals in Soil. New York, NY: International Copper Association.

294. Patel, P., and Bhattacharya, P.K. 1995. Study of binary and ternary complexes of some
binucleating ligands as models of binuclear copper proteins. Indian J. Chem. 34A (3):196-200.
ICA no. P-2294A.

295. Patrick, F.M., and Loutit, M. 1976. Passage of metals in effluents, through bacteria to
higher organisms. Water Res 10:333-335. ICA no. P-101.

296. PEDCo Environmental Inc. 1980. Primary copper industry. In Industrial process profiles
for environmental use. Cincinnati, OH: Industrial Environmental Research Laboratory, U.S.
Environmental Protection Agency.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 195
REFERENCES

297. Pedersen, K., Brandstrom, L., and Olsson, A.K. 1986. Do some tubing materials favor
attachment and growth of bacteria in drinking water systems? Vatten 42:21-24. ICA no. P-749.

298. Pennington, J.A.T., and Schoen, S.A. 1996. Total diet study: Estimated dietary intakes of
nutritional elements, 1982-1991. Int. J. Vitamin Nutr. Res. 66 (4):350-362. ICA no. P-2147.

299. Pennington, J.A.T., Young, B.E., Wilson, D.B., Johnson, R.D., and Vanderveen, J.E. 1986.
Mineral content of foods and total diets: The selected minerals in foods survey, 1982 to 1984. J.
Amer. Diet Assoc. 86:876-891. ICA no. P-659.

300. Peplow, G., and Vernon, F. 1987. Trace metal fouling and cleaning of seawater R O
membranes. In Proceedings of the Third World Congress on Desalination on Water Reuse Vol 3,
edited by Balaban, M. Amsterdam, Netherlands: Elsevier Sci. Publ. B.V. ICA no. P-884.

301. Perrone, L., Moro, R., Caroli, M., DiToro, R., and Gialanella, G. 1996. Trace elements in
hair of healthy children smapled by age and sex. Biol. Trace Elem. Res. 51 (1):71-76. ICA no. P-
1976.

302. Perwak, J., Bysshe, S., Goyer, M., Nelken, L., and Scow, K. 1980. Exposure and Risk
Assessment for Copper. Washington, DC, USA: Office Water Regul. Stand., Environmental
Protection Agency. ICA no. P-477.

303. Pettersson, R., and Rasmussen, F. 1999. Daily Intake of Copper from Drinking Water
among Young Children in Sweden. Environmental Health Perspectives 107 (6):441-446.

304. Phelps, H.L. 1984. A Research Program in Determination of Heavy Metals in Sediments
and Benthic Species in Relation to Nuclear Power Plant Operation. In Final Technical Report
1974-1982. Washington, DC, USA: Univ District of Columbia. ICA no. P-398.

305. Phillips, C., and Hershelman, G.P. 1996. Recent temporal trends in sediment metal
concentrations near a large wastewater outfall off Orange County, California. Water Environ.
Resour. 68 (1):105-114. ICA no. P-2097.

306. Poulsen, O.M., Christensen, J.M., Sabbioni, E., and Van der Venne, M.T. 1994. Trace
element reference values in tissues from inhabitants of the European Community V. Review of
trace elements in blood, serum and urine and critical evaluation of reference values for the
Danish population. Sci. Total Environ. (141):197-215.

307. Poulton, D.J., Morris, W.A., and Coakley, J.P. 1996. Zonation of contaminated bottom
sediments in Hamilton Harbour as defined by statistical classification techniques. Water Qual.
Res. J. Can. 31 (3):505-528. ICA no. P-2069.

308. Poulton, D.J., Simpson, K.J., Barton, D.R., and Lum, K.R. 1988. Trace metals and benthic
invertebrates in sediments of nearshore Lake Ontario at Hamilton Harbour. J. Great Lakes
Research 14:52-65. ICA no. P-1044.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 196
REFERENCES

309. Pyle, B.H., Broadway, S.C., and McFeters, G.A. 1992. Efficacy of copper and silver ions
with iodine in the inactivation of Pseudomonas cepacia. J. Appl. Bacteriol. 72 (71-79). ICA no.
P-1398.

310. Que, H.S.S., Finelli, V.N., Fricke, F.L., and Wolnik, K.A. 1982. Metal content of stack
emissions, coal and fly ash from some eastern and western power plants in the USA as obtained
by ICP-AES. Int J Environ Anal Chem 13:1-18. ICA no. Q-005.

311. Quemerais, B., Lum, K.R., and Lemieux, C. 1996. Concentrations and transport of trace
metals in the St. Lawrence River. Aquat. Sci. 58 (1):52-68. ICA no. Q-116.

312. Qureshi, I.H., Waheed, S., Rehman, A., and Ahmad, S. 1995. Intake of trace elements
through major food articles. Paper read at Trace and Toxic Elements in Nutrition and Health.
Proceedings of the Fourth Annual Conference. ICA no. Q-127.

313. Ralph, A., and McArdle, H. 2001. Copper Metabolism and Requirements in the Pregnant
Mother, her Fetus, and Children -- A Critical Review. New York, NY: International Copper
Association.

314. Raspor, B., Nurnberg, H.W., Valenta, P., and Branica, M. 1984a. Studies in seawater and
lake water on interactions of trace metals with humic substances isolated from marine and
estuarine sediments I Characterization of humic substances. Mar Chem 15:217-230. ICA no. R-
148.

315. Raspor, B., Nurnberg, H.W., Valenta, P., and Branica, M. 1984b. Studies in seawater and
lake water on interactions of trace metals with humic substances isolated from marine and
estuarine sediments II Voltammetric investigations on trace metal complex formation in the
dissolved phase. Mar Chem 15:231-249. ICA no. R-149.

316. Reed, J.S., and Henningson, J.C. 1984. Acid precipitation and drinking water supplies. J.
Amer. Water Works Assoc. 76:60-65. ICA no. R-373.

317. Reiber, S.H. 1989. Copper plumbing surfaces: An electrochemical study. Amer. Water
Works Assoc. J. 81:114-122. ICA no. R-932.

318. Reisenhofer, E., Adami, G., and Favretto, A. 1996. Heavy metals and nutrients in coastal,
surface sea-waters (Gulf of Trieste, Northern Adriatic Sea): An environmental study by factor
analysis. Fresenius J. Anal. Chem. 354 (5-6):729-734. ICA no. R-2114.

319. Renwick, W.H., and Edenborn, H.M. 1983. Metal and bacterial contamination in New
Jersey estuarine sediments. Environ. Pollut. Ser. B (5):175-185. ICA no. R-238.

320. Reunanen, A., Knekt, P., Marniemi, J., Maki, J., Maatela, J., and Aromaa, A. 1996. Serum
calcium, magnesium, copper and zinc and risk of cardiovascular death. Eur. J. Clin. Nutr. 50
(7):431-437. ICA no. R-2072.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 197
REFERENCES

321. Richardson, H.W., ed. 1997. Handbook of Copper Compound and Applications. NY:
Marcel Dekker.

322. Ritchie, G.S.P., and Sposito, G. 1995. Speciation in soils. Edited by Ure, A. M. and
Davidson, C. M., Chemical Speciation in the Environment. Glasgow, Scotland, UK: Blackie &
Son Ltd. ICA no. R-2010.

323. Ritter, W.F., and Eastburn, R.P. 1978. Leaching of heavy metals from sewage sludge
through coastal plain soils. Commun Soil Sci Plant Anal 9:785-798. ICA no. R-064.

324. Riveros-Rosas, H., Pfeifer, G.D., Lynam, D.R., Pedroza, J.L., Julian-Sanchez, A., Canales,
O., and Garfias, J. 1997. Personal exposure to elements in Mexico City air. Sci. Total Environ.
198:79-96.

325. Robberecht, H.J., Cauwenbergh, R.v., Hendrix, P., and Delstra, H.A. 1996. Daily dietary
intake of calcium, magnesium, copper, manganese, choromium and selenium in Belgium, using
duplicate portion sampling. Paper read at Metal Ions in Biology and Medicine. Proceeding of the
Fourth National Symposium, May 19-22, at Barcelona, Spain. ICA no. R-2147.

326. Rodriguez Tuya, I., Pinilla Gil, E., Maynar-Marino, M., Garcia Monco Carra, R.M., and
Sanchez Misiego, A. 1996. Evaluation of the influence of physical activity on the plasma
concentrations of several trace metals. Eur. J. Appl. Physiol. Occup. Physiol. 73 (3-4):299-303.
ICA no. R-1989.

327. Rope, S.K., Arthur, W.J., III, Craig, T.H., and Craig, E.H. 1988. Nutrient and trace elements
in soil and desert vegetation of Southern Idaho. Environ Monit Assess 10:1-24. ICA no. R-801.

328. Rossotti, H. 1998. Diverse Atoms - Profiles of the chemical elements, Oxford Chemistry
Guides. Oxford, UK: Oxford University Press.

329. Ruhling, A. 1995. Atmospheric heavy metal deposition in Europe estimated by moss
analysis. Edited by Munawar, M., Hanninen, O., Roy, S., Munawar, N., Karenlampi, L. and
Brown, D., Ecovision World Monograph Series. Bioindicators of Environmental Health.
Amsterdam, The Netherlands: SPB Acad. Publ. bv. ICA no. R-2101.

330. Saad, M.A.H. 1987. Limnological studies on the Nozha Hydrodrome, Egypt, with special
reference to the problems of pollution. Sci Total Envir 67:195-214. ICA no. S-1693.

331. Sadiq, M., Zaidi, T.H., and Sheikheldin, S. 1995. Concentration of metals of health
significance in commonly consumed shrimps in the eastern province of Saudia Arabia. J.
Environ. Sci. Health A30 (1):15-30. ICA no. S-5265.

332. Saltzer, E.I., and Wilson, J.W. 1968. Allergic contact dermatitis due to copper. Arch. Derm.
(98):375-376. ICA no. S-4565B.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 198
REFERENCES

333. Samet, J. 2000. A Technical Guide for the Study of Acute Gastrointestinal Effects of Copper
in Drinking Water: Methods for Public Health Investigators, Copper in the Environment and
Health. New York, NY: International Copper Association.

334. Santschi, P.H., Nixon, S., Pilson, M., and Hunt, C. 1984. Accumulation of sediments, trace
metals (Pb, Cu) and total hydrocarbons in Narragansett Bay, Rhode Island. Estuar Coastal Shelf
Sci 19:427-449. ICA no. S-765.

335. Savvides, C., Papadopoulos, A., Haralambous, K.J., and Loizidou, M. 1995. Sea sediments
contaminated with heavy metals: Metal speciation and removal. Water Sci. Technol. 32 (9-
10):65-73. ICA no. S-5138.

336. Scharenberg, W., and Ebeling, E. 1996. Distribution of heavy metals in a woodland food
web. Bull. Environ. Contam. Toxicol. 56 (3):389-396. ICA no. S-5128.

337. Scheinberg, I.H. 1979. Human health effects of copper. In Copper in the Environment, Part
II: Health Effects, edited by Nriagu, J. O. New York, NY, USA: John Wiley & Sons.

338. Scheinberg, I.H., and Sternlieb, I. 1994. Is non-Indian childhood cirrhosis caused by excess
dietary copper? Lancet (North American Edition) 344 (8928):1002-1004. ICA no. S-4135.

339. Schintu, M., Sechi, N., Sarritzu, G., and Contu, A. 1989. Reservoir sediments as potential
source of heavy metals in drinking water (Sardinia, Italy). Water. Sci. Technol. 21:1891-1894.
ICA no. S-2798.

340. Schneider, P.M., and Davey, S.B. 1995. Sediment contaminants off the coast of Sydney,
Australia: A model for their distribution. Mar. Pollut. Bull. 31 (4-12):262-272. ICA no. S-5099.

341. Schock, M.R. 2001. Personal communication.

342. Schock, M.R., Edwards, M., Powers, K., Hidmi, L., and Lytle, D.A. 2000. The Chemistry
of New Copper Plumbing. Paper read at Proc. AWWA Water Quality Technology Conference,
November 5-9, at Salt Lake City, UT.

343. Schock, M.R., Lytle, D.A., and Clement, J.A. 1995. Effect of pH, DIC,

Orthophosphate and Sulfate on Drinking Water Cuprosolvency. Cincinnati,

OH: USEPA Office of Research and Development.

344. Schock, M.R., and Neff, C.H. 1988. Trace metal contamination from brass fittings. Amer.
Water Works Assoc. J. 80:47-56. ICA no. S-2316.

345. Schoenen, D., and Schlomer, G. 1989. Microbial contamination of water by materials of
pipes and hoses; Third communication: Reaction of Escherichia coli, Citrobacter freundii and
Klebsiella pneumoniae. Zentralb Hyg. Umweltmed. 188:475-480. ICA no. S-2612.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 199
REFERENCES

346. Schoenen, D., Schulze-Robbecke, R., and Schirdewahn, N. 1988. Microbial contamination
of water by materials of pipes and hoses; 2nd Communication: Growth of Legionella
pneumophila. Zentralbl Bakteriol. Mikrobiol. Hyg. Ser. B. 186:326-332. ICA no. S-1822.

347. Schoenen, D., and Wehse, A. 1988. Microbial colonization of water by the materials of
pipes and hoses; 1st Communication: Changes in colony counts. Zentralbl Bakteriol. Mikrobiol.
Hyg. Ser. B. 186:108-117. ICA no. S-1823.

348. Schroeder, H.A., Nason, A.P., Tipton, I.H., and Balassa, J.J. 1966. Essential trace metals in
man: Copper. J Chronic Dis 19:1007-1034. ICA no. S-4558B.

349. Schroeder, W.H., Dobson, M., Kane, D.M., and Johnson, N.D. 1987. Toxic trace elements
associated with airborne particulate matter: A review. JAPCA 37:1267-1285. ICA no. S-2211.

350. Schulze-Robbecke, R., Jung, K.D., Pullmann, H., and Hundgeburth, J. 1990. Control of
Legionella pneumophila in a hospital hot water system. Zentralb Hyg. Umweltmed. 190:84-100.
ICA no. S-2616.

351. Schwartz, J., and Weiss, S.T. 1990. Dietary factors and their relation to respiratory
symptoms; The Second National and Nutrition Examination Survey. Amer. J. Epidemiol. 132:67-
76. ICA no. S-2618.

352. Scoullos, M., Dassenakis, M., and Zeri, C. 1996. Trace metal behaviour during summer in a
stratified Mediterranean system: The Louros Estuary (Greece). Water Air Soil Pollut. 88 (3-
4):269-295. ICA no. S-5070.

353. Sebestova, E., Machovic, V., Pavlikova, H., Lelak, J., and Minarik, L. 1996. Environmental
impact of brown coal mining in sokolova basin with especially trace metal mobility. J. Environ.
Sci. Health 31A (10):2453-2463. ICA no. S-5385.

354. Seisuma, Z., Legzdina, M., and Kulikova, I. 1993. The impact of the Riga waste treatment
plant on the level of heavy metals in the "Boldaraja study area" ecosystem (Gulf of Riga). Latv.
Zinat. Akad. Vestis. -B-Dala, -Dabaszinatnes (10):60-65. ICA no. S-5171.

355. Semple, A.B., Parry, W.H., and Phillips, D.E. 1960. Acute copper poisoning: An outbreak
traced to contaminated water from a corroded geyser. Lancet 2:700-701. ICA no. S-4567B.

356. Sen, J., and Chaudhuri, A.B.D. 1996. Human hair lead and copper levels in three
occupationally unexposed population groups in Calcutta. Bull. Environ. Contam. Toxicol. 57
(2):321-326. ICA no. S-5127.

357. Sharma, V.K., and Millero, F.J. 1988. Effect of ionic interactions on the rates of oxidation
of Cu(I) with O2 in natural waters. Mar. Chem. (25):141-161. ICA no. S-2047.

358. Shaw, J.C.L. 1992. Copper deficiency in term and preterm infants. In Nutritional Anemias
Workshop, edited by Fomon, S. J. and Zlotkin, S. New York, NY, USA: Raven Press. ICA no.
S-4266.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 200
REFERENCES

359. Shen, Z.L., and Liu, M.X. 1995. The removal of Zn, Cd, Pb and Cu in the Changjiang
Estuary. Chin. J. Oceanol. Limnol. 13 (4):332-337. ICA no. S-5474.

360. Shils, M.E., Olson, S.A., Shike, M., and Ross, A.C., eds. 1999. Modern Nutrition in Health
and Disease, 9th Edition. Baltimore, MD: William & Wilkins.

361. Shimbo, S., Hayase, A., Murakami, M., Hatai, I., Higashikawa, K., Moon, C.S., Zhang,
Z.W., Watanabe, T., Iguchi, H., and Ikeda, M. 1996. Use of a food composition database to
estimate daily dietary intake of nutrient or trace elements in Japan, with reference to its
limitations. Food Additives Contam. 13 (7):775-786. ICA no. S-5476.

362. Shine, J.P., and Wallace, G.T. 1996. Flux of surface-active oranic complexes of copper to
the air sea interface in coastal marine waters. J. Geophys. Res. 101 (C5):12,017 - 12,026. ICA
no. S-5133.

363. Sholkovitz, E.R. 1980. Removal of 'soluble' iron in the Potomac River estuary: Comments.
Estuar Coastal Mar Sci 11:585-587. ICA no. S-328.

364. Shotyk, W., Cheburkin, A.K., and Andreyev, A.V. 1993. Evaluation of anthropogenic
enrichments of heavy metals in soils of Kiev, Ukraine. Paper read at Heavy Metals in the
Environment. International Conference, 1993, at Edingurgh, UK. ICA no. S-5145.

365. Shulkin, V.M., and Kavun, V.I. 1995. The use of marine bivalves in heavy metal
monitoring near Vladivostok, Russia. Mar. Pollut. Bull. 31 (4-12):330-333. ICA no. S-5101.

366. Silver, S., Schottel, J., and Weiss, A. 1975. Bacterial resistance to toxic metals determined
by extrachromosomal R factors. Paper read at Third International Biodegradation Symposium
Biodeterioration Society, 1975, at London, England. ICA no. S-117.

367. Simpson, K., Morris, E.T.D., and Cook, J.D. 1981. The inhibitory effect of bran in iron
absorption in man. Amer J Clin Nutr 34:1469-1478.

368. Singh, I. 1990. Significance of Building and Plumbing Specifics on Trace Metal
Concentrations in Drinking Water. M.A. Sc., Univ. British Columbia, Vancouver, BC, Canada.
ICA no. S-3319.

369. Singh, I., and Mavinic, D.S. 1990. Significance of building and plumbing specifics on trace
metal concentrations in drinking water. Paper read at Annual Conference and 1st Biennial
Environmental Speciality Conference, 1990, at Hamilton, ON, Canada. ICA no. S-3302.

370. Sloof, W., Cleven, R.F.M.J., Janus, J.A., and Ros, J.P.M. 1989. Integrated criteria
document; Copper. Bilthoven, The Netherlands: Rijksinstituut voor de Volksgezondheid en
Milieuhygiene. ICA no. S-2608.

371. Smith, C.H., and Bidlack, W.R. 1980. Interrelationship of dietary ascorbic acid and iron on
the tissue distribution of ascorbic acid, iron and copper in female guinea pigs. J Nutr 110:1398-
1408. ICA no. S-4571B.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 201
REFERENCES

372. Smith, G.C., Clegg, M.S., Keen, C.L., and Grivetti, L.E. 1996. Mineral values of selected
plant foods common to southern Burkina Faso and to Niamey, Niger, West Africa. Int. J. Food
Sci. 47 (1):41-53. ICA no. S-5075.

373. Sorg, T.J., and Bell, F.A., Jr., eds. 1986. Plumbing Materials and Drinking Water Quality.
Vol. 128, Pollution Technology Review. Park Ridge, NJ, USA: Noyes Publ. ICA no. S-3853.

374. Soudine, A. 1989. WMO efforts towards the assessment of atmospheric fluxes of heavy
metals to the marine environment. Paper read at Heavy Metals in the Environment, 7th
International Conference, 1989, at Edinburgh, U.K. ICA no. S-2653.

375. Spitalny, K.C., Brondum, J., Vogt, R.L., Sargent, H.E., and Kappel, S. 1984. Drinking-
water-induced copper intoxication in a Vermont family. Pediatrics 74:1103-1106. ICA no. S-
1123.

376. St-Cyr, L., and Crowder, A.A. 1987. Relation between Fe, Mn, Cu and Zn in root plaque
and leaves of Phragmites australis. In Heavy Metals in the Environment, edited by Lindberg, S.
E. and Hutchinson, T. C. Edinburgh, U.K.: CEP Consult Ltd. ICA no. S-2763.

377. Stekel, A., Olivares, M., Pizarro, F., Amar, M., Chadud, P., Cayazzo, M., Llaguno, S.,
Vega, V., and Hertrampf, E. 1985. The role of ascorbic acid in the bioavailabilitv of iron from
infant foods. Int J Vitamin Nutr Res 55 (Supp 27):167-175.

378. Stephenson, T., and Lester, J.N. 1987. Heavy metal behavior during the activated sludge
process II Insoluble metal removal mechanisms. Sci Total Environ 63:215-230. ICA no. S-1807.

379. Sternlieb, I. 1982. Pathobiology of metals. In The Liver: Biology and Pathobiology, edited
by Arias, I. M., Popper, H., Schachter, D. and Shafritz, D. A. New York, NY, USA: Raven Press.
ICA no. S-438.

380. Sternlieb, I., and Scheinberg, I.H. 1977. Human copper metabolism. In Medical and
Biologic Effects of Environmental Pollutants - Copper. Washington D.C., USA: National
Academy of Sciences.

381. Stiff, M.J. 1971a. The chemical states of copper in polluted water and a scheme of analysis
to differentiate them. Water Res 5:585-599. ICA no. S-280.

382. Stiff, M.J. 1971b. Copper bicarbonate equilibria in solutions of bicarbonate ion at
concentrations similar to those found in natural water. Water Res 5:171-176. ICA no. S-022.

383. Stokinger, H.E. 1981. Copper. In Patty's Industrial Hygiene and Toxicology, edited by
Clayton, G. D. and Clayton, E. New York, NY, USA: John Wiley & Sons.

384. Strain, W.H., Hershey, C.O., McInnes, S., Breslau, D., Hershey, L.A., McKinney, B.M.,
Varnes, A.W., and Khourey, C.J. 1984. Hazards to groundwater from acid rain. Trace Subst.
Environ. Health 18:178-184. ICA no. S-4573B.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 202
REFERENCES

385. Strickland, G.T., Beckner, W.M., and Leu, M.-L. 1972. Absorption of copper in
homozygotes and heterozygotes for Wilson's disease and controls: Isotope tracer studies with
67Cu and 64Cu. Clin Sci 43:617-625. ICA no. S-4575B.

386. Suciu, I., Prodan, L., Lazar, V., Ilea, E., Cocirla, A., Olinici, L., Paduratu, A., Zagreanu, O.,
Lengyel, P., Gyorffi, L., and Andru, D. 1981. Research on copper poisoning. Med. Lavoro
3:190-197. ICA no. S-4560B.

387. Symons, J.M. 1986. Summation: Plumbing materials and drinking water quality. In
Plumbing Materials and Drinking Water Quality, edited by Sorg, T. J. and Bell, F. A., Jr. Park
Ridge, NJ, USA: Noyes Publ. ICA no. S-3857.

388. Tankere, S.P.C., and Statham, P.J. 1996. Distribution of dissolved Cd, Cu, Ni and Zn in the
Adriatic Sea. Mar. Pollut. Bull. 32 (8-9):623-630. ICA no. T-1695.

389. Tariq, J., Ashraf, M., Jaffar, M., and Afzal, M. 1996. Pollution status of the Indus River,
Pakistan, through heavy metal and macronutrient contents of fish, sediment and water. Water
Res. 30 (6):1337-1344. ICA no. T-1714.

390. Tariq, M.A., Qamar-un-Nisa, and Fatima, A. 1995. Concentrations of Cu, Cd, Ni, and Pb in
the blood and tissues of cancerous persons in a Pakistani population. Sci. Total Environ. 175
(1b):43-48. ICA no. T-1693.

391. Taylor, G.J., and Crowder, A.A. 1983a. Accumulation of atmospherically deposited metals
in wetland soils of Sudbury, Ontario. Water Air Soil Pollut. 19:29-42. ICA no. T-213.

392. Taylor, G.J., and Crowder, A.A. 1983b. Uptake and accumulation of copper, nickel and iron
by Typha latifolia grown in solution culture. Amer J Bot 70 (5 Part 2):56. ICA no. T-279.

393. Taylor, G.J., and Crowder, A.A. 1983c. Uptake and accumulation of copper, nickel and iron
by Typha latifolia grown in solution culture. Can J Bot 61:1825-1830. ICA no. T-298.

394. Tkalin, A.V., Presley, B.J., and Boothe, P.N. 1996. Spatial and temporal variations of trace
metals in bottom sediments of Peter the Great Bay, the Sea of Japan. Environ. Pollut. 92 (1):73-
78. ICA no. T-1761.

395. Tobias, F.J., Bech, J., and Sanchez Algarra, P. 1997. Establishment of the background
levels of some trace elements in soils of NE Spain with probability plots. Sci. Total Environ.
206:255-265.

396. Truitt, R.E., and Weber, J.H. 1981a. Copper(II) and cadmium(II) binding abilities of some
New Hampshire freshwaters determined by dialysis titration. Environ Sci Technol 15:1204-1208.
ICA no. T-120.

397. Truitt, R.E., and Weber, J.H. 1981b. Determination of complexing capacity of fulvic acid
for copper (II) and cadmium (II) by dialysis titration. Anal Chem 53:337-342. ICA no. T-119.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 203
REFERENCES

398. Turner, A.P., Watmough, S.A., and Dickinson, N.M. 1993. Ecotoxicology of 60 years of
copper, cadmium and zinc processing in north-west England. Paper read at Heavy Metals in the
Environment, International Conference, 1993, at Edinburgh, UK. ICA no. T-1717.

399. Turnlund, J.R. 1999. Copper. In Modern Nutrition in Health and Disease, 9th Edition,
edited by Shils, M. E., Olson, S. A., Shike, M. and Ross, A. C. Baltimore, MD: William &
Wilkins.

400. Turnlund, J.R., Keyes, W.R., Anderson, H.L., and Acord, L.L. 1989. Copper absorption and
retention in young men at three levels of dietary copper by use of the stable isotope 65Cu. Amer
J Clin Nutr 49:870-878. ICA no. T-797.

401. Tuschewitzki, G.J. 1990. Induction of microbial surface colonization in copper pipes by
drinking water. Zentralbl. Hyg. Umweltmed. 190:62-71. ICA no. T-875.

402. Tyler, L.D., and McBride, M.B. 1982. Mobility and extractability of cadmium, copper,
nickel, and zinc in organic and mineral soil columns. Soil Sci. 134 (3):198-205. ICA no. T-
1508B.

403. Uauy, R., Castillo-Duran, C., Fisberg, M., Fernandez, N., and Valenzuela, A. 1985. Red cell
superoxide dismutase activity as an index of human copper nutrition. J Nutr 115:1650-1655. ICA
no. U-062.

404. Unlu, E., Akba, O., Sevim, S., and Gumgum, B. 1996. Heavy metal levels in mullet, Liza
abu (Heckel, 1843) (Mugilidae) from the Tigris River, Turkey. Fresenius Environ. Bull. 5 (1-
2):107-112. ICA no. U-458.

405. USEPA. 1987. Superfund Record of Decision (EPA Region 2): Volney Landfill Site,
Volney, New York (First Remedial Action), July 1987. Washington, DC: United States
Environmental Protection Agency. ICA no. U-125.

406. van Campen, D.R., and Scaifi, P.U. 1967. Zinc interference with copper absorption in rats. J
Nutr 91:473-476.

407. Van den Berg, C.M.G. 1992. Effect of the deposition potential on the voltammetric
determination of complexing ligand concentrations in sea-water. Analyst 117:589-593. ICA no.
V-843.

408. Van den Berg, C.M.G., Nimmo, M., Daly, P., and Turner, D.R. 1990. Effects of detection
window on the determination of organic copper speciation in estuarine waters. Anal. Chim. Acta
232:149-159. ICA no. V-549.

409. Van Hattum, B., van Straalen, N.M., and Govers, H.A.J. 1996. Trace metals in populations
of freshwater isopods: Influence of biotic and abiotic variables. Arch. Environ. Contamin.
Toxicol. 31 (3):303-318. ICA no. V-1121.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 204
REFERENCES

410. Vaughan, L.A., Weber, C.W., and Kemberling, S.R. 1979. Longitudinal changes in the
mineral content of human milk. Amer. J. Clin. Nutr. 32:2301-2306. ICA no. V-999B.

411. Versteegh, J.F.M., Havelaar, A.H., Hoekstra, A.C., and Visser, A. 1989. Complexing of
copper in drinking water samples to enhance recovery of Aeromonas and other bacteria. J. Appl.
Bacteriol. 67:561-566. ICA no. V-589.

412. Viraraghavan, T., Subramanian, K.S., and Rao, B.V. 1996. Drinking water at the tap:
Impact of plumbing material on water quality. J. Environ. Sci. Health 31A (8):2005-2016. ICA
no. V-1182.

413. Wagner, P., Little, B., and Janus, L. 1987. An investigation of microbiologically mediated
corrosion of copper-nickel piping systems selectively treated with ferrous sulfate. Paper read at
Oceans '87 The Ocean - An International Workplace Proceedings: Marine Engineering Policy,
Education and Technology Transfer, 1987, at Washington, DC, USA. ICA no. W-974.

414. Wake, S.A., and Mercer, J.F.B. 1985. Induction of metallothionein mRNA in rat liver and
kidney after copper chloride injection. Biochem J 228:425-432. ICA no. W-501.

415. Walker, W.R., Reeves, R.R., Brosnan, M., and Coleman, G.D. 1977. Perfusion of intact
skin by a saline solution of bis(glycinato) copper(II). Bioinorg Chem 7:271-276. ICA no. W-
1678B.

416. Weant, G.E. 1985. Sources of Copper Air Emissions. Research Triangle Park, NC, USA:
Air Energy Engn. Res. Lab., Off. Res. Dev., Environmental Protection Agency. ICA no. W-
1674B.

417. Weber, P.M., O'Reilly, S., Pollycove, M., and Shipley, L. 1969. Gastrointestinal absorption
of copper: Studies with 64Cu, 95Zr, a whole- body counter and the scintillation camera. J
Nuclear Med 10:591-596. ICA no. W-1680B.

418. Weiss, H., Bertine, K., Koide, M., and Goldberg, E.E. 1975. Chemical Composition of
Greenland Glacier. Geochim. Cosmochim. Acta 39:1-10.

419. Weiss, K.C., and Linder, M.C. 1985. Copper transport in rats involving a new plasma
protein. Amer J Physiol 249 (1 Part 1):E77-E88. ICA no. W-492.

420. Wendt, P.H., van Dolah, R.F., Bobo, M.Y., Mathews, T.D., and Levisen, M.V. 1996. Wood
preservative leachates from docks in an estuarine environment. Arch. Environ. Contam. Toxicol.
31 (1):24-37. ICA no. W-1907.

421. Wenzel, C., and Gabrielsen, G.W. 1995. Trace element accumulation in three seabird
species from Hornoya, Norway. Arch. Environ. Contam. Toxicol. 29 (2):198-206. ICA no. W-
1919.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 205
REFERENCES

422. Wenzel, W.W., Alge, G., Blum, W.E.H., Brandstetter, A., Pollak, M.A., Riedler, C., and
Schulte, A. 1996. Seasonal and spatial variation of extractable trace metal fractions in topsoils
under mixed forest. Z. Pflanzenernaehrung Bodenkd. 159 (4):333-336. ICA no. W-1969.

423. WHO-IPCS. 1998. Copper - Environmental Health Criteria 200: World Health Organization
and the International Programme on Chemical Safety.

424. Widerlund, A. 1996. Early diagenetic remobilization of copper in near-shore marine


sediments: A quantitative pore-water model. Mar. Chem. 54 (1-2):41-53. ICA no. W-1904.

425. Williams, D.M. 1982. Clinical significance of copper deficiency and toxicity in the world
population. In Clinical, Biochemical, and Nutritional Aspects of Trace Elements, edited by
Prasad, A. S. New York, NY, USA: Alan R Liss, Inc. ICA no. W-293.

426. Williams, M.L., Short, R.J., O'Neal, P.L., and Oski, F.A. 1975. Role of dietary iron and fat
on vitamin E deficiency anemia of infancy. N Engl J Med 292:897-890.

427. Williams, S.C., Simpson, H.J., Olsen, C.R., and Bopp, R.F. 1978. Sources of heavy metals
in sediments of the Hudson River estuary. Mar Chem 6:195-214. ICA no. W-061.

428. Wong, J.W.C. 1996. Heavy metal contents in vegetables and market garden soils in Hong
Kong. Environ. Technol. 17 (4):407-414. ICA no. W-1901.

429. World Health Organization. 1993. Our Planet, Our Health: Report of the WHO Commission
on Health and Environment. Geneva, Switzerland: World Health Organization. ICA no. W-1339.

430. World Health Organization. 1996. Trace Elements in Human Nutrition and Health. Geneva,
Switzerland: World Health Organization. ICA no. W-2084.

431. Wyllie, J. 1957. Copper poisoning at a cocktail party. Amer. J. Public Health 47:617. ICA
no. W-1683B.

432. Xue, H.-B., Gachter, R., and Sigg, L. 1997. Comparison of Cu and Zn cycling in eutrophic
lakes with oxic and anoxic hypolimnion. Aquat. Sci. 59:176-189. ICA no. X-110.

433. Yahya, M.T., Landeen, L.K., Thurman, R.B., and Gerba, C.P. 1989. Inactivation of
bacteriophage MS-2 using copper and silver ions with low levels of free chlorine. Paper read at
Abstr. Annu. Meet. Amer. Soc. Microbiol., 1989. ICA no. Y-222.

434. Yahya, M.T., and Straub, T.M. 1990. Inactivation of bacteriophage MS-2 and poliovirus in
metal and plastic domestic water pipes. Paper read at Abstr. Annu. Meet. Amer. Soc. Microbiol.,
1990. ICA no. Y-268.

435. Yeats, P.A. 1988. The distribution of trace metals in ocean waters. Sci. Total Environ.
72:131-149. ICA no. Y-210.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 206
REFERENCES

436. Ylaranta, T. 1995. Effect of road traffic on heavy metal concentrations of plants. Agric. Sci.
Finland 4 (1):35-48. ICA no. Y-700.

437. Yousef, Y.A., Baker, D.M., and Hvitved-Jacobsen, T. 1996. Modeling and impact of metal
accumulation in bottom sediments of wet ponds. Sci. Total Environ. 189-190:349-354. ICA no.
Y-720.

438. Zahn, M.T. 1988. Distribution of Heavy Metals and Anions in the Groundwater of
Quaternary Gravel Deposits in the Munich Area (Dornach); Laboratory and Field Test Results,
Fak Geowissensch, Muenchen Univ., Muenchen, Germany. ICA no. Z-243.

439. Zhang, J. 1995. Geochemistry of trace metals from Chinese river/estuary systems: An
overview. Estuar. Coastal Shelf Sci. 41 (6):631-658. ICA no. Z-745.

440. Zolotukhina, E.Y., Radzinskaya, N.V., Tropin, I.V., and Gavrilenko, E.E. 1993. A study of
the macroalgae Chordaria flagelliformis (Mull.) Ag. and Rhodomela larix (Turn.) C. Ag. as
objects of heavy metals monitoring in the Peter the Great Bay, Sea of Japan: Vestn. Mosk. Univ.
(Biol.). ICA no. Z-687.

COPPER: Environmental Dynamics and Human Exposure Issues


Page 207

Anda mungkin juga menyukai