Download as pdf or txt
Download as pdf or txt
You are on page 1of 48

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/336538333

Sediment-Hosted Lead-Zinc DepositsA Global Perspective

Chapter · January 2005


DOI: 10.5382/AV100.18

CITATIONS READS
88 4,627

8 authors, including:

K. D. Kelley Ross Large


United States Geological Survey University of Tasmania
102 PUBLICATIONS 2,915 CITATIONS 291 PUBLICATIONS 13,829 CITATIONS

SEE PROFILE SEE PROFILE

J. Gutzmer Steve Walters


Helmholtz-Zentrum Dresden-Rossendorf MINOREvation
349 PUBLICATIONS 8,597 CITATIONS 43 PUBLICATIONS 1,943 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by J. Gutzmer on 01 December 2019.

The user has requested enhancement of the downloaded file.


©2005 Society of Economic Geologists, Inc.
100th Anniversary Volume, pp. 561–607

Sediment-Hosted Lead-Zinc Deposits: A Global Perspective


DAVID L. LEACH,†
U.S. Geological Survey, Mail Stop 973, Box 25046, Denver, Colorado 80225

DONALD F. SANGSTER,
2335 Russvern Drive, North Gower, Ontario, Canada K0A 2T0

KAREN D. KELLEY,
U.S. Geological Survey, Mail Stop 973, Box 25046, Denver, Colorado 80225

ROSS R. LARGE,
University of Tasmania, CODES, Private Bag 79, Hobart, Tasmania 7001, Australia

GRANT GARVEN,
Johns Hopkins University, Department of Earth and Planetary Sciences, 3400 North Charles Street, Baltimore, Maryland 21218

CAMERON R. ALLEN,
Teck Cominco American Inc., 15918 E. Euclid, Spokane, Washington 99216

FT JENS GUTZMER,
Rand Afrikaans University, Department of Geology, P.O. Box 524, Auckland Park, Gauteng South Africa 2006

A
AND STEVE WALTERS

R
University of Tasmania, CODES, Private Bag 79, Hobart, Tasmania 7001, Australia

Abstract

D
Sediment-hosted Pb-Zn deposits contain the world’s greatest lead and zinc resources and dominate world
production of these metals. They are a diverse group of ore deposits hosted by a wide variety of carbonate and
siliciclastic rocks that have no obvious genetic association with igneous activity. A range of ore-forming
processes in a variety of geologic and tectonic environments created these deposits over at least two billion
years of Earth history. The metals were precipitated by basinal brines in synsedimentary and early diagenetic
to low-grade metamorphic environments. The deposits display a broad range of relationships to enclosing host
rocks that includes stratiform, strata-bound, and discordant ores.
These ores are divided into two broad subtypes: Mississippi Valley-type (MVT) and sedimentary exhalative
(SEDEX). Despite the “exhalative” component inherent in the term “SEDEX,” in this manuscript, direct evi-
dence of an exhalite in the ore or alteration component is not essential for a deposit to be classified as SEDEX.
The presence of laminated sulfides parallel to bedding is assumed to be permissive evidence for exhalative ores.
The distinction between some SEDEX and MVT deposits can be quite subjective because some SEDEX ores
replaced carbonate, whereas some MVT deposits formed in an early diagenetic environment and display lam-
inated ore textures.
Geologic and resource information are presented for 248 deposits that provide a framework to describe and
compare these deposits. Nine of the 10 largest sediment-hosted Pb-Zn deposits are SEDEX. Of the deposits
that contain at least 2.5 million metric tons (Mt), there are 35 SEDEX (excluding Broken Hill-type) deposits
and 15 MVT (excluding Irish-type) deposits. Despite the skewed distribution of the deposit size, the two de-
posits types have an excellent correlation between total tonnage and tonnage of contained metal (Pb + Zn),
with a fairly consistent ratio of about 10/1, regardless of the size of the deposit or district. Zinc grades are ap-
proximately the same for both, whereas Pb and Ag grades are about 25 percent greater for SEDEX deposits.
The largest difference between SEDEX and MVT deposits is their Cu content. Three times as many SEDEX
deposits have reported Cu contents, and the median Cu value of SEDEX deposits is nearly double that of MVT
deposits. Furthermore, grade-tonnage values for MVT deposits compared to a subset of SEDEX deposits
hosted in carbonate rocks are virtually indistinguishable.
The distribution of MVT deposits through geologic time shows that they are mainly a Phanerozoic phenom-
enon. The ages of SEDEX deposits are grouped into two major groups, one in the Proterozoic and another in
the Phanerozoic. MVT deposits dominantly formed in platform carbonate sequences typically located within
† Corresponding author: e-mail, dleach@usgs.gov

0361-0128/01/0000/561-47 561
562 LEACH ET AL.

extensional zones inboard of orogenic belts, whereas SEDEX deposits formed in intracontinental or failed rifts,
and rifted continental margins. The ages of MVT ores are generally tens of millions of years younger than their
host rocks; however, a few are close (<~5 m.y.) to the age of their host rocks. In the absence of direct dates for
SEDEX deposits, their age of formation is generally constrained by relationships to sedimentary or diagenetic
features in the rocks. These studies suggest that deposition of SEDEX ores was coeval with sedimentation or
early diagenesis, whereas some deposits formed at least 20 m.y. after sedimentation.
Fluid inclusion, isotopic studies, and deposit modeling suggest that MVT and SEDEX deposits formed from
basin brines with similar temperatures of mainly 90° to 200°C and 10 to 30 wt percent NaCl equiv. Lead iso-
tope compositions for MVT and SEDEX deposits show that Pb was mainly derived from a variety of crustal
sources. Lead isotope compositions do not provide criteria that distinguish MVT from SEDEX subtypes. How-
ever, sulfur isotope compositions for sphalerite and galena show an apparent difference. SEDEX and MVT sul-
fur isotope compositions extend over a large range; however, most data for SEDEX ores have mainly positive
isotopic compositions from 0 to 20 per mil. Isotopic values for MVT ores extend over a wider range and include
more data with negative isotopic values.
Given that there are relatively small differences between the metal character of MVT and SEDEX deposits
and the fluids that deposited them, perhaps the most significant difference between these deposits is their de-
positional environment, which is determined by their respective tectonic settings. The contrasting tectonic set-
ting also dictates the fundamental deposit attributes that generally set them apart, such as host-rock lithology,
deposit morphology, and ore textures.
Brief discussions are also presented on two controversial sets of deposits: Broken Hill-type deposits and a
subset of deposits in the MVT group located in the Irish Midlands, considered by some authors to be a distinct
ore type (Irish type). There are no significant differences in grade tonnage values between MVT deposits and
the subset that is described as Irish type. Most features of the Irish deposits are not distinct from the family of
MVT deposits; however, the age of mineralization that is the same as or close to the age of the host rocks and
the anomalously high fluid inclusion temperatures (up to 250°C) stand out as distinctly different from typical
MVT ores. The dominance of bacteriogenic sulfur in the Irish ores commonly ascribed as uniquely Irish type
is in fact no different from several MVT deposits or districts.
A comparison of SEDEX and Broken Hill-type deposits shows that the latter deposits contain significantly
higher contents of Ag and Pb relative to SEDEX deposits. In terms of median values, Broken Hill-type de-
posits are almost three times more enriched in Ag and one and a half times more enriched in Pb compared to
other SEDEX deposits. Metamorphism is a characteristic feature but not a prerequisite for inclusion in the
Broken Hill-type category, and known Broken Hill-type examples appear to occur in Paleo- to Mesoprotero-
zoic terranes. Broken Hill-type deposits remain an enigmatic grouping; however, there is sufficient evidence to
support their inclusion as a separate category of SEDEX deposits.

Introduction deposits. We will not discuss the fracture-controlled deposits


SEDIMENT-HOSTED Pb-Zn deposits are a diverse group of ore in which fluorite is dominant and barite typically abundant
deposits hosted by a wide variety of carbonate and siliciclastic (e.g., Central Kentucky, Hansonburg, New Mexico, or the
rocks that have no obvious genetic association with igneous English Pennines) or the fluorite-rich deposits such as those
activity. These deposits have historically been the world’s in southern Illinois, considered a genetic variant of carbon-
most important lead and zinc resources and continue to dom- ate-hosted Pb-Zn deposits (Leach and Sangster, 1993). We
inate world production of these metals. Sphalerite and galena exclude the Pb-Zn vein districts such as Freiberg, Germany,
are the major sulfide minerals, although, in certain deposits, Coeur d’Alene, Idaho, Woodcutters in northern Australia,
significant amounts of iron sulfide are present. The deposits and the veins of the Rhenish massif in Germany, because
are typically Cu poor and some contain economically impor- some may be the product of deep-seated igneous or meta-
tant amounts of Ag and Ge. The nonsulfide gangue minerals morphic fluids. Skarn and igneous-related Pb-Zn manto de-
are mainly dolomite, siderite, ankerite, calcite, barite, and posits are also excluded from this paper (see Meinert et al.,
quartz (including chert and ore-related silicification). These 2005). We do not discuss oxide zinc deposits because they
deposits formed from a range of ore-forming processes in a have been discussed in a special issue devoted to “Nonsulfide
variety of geologic and tectonic environments over at least the Zinc Deposits: A New Look” (Economic Geology, v. 98, no. 4,
last two billion years of Earth history. The metals were de- 2003).
posited by basinal brines in synsedimentary and early diage-
netic to low-grade metamorphic environments. The deposits Objectives
display a broad range of relationships to enclosing host rocks, The main objectives of this paper are to summarize the
including stratiform, strata-bound, and discordant ores. The salient features of sediment-hosted Pb-Zn deposits and to de-
sulfides may be present as open-space fill and replacement of fine the current understanding on why and where these de-
sediments or sedimentary rocks, filling porosity created by posits formed in the Earth’s crust. The authors of this paper
displaced pore fluids in sediments, or precipitated directly on share broad agreement on many aspects of sediment-hosted
the sea floor. Pb-Zn deposits; however, there are differing opinions and in-
terpretations on a number of issues that required consider-
Excluded ore types able compromise and acceptance of alternative views. A fun-
Considering the large task of describing the broad subject damental contribution of this paper is the compilation of
matter, we excluded certain types of sediment-hosted Pb-Zn basic mineral resource data on known deposits, age of host

0361-0128/98/000/000-00 $6.00 562


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 563

rocks, age of ore formation, host-rock lithology, and related MVT deposits: The most important characteristics of MVT
deposit attributes in the Appendix A (Tables A1–A4). deposits, modified from Leach and Sangster (1993) are (1)
they are epigenetic; (2) they are not associated with igneous
The problem of classification activity; (3) they are hosted mainly by dolostone and lime-
Sediment-hosted Pb-Zn deposits are best described as a stone, rarely in sandstone; (4) the dominant minerals are spha-
family of related subtypes having sets of attributes that, to a lerite, galena, pyrite, marcasite, dolomite, and calcite; barite is
limited extent, set them apart from others. However, the di- typically minor to absent and fluorite is rare; (5) they occur in
versity of ore-forming environments and ore-deposit mor- platform carbonate sequences commonly at flanks of basins or
phologies, together with the overlapping attributes among the foreland thrust belts; (6) they are commonly strata bound but
deposit subtypes, has prevented establishing descriptive or ge- may be locally stratiform; (7) they typically occur in large dis-
netic models that uniquely separate the subtypes. The litera- tricts; (8) the ore fluids were basinal brines with ~10 to 30 wt
ture contains many classifications of sediment-hosted Pb-Zn percent salts; (9) they have crustal sources for metals and sul-
deposits, including use of descriptive attributes such as host- fur; (10) temperatures of ore deposition are typically 75° to
rock lithology (e.g., carbonate-hosted, shale-hosted), deposit about 200°C; (11) the most important ore controls are faults
morphologies (e.g., strata-bound, stratiform) or genetic char- and fractures, dissolution collapse breccias, and lithological
acteristics (e.g., syngenetic, synsedimentary, syndiagenetic). transitions; (12) the sulfides are coarsely crystalline to fine-
Previous reviews of sediment-hosted Pb-Zn deposits, in- grained, massive to disseminated; (13) the sulfides occur
cluding the Seventy-fifth Anniversary Volume of Economic mainly as replacement of carbonate rocks and to a lesser ex-
Geology (1981), Sangster (1996), Fontbote and Boni (1994), tent, open-space fill; and (14) alteration consists mainly of
and the 1966 Symposium on Carbonate-hosted Pb-Zn de- dolomitization, host-rock dissolution, and brecciation.
posits (Brown, 1967), concluded that no single classification Diversity between MVT districts resulted in the historic ap-
scheme provided rigorous classification criteria. Consistent plication of MVT subsets or alternative classifications to char-
with this “historical controversy”, the authors of this paper acterize specific sets of deposit attributes deemed unique
were unable to establish a classification scheme that uniquely within the broader family of MVT deposits. The use of sub-
separates each subtype. Consequently, we used a broad clas- sets mainly reflects a philosophical difference between geolo-
sification scheme that is biased toward descriptive attributes gists that are “splitters” and focus on the differences between
over genetic features, and provides alternative classifications ore deposits versus the “lumpers” that view diversity among
for many deposits. This classification scheme is used in the the ores as the norm. Important subsets historically used in-
text and is used for the 248 deposits described in Appendix A clude the Appalachian-, Alpine-, Bleiberg-, Upper Silesia-,
(Table A1: MVT and SEDEX, and Table A4: sandstone-lead Reocin-, and Irish-type deposits.
and sandstone-hosted lead). Irish-type: Of the MVT subsets or alternative classifications
mentioned, deposits in the Irish Midlands are debated as a
Classification and terminology distinct ore style referred to as Irish type (Table A1, column
The sediment-hosted Pb-Zn deposits are divided primarily G, “deposit type B”). The term “Irish-type” deposit is based
among four subtypes: Mississippi Valley-type (MVT), sedi- on current usage in the literature (e.g., Hitzman and Large,
mentary exhalative (SEDEX; Tables A1–A3), sandstone-lead 1986; Hitzman, 1999; Wilkinson, 2003). Hitzman and Beaty
(Sst-Pb; Table A4), and sandstone-hosted deposits. None of (1996, p. 499) include the following abbreviated characteris-
these deposits has any obvious genetic connection to igneous tics: (1) ore is located in the lowest nonargillaceous unit; (2)
activity. The sandstone-lead and sandstone-hosted lead de- ore is associated with normal faults; (3) sphalerite and galena
posits are only briefly discussed because they form small sub- dominate, pyrite may be abundant, and minor to abundant
types of sediment-hosted Pb-Zn deposits, and also because barite is present; (4) they are strata bound; and (5) they have
Sst-Pb deposits were extensively described by Bjørlykke and complex textures including replacement, colloform, and fine-
Sangster (1981). to coarse-grained sulfides and cavity fillings. Wilkinson (2003,
Neither the MVT nor SEDEX subtypes are ideal classifica- p. 980) added the following to the above list of characteristics:
tion terms nor do these categories have criteria that allow un- (1) the normal faults are synsedimentary; (2) ore formation
equivocal classification for all sediment-hosted Pb-Zn de- occurred during diagenesis; (3) orebodies are dominated by
posits. Our inability to select more appropriate terms, massive sulfides with lateral metal zonation; and finally (4) re-
together with the widespread use of these terms in the litera- duced sulfur is dominantly of bacteriogenic origin. In Appen-
ture, led us to continue using these two terms. Criteria used dix A, we included these deposits in the broad classification of
to classify the deposits and districts in Tables A1 to A3 were MVT ores but offer an alternative classification as Irish type.
based on the specific subtype assigned to the deposits in the We include them in the general discussion of MVT ores that
literature and, in some cases, the opinions of the authors. We follows with a discussion of the specific attributes that some
attempt here to minimize “process-related,” “interpretive,” believe uniquely define a different subtype of sediment-
and “model-driven” features to classify the deposits. The par- hosted Pb-Zn deposits.
ticular assignment of a subtype to some deposits may not be SEDEX: The term SEDEX evolved from the original term
universally accepted; therefore, we provided alternative clas- proposed by Carne and Cathro (1982) that included lami-
sifications (deposit type B and deposit type C, Tables A1, A4). nated, exhalative sulfides in fine-grained clastic rocks to a di-
Characteristics of the deposit classifications are briefly pre- verse group of deposits containing laminated ores in clastic,
sented below and discussed in more detail elsewhere in the carbonate, and metasedimentary rocks. Despite the exhala-
text. tive component inherent in the term SEDEX, we do not

0361-0128/98/000/000-00 $6.00 563


564 LEACH ET AL.

require direct evidence of an exhalite in the ore or alteration (Sinclair et al., 1999). The second author of this paper com-
component to be classified as a SEDEX. Rather, the presence piled the information on Pb-Zn deposits in the World Miner-
of laminated sulfides parallel to bedding is assumed to be per- als Geoscience database project and led the effort for the ex-
missive evidence for exhalative ores. Some SEDEX deposits panded database in Table A1.
are known to have replaced carbonate host rocks (e.g., Kelley We decided not to exclude deposits or prospects in the data
et al., 2004a). Thus, the distinction between some SEDEX compilation that were below some arbitrary resource thresh-
and MVT deposits can be quite subjective. old or a perceived value-in-the-ground, because a major ob-
The general characteristics of SEDEX deposits are sum- jective was the usefulness of the table to a diverse group of
marized in Gustafson and Williams (1981), Large (1983), geologists with presumably different economic objectives.
Sangster (1990), Goodfellow et al. (1993), Lydon (1996, The data will hopefully be used by exploration companies,
2004b), Sangster and Hillary (1998), Large et al. (2002), and academic researchers, and government agencies alike. Al-
Goodfellow (2004. The key classification features are (1) they though most of the small deposits and prospects will never be
occur as tabular Zn-Pb-Ag deposits that contain laminated, anything more than small deposits, many small deposits or
stratiform mineralization; (2) they may be hosted in shale, prospects have pointed the way for some major discoveries.
carbonate, or carbonate- or organic-rich clastic rocks (silt- Table 1 and Appendix B Figure B1 summarize the metal
stone and less commonly sandstone and conglomerate rich); endowments for the subtypes and these will be referred to
(3) spatially and/or genetically associated igneous rocks are throughout this paper. In terms of metal character, the most
generally absent or volumetrically minor; and (4) they form in striking difference between SEDEX and MVT deposits is
intra- and/or epicratonic rift and passive margin environ- that there are 13 SEDEX deposits (excluding Broken Hill-
ments (Large, 1980; Large et al., 2002). type deposits) with 10 Mt or more contained metal relative to
The divisions of the SEDEX deposits used in this report two MVT deposits of this size (excluding Irish-type). How-
(Table A1, column G) include Broken Hill-type (Broken Hill- ever, this comparison does not consider the resource data for
type), carbonate replacement (CR), carbonate-hosted (carb- two giant MVT districts (e.g., Viburnum Trend and Upper
hst), shale-hosted (sh-hst), and coarse clastic-hosted (cc-hst). Silesisa) that report data for “mines” that are actually “pro-
The host-rock terminology is based on the dominant rock duction units” within a more or less continuous mineralized
lithology that hosts the ore. “Unclassified” is used if meta- zone. In terms of median values for deposits, MVT size is
morphism and deformation of the proto-ore obscured the about 72 percent of SEDEX, Zn grade is approximately the
original deposit morphology and ore fabric so that it is impos- same, MVT Pb grade is 75 percent of SEDEX, and Ag grade
sible to determine the original ore subset. VMS (vol- is 82 percent that of SEDEX. Three times as many SEDEX
canogenic-hosted massive sulfide) is used for a few deposits deposits have reported Cu contents and the median Cu value
that are described in the literature as having possible igneous of SEDEX deposits is nearly double that of MVT deposits
affinities and that may reflect a transition between SEDEX (Table A1). Broken Hill-type deposits are significantly larger
and volcanogenic massive sulfide environments. than SEDEX in terms of size and Pb and Ag content. In
Broken Hill-type deposits: Some workers (including some terms of median values, MVT deposits (excluding Irish) are
authors of this paper) consider Broken Hill-type deposits to about 30 percent larger, contain 10 percent more Pb, 30 per-
be SEDEX deposits that formed from fluids, or in deposi- cent more Ag grade, and about the same Zn grade relative to
tional conditions, slightly different from typical SEDEX de- Irish type. However, the few Irish-type deposits do not yield
posits, whereas others (including other authors of this paper) a robust comparison.
consider Broken Hill-type to be a unique class and represent
a different genetic model from SEDEX deposits (see Large et Mississippi Valley-Type Deposits
al., 2005). In Table A1 (column G), we use Broken Hill type Mississippi Valley-type deposits are found throughout the
for deposits that have been considered in the literature to be world (App. A Map Files, App. B, Fig. B2) but are most abun-
Broken Hill type or if some authors of this paper preferred to dant in North America and Europe. Previous reviews of MVT
apply Broken Hill type to a particular deposit. deposits include Ohle (1959, 1980), Callahan (1964), Snyder
Mineralization is dominantly strata bound and hosted by (1967), Heyl et al. (1974), Hoagland (1976), Anderson and
amphibolite-granulite facies of Paleo- to Mesoproterozoic Macqueen (1982), Sverjensky (1986), Leach and Sangster
clastic metasedimentary host sequences and is characterized (1993), and papers in Brown (1967), Kisvarsanyi et al. (1983),
by complex metamorphic, metasomatic, and structural over- and Sangster (1996). Mississippi Valley-type deposits owe
prints. Consequently, protolith interpretations and definitions their commonly accepted name to the fact that several classic
of genetic models are varied and often poorly constrained. districts are located in the drainage basin of the Mississippi
Published genetic models range from syngenetic exhalative or River in the central United States. Early models called upon
subsea-floor replacement, distal skarn, metamorphogenic, or ore formation by dilute meteoric ground water, magmatic,
hybrid combinations (Walters et al., 2002). and connate fluids and invoked processes that ranged from
synsedimentary-exhalative to supergene activity involving
Data Compilations meteoric water. Sharp divisions of opinion existed between
The compilation of information on 248 deposits and dis- geologists who believed the ores were syngenetic, early dia-
tricts presented in Table A1 provides a framework and basis genetic, or epigenetic in origin. The syngenetic interpreta-
for describing and comparing the world’s sediment-hosted tions were mainly based on the stratiform nature of some ores
Pb-Zn deposits. The data compilations in these tables were and ore textures that are now known to be the result of host-
based on the World Minerals Geoscience database project rock replacement that mimic sedimentary features. Excellent

0361-0128/98/000/000-00 $6.00 564


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 565

TABLE 1. Summary Statistics for Select Parameters of Sediment-Hosted Pb-Zn Deposits,


including Subgroups MVT, SEDEX, Broken Hill-, and Irish-Type Deposits (data from Table A1)

Min 10th 25th 50th 75th Max n

MVT deposits (110 deposits, includes Irish)


Size (Mt) 0.6 0.65 2.0 7.2 19 300 110
Pb (Mt) 0 0.01 0.04 0.14 0.4 4.99 96
Zn (Mt) 0 0.04 0.13 0.37 1.08 15.62 104
Pb (%) 0.18 0.7 1.165 1.75 3.2 13 96
Zn (%) 0.23 1.9 3.7 6 9.58 24 104
Cu (%) 0.1 0.1 0.17 0.28 0.4 0.52 6
Ag (g/t) 1.2 10 17 37 69.7 485 31

SEDEX deposits (137 deposits, includes BHT)


Size (Mt) 0.01 1 2.78 11.5 42 476 137
Pb (Mt) 0 0.03 0.08 0.29 1 28 120
Zn (Mt) 0 0.05 0.15 0.58 1.97 27.39 130
Pb (%) 0.2 0.78 1.39 2.37 4.6 11.6 120
Zn (%) 0.4 1.66 3.11 5.12 8 21.7 130
Cu (%) 0.09 0.1 0.19 0.5 0.74 4.39 23
Ag (g/t) 1.1 10 23 46 71 538 76

MVT deposits (94 deposits, excludes Irish)


Size (Mt) 0.06 0.65 2 7.26 20 300 94
Pb (Mt) 0 0.01 0.05 0.15 0.46 5.0 80
Zn (Mt) 0 0.04 0.13 0.37 1.08 15.62 88
Pb (%) 0.3 0.7 1.17 2.0 3.5 13 80
Zn (%) 0.23 1.68 3.5 6 9.94 24 88
Cu (%) 0.1 0.1 0.15 0.26 0.33 0.4 5
Ag (g/t) 1.2 8.5 17 42 75 485 22

Irish deposits (16 deposits)


Size (Mt) 0.15 1.04 2.03 5.3 13.95 95.3 16
Pb (Mt) 0 0.002 0.02 0.08 0.29 2 16
Zn (Mt) 0.01 0.04 0.11 0.34 1.01 7.9 16
Pb (%) 0.18 0.64 1.2 1.5 1.8 6.19 16
Zn (%) 2.67 3.5 4.65 5.98 7.65 12.43 16
Cu (%) 0.52 NA NA NA NA 0.52 1
Ag (g/t) 10 11.6 21.6 27 39.7 66 9

SEDEX deposits (129 deposits, excludes BHT)


Size (Mt) 0.01 1 2.72 10 31 476 129
Pb (Mt) 0 0.03 0.08 0.26 0.95 19.5 112
Zn (Mt) 0 0.05 0.15 0.57 1.96 27.39 122
Pb (%) 0.2 0.77 1.39 2.26 4.1 8 112
Zn (%) 0.4 1.68 3.25 5.55 8.5 21.7 122
Cu (%) 0.1 0.1 0.2 0.6 0.9 4.39 19
Ag (g/t) 1.1 7.5 22 45 67 224 70

BHT deposits (8 deposits)


Size (Mt) 0.8 3.14 26.2 83.3 125.5 280 8
Pb (Mt) 0.09 0.26 0.75 1.6 4.06 28 8
Zn (Mt) 0.02 0.10 0.39 1.72 6.56 23.8 8
Pb (%) 0.55 0.69 1.84 5.64 10.5 11.6 8
Zn (%) 0.59 0.94 2.1 2.95 5.75 8.5 8
Cu (%) 0.09 NA 0.09 0.22 0.54 0.75 4
Ag (g/t) 12.9 14.59 29.83 98.05 400 538 6

Notes: Max = maximum value; Min = minimum value; NA = not applicable; 10th = 10th percentile; 25th = 25th percentile; 75th = 75th percentile, n = num-
ber of deposits with data for given parameter; (g/t) = grams per ton

perspectives of this debate are summarized by Snyder (1967), understanding their genesis was the suggestion that they
Brown (1970), and Ohle (1970). formed by the mixing of multiple basinal fluids with contrast-
Modern thoughts on the origin of MVT deposits began with ing reduced sulfur and metal content (e.g., Dunham, 1966).
fluid inclusion studies that showed the ore fluids were similar Early models assumed that MVT deposits had no connection
to oil-field brines (Newhouse, 1933; Hall and Friedman, with tectonic processes. However, in the last 25 years, studies
1963; Roedder et al., 1963; White, 1968). A major step in have noted the association of MVT ore genesis to major

0361-0128/98/000/000-00 $6.00 565


566 LEACH ET AL.

crustal tectonic events (e.g., Garven, 1985; Leach and Rowan izing the ore or if the dolostone was a product of the miner-
1986; Oliver, 1986; Bethke and Marshak, 1990; Oliver, 1992). alization event. For example, all of the ore produced from the
Perhaps the most significant achievements in understanding Upper Silesia district is hosted in what is locally called the
MVT ore genesis have been from advances in dating MVT “ore-bearing dolostone.” Two main opinions are debated for
ores (e.g., Leach et al. 2001, and references therein). the origin of the ore-bearing dolostone: pre- and synmineral-
ization (e.g., Sass-Gustkiewicz and Dzulynski, 1998; Leach et
Scale of mineralization and diversity al., 2003). In the Viburnum Trend, 95 percent of the ore is
Some MVT hydrothermal events formed multiple districts hosted in a regional dolostone, located less than 2 km from a
in a large area that define huge metallogenic provinces (e.g., facies transition to limestone. The host dolostone formed dur-
Ozark and the Appalachian provinces: Ohle, 1959; Leach, ing several preore diagenetic events (e.g., Gerdemann and
1994, Kesler, 1996). Deposits within districts usually have Meyers, 1972; Lyle, 1977). However, there is abundant re-
similar deposit attributes and ore controls, whereas districts gional hydrothermal dolomite that is pre-, syn-, and postmin-
can be quite distinct. This diversity is well illustrated by the eralization (Gregg, 1985; Voss and Hagni, 1985; Rowan,
Ozark MVT province in the United States midcontinent (Fig. 1986).
B3) that includes the world-class districts of the Old Lead
Belt and Viburnum Trend (galena dominant), the Tri-State Grade and tonnage
district, and the smaller Northern Arkansas, Central Missouri Mississippi Valley-type deposits are characteristically dis-
(sphalerite, galena, plus barite), and Southeast Missouri tributed over hundreds of square kilometers that define indi-
barite districts (barite dominant), and regional iron sulfide vidual ore districts. Large MVT districts include Southeast
mineralization. This metallogenic province was the product of Missouri (3,000 km2), Tri-State (1,800 km2), Pine Point (1,600
an enormous hydrothermal system that affected more than km2), Alpine (10,000 km2), Upper Silesia (2,800 km2), Irish
350,000 km3 of rock in the late Paleozoic in response to con- Midlands (8,000 km2), and the Upper Mississippi Valley
vergent plate tectonics in the Ouachita fold belt (Leach, 1994, (7,800 km2). Pine Point contains more than 80 deposits and
and references therein). Each ore district has a unique com- the Upper Mississippi Valley nearly 400 deposits. Individual
bination of ore assemblages, ore controls, host rocks, alter- deposits can be small in some districts. Most deposits in the
ation, isotopic characteristics, and geochemistry (Table B1; Pine Point district contain between 0.2 and 2 million tons
Fig. B4). The diversity shown by these ore districts is the re- (Mt) of ore, and the largest has nearly 18 Mt (Sangster, 1990).
sult of the interplay between a range of ore depositional Heyl et al. (1959) reported that the average deposit size in the
processes (Fig. B5) and paleohydrological controls on the mi- Upper Mississippi Valley district was between 0.1 and 0.5 Mt,
gration of the ore fluids, together with rock-water interactions although a few contained as much as 3 Mt of ore.
in multiple aquifers. Because production and reserve data for MVT deposits,
The Appalachian province consists of several districts and particularly in the United States, are commonly presented
widespread occurrences in the eastern United States and only as district totals, the size, grade, and metal-ratio parame-
Canada that formed during a series of Paleozoic tectonic ters for individual deposits are difficult to compile. For exam-
events in the Appalachian orogen (Kesler, 1996). Kesler (1996) ple, the Viburnum Trend deposits in Table A1 are effectively
proposed that mineralization was controlled by four different production units within a mineralized zone that continues for
paleoaquifers, within some 10 Appalachian MVT brine more than 60 km (Fig. B6).
provinces, which produced numerous deposits by the mixing The largest 30 MVT deposits are shown in Figure 1. Figure
of metal-bearing brines with fluids containing reduced sulfur. 2 compares the size of MVT districts to individual deposits
Included in the Appalachian province are the East Tennessee and exhibits a correlation between size and grade that will be
district (including the Mascot-Jefferson City district), New- addressed below. Navan (95.3 Mt at 2.1 wt % Pb and 8.3 wt
foundland Zinc, and the Austinville-Ivanhoe district. % Zn) is the largest active mine, with a metal endowment
greater than most districts.
Host rocks
The age of MVT host rocks peaks in the Cambrian to Late Mineralogy, zoning, and geochemical halos
Carboniferous and Triassic to Cretaceous with a paucity of Most MVT deposits have a simple mineralogy, consisting
deposits in Silurian and Permian rocks. Despite the abun- primarily of sulfide minerals that are dominated by sphalerite,
dance of Proterozoic carbonate rocks, there are relatively few galena, and iron sulfides. Barite may be present in minor to
MVT deposits in these rocks. The significance of these obser- trace amounts and fluorite is rare. Barite dominates over sul-
vations is discussed below. fides in a few smaller districts that include the Central Mis-
More MVT deposits are hosted in dolostone relative to souri and Southeast Missouri barite districts (Leach, 1980)
limestone, and dolostone-hosted deposits are generally larger and the Sweetwater district in Tennessee (e.g., Kesler, 1996).
and contain higher Pb, Zn, and Ag grades. The preference for The Viburnum Trend has one of the most complex mineralo-
mineralization to be hosted by dolostone may be related to a gies, consisting of a variety of Cu, Co, Ni, Fe, Ag, and Sb sul-
higher transmissivity to fluid flow relative to limestone. An ex- fides and sulfosalts (e.g., Heyl, 1983). Cadmium, Ge, Ga, In,
ception to this tendency is the large and high-grade Navan barite, and fluorite have been recovered from some districts.
deposit in the Irish Midlands that is hosted by limestone. Copper and Ni are not common constituents of MVT deposits
Most dolostones in MVT districts have a complex history except those in the Southeast Missouri lead district. Silver is
that can include pre-, syn-, and postore dolomitizing events. reported for 31 of 110 deposits in Table A1, with a median
It is commonly unclear if the dolostone played a role in local- value of 37 g/t and the 90th percentile value of 41 g/t. Total

0361-0128/98/000/000-00 $6.00 566


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 567

22.5 Missouri district. Some ores, such as those in the East Ten-
nessee and Newfoundland Zinc districts, are essentially Pb
20 free.
17.5
In the Ozark region (Fig. B3), Erickson et al. (1983) iden-
Total Pb + Zn Metal (Mt)

tified regional geochemical halos of trace elements in insolu-


15 ble residues of barren carbonate rocks that extend for several
hundred kilometers from the MVT districts. Primary disper-
12.5 sion halos of metals around deposits are generally small (less
than 50–75 m in the Upper Mississippi Valley districts and
10
less than 125 cm in Central Tennessee; Lavery et al., 1994,
7.5 and references therein). Zones of Zn-bearing hydrothermal
dolomite (Zn bound in the crystal lattice) extend to 50 m from
5 ore in the Viburnum Trend (Viets et al., 1983) and several
kilometers in the Upper Silesia district (Bak, 1986).
2.5
Geochemical zoning in MVT deposits has been described
0 in only a few areas (e.g., in the Irish Midlands, Southeast Mis-
Touissit-Bou Beker
Mehdiabad
Admiral Bay
Navan
Reocin
Fankou
Pavlovskoye
Buick
Komdok
Daliangzi

Polaris

Florida Canyon

Tianbaoshan
Lisheen

El Abed
Pillara (Blendvale)
Urultun
Magmont
Nanisivik
Huayuan
Silvermines
Galmoy
Emarat
Raibl
Gayna River
Sumsar
Prairie Creek
Bleiberg

Cadjebut Trend

souri, Pine Point, and Upper Mississippi Valley districts). Per-


San Vicente

haps the best-developed zoning is for some deposits in the


Irish Midland, where metal-bearing fluids were highly fo-
cused into limestone by steeply dipping basement-rooted
faults. Deposits in the Southeast Missouri districts have zon-
FIG. 1. Bar diagram showing largest 30 MVT deposits based on total Pb +
ing patterns of Pb, Zn, Fe, Cu, Ni, and Co; however, these
Zn content (Mt). Data listed in Table A1. patterns are highly irregular along the trend of deposits (e.g.,
Hagni, 1983; Mavrogenes et al., 1992, and references
therein). At Pine Point, Fe/(Fe + Zn + Pb) and Zn/(Zn + Pb)
silver production can be substantial; for example, during the ratios increase outward from prismatic brecccia (high verti-
period 1960 to 1990, silver production from the Viburnum cal-lateral aspect ratio) orebodies (Kyle, 1981).
Trend averaged 45.9 t/yr. With the exception of the Southeast
Missouri district, all major MVT districts are Zn rich relative Alteration
to Pb and have Zn/(Zn + Pb) ratios greater than 0.5. By Dissolution of carbonate host rocks: The most common al-
contrast, this metal ratio is less than 0.1 for the Southeast teration associated with MVT deposits is dissolution and hy-
drothermal brecciation of the carbonate host rocks that result
from acid-producing reactions commonly associated with
Deposits fluid mixing (Corbella et al., 2004). The most significant acid
100 production occurs from the mixing of reduced sulfur-bearing
Mehdiabad fluids with a metal-bearing fluid and, to a lesser extent, by sul-
Reocin fate reduction in the ore zone.
Navan Hydrothermal carbonates: Hydrothermal dolomite occurs
Fankou in most MVT deposits as replacement of the host carbonates
10 Buick Admiral Bay and/or as cement in intergranular porosity and open-space
Pavlovskoye fill. Calcite is common in a few MVT districts (e.g., Central
Pb +Zn (Mt)

Polaris Komdok Tennessee) and is more likely to be present in limestone-


Daliangzi
dominated lithologies (e.g., Irish Midlands). Hydrothermal
dolomite may be pre-, syn-, or postore and form distinctive al-
teration halos around the deposits (e.g., Irish Midlands, East
1
Tennessee, and Tri-State districts). At Pine Point, Krebs and
Macqueen (1984) reported eight major paragenetic stages,
within which there were at least seven periods of dolomite or
calcite deposition. Hydrothermal dolomite cement, com-
monly associated with trace amounts of sulfides, may be pre-
sent in ore-barren carbonate rocks hundreds of kilometers
1 10 100 1000
Size (Mt) from the ore deposits, reflecting the regional scale of some
Districts MVT ore events, such as in the Ozark region (Rowan, 1986),
Austinville-Ivanhoe Lennard Shelf Sanandaj - Sirjan
Central Tennessee Metaline SE Missouri Lead
Central and East Tennessee districts (e.g., Montanez, 1996),
Cevennes Morrocco Tri-State and the Western Canada basin (e.g., Qing and Mountjoy,
Cornwallis Nanling Upper Mississippi Valley 1992).
Irish Midlands Pine Point Upper Silesia Although hydrothermal dolomite is usually associated with
Kildare Red Sea Touissit-Bou Beker MVT mineralizing events, its complex relationship with sulfides
FIG. 2. Size (Mt) of MVT deposits vs. total metal (Pb + Zn) content (Mt) and common widespread distribution without associated min-
for MVT deposits and districts. Data listed in Table A1. eralization restricts its use in exploration. The Irish Midlands

0361-0128/98/000/000-00 $6.00 567


568 LEACH ET AL.

is an exception because specific types of hydrothermal lithological controls. Important variations on fault-controlled
dolomite (dolomite-matrix breccias) are usually associated morphology range from the steeply dipping fracture-con-
with mineralization (Hitzman et al, 1992; Doyle and Bowden, trolled ores of Raible and the Irish Midlands (discussed in
1995; Hitzman, 1996; Wilkinson and Earls, 2000). more detail below), to the graben-controlled ores of Upper
Silicification: Quartz and hydrothermal silicification of the Silesia (Fig. B7), Morocco (Fig. 3B), and Reocin (Fig. 3C).
host carbonate rocks are generally minor in most MVT de- Salt-dome environments provide a minor but interesting
posits. The amount of quartz present depends largely on the form of MVT ores. Sulfides are present in the cap-rock min-
temperature of ore deposition and the amount of cooling that eral assemblage, as fracture fill and as replacement of car-
occurred during ore formation. Silica precipitation can be bonate host rocks. The most important deposits are the Fedj
suppressed by dilution and cooling as a consequence of fluid el Adoum and Bou Grine deposits (Fig. B8) in Tunisia. Other
mixing during ore deposition (e.g., Plumlee et al., 1994), due significant but presently noneconomic occurrences are in the
to kinetic inhibition of precipitation at low temperatures. Gulf Coast of the United States (Kyle and Saunders, 1996)
However, if the dilutant is above about 200°C, the rate of sil- and Tunisia (Rouvier et al., 1985; Orgeval, 1994).
ica precipitation is higher, and thus there is more quartz in
hotter systems (Rimstidt, 1997). Extensive silicification is Ore textures
most abundant in the Tri-State (Brockie et al., 1968) and the Deposition of ore in MVT districts throughout the world
Northern Arkansas districts (McKnight, 1935). involves an inextricable link between sulfide precipitation,
Organic matter: Although organic matter is present in vari- dissolution, host-rock replacement, open-space filling, and
able amounts and types in MVT deposits (Leventhal, 1990), solution-collapse brecciation. This interrelationship exists be-
its relationship to ore genesis is generally uncertain. Liquid or cause sulfide precipitation is nearly always an acid-generating
solid petroleum-type organic matter, if present in the ores, process. Ore textures from MVT deposits can be highly com-
has rarely been observed in primary fluid inclusions in main- plex due to commonly inherited fabrics of the replaced host
stage minerals. Studies of organic material from the Vibur- rock and the ore-forming processes. The ores can be ex-
num Trend district (e.g, Leventhal, 1990) show that organic tremely fine grained but in some districts crystals up to a
matter in the ore was thermally and compositionally altered meter or more occur (e.g., Tri-State and Central Tennessee
by the ore fluids. Thermal alteration of biomarkers in organic districts).
matter was used to calculate the duration of the ore-forming The large variety of ore textures present in MVT deposits
event in the Upper Mississippi Valley district (Rowan and are described in detail in Leach and Sangster (1993) and ref-
Goldhaber, 1995). erences therein. The most important textures include the fol-
Clay, mica, and feldspar diagenesis: Formation or destruc- lowing features. Fine-grained banded ore: Fine-scale banding
tion of clay, mica, and feldspar minerals is recognized in a few (millimeter- to micron-size bands) is common in individual
deposits. Authigenic K silicates related to ore formation are crystals and in some deposits (e.g., Upper Mississippi Valley
reported in the Upper Mississippi Valley (Hay et al., 1988), district) and can be correlated throughout the district (McLi-
East Tennessee (Hearn et al., 1987), Newfoundland Zinc mans et al., 1980; Fig. 4); colloform and dendritic ores: These
(Hall et al., 1989), and Gays River (Ravenhurst et al., 1987) are fine-scale (<1-mm) banded sulfides, spherolitic aggre-
districts. Alteration of feldspars has been observed in the gates of colloform sphalerite, dark bands of fibrous aggregates
Southeast Missouri lead district (Stormo and Sverjensky, of crystals (schalenblende), zones of dendritic galena, and
1983) and regionally in the basal sandstone (Diehl et al., iron sulfides in discrete bands or botryoidal masses. This tex-
1989). ture provides evidence for open-space deposition. (Fig. 4D-
F); replacement ores: Nearly complete replacement of host
Deposit morphology rock can produce massive sulfides (e.g., Nanisivik, Fig. 4A, C,
Deposits are commonly discordant on a deposit scale but G, H). Replacement can be remarkably selective (Fig. 4H, I)
strata bound on a district scale. The location and geometry of for a specific rock fabric; sulfides can replace fossils, stylolitic
most MVT deposits reflect the interplay between faults, pre- seams, thin organic-rich or stromatolitic layers (Fig. 4I); dis-
ore carbonate dissolution features, and permeable strati- solution collapse breccias: These are commonly multistage
graphic units. In some deposits, extensive mineralized zones breccias that may include pre- and synore breccias. Textures
are stratiform in nature (Fig. 3A). in dissolution collapse bodies include (1) rock-matrix brec-
One of the most characteristic morphologies of many MVT cias, composed of fragments of the host carbonate rocks sup-
deposits is dissolution-collapse breccias. They vary from the ported by a matrix of finer grained carbonate rock fragments
prismatic breccias of Pine Point (Rhodes et al., 1984; Kyle, (Fig. 4J); (2) crackle breccias, which may occur in the upper
1981) and columnar breccias of East Tennessee (Ohle, 1985) part of dissolution collapse bodies, which consist of highly
to the low domal breccias and stratiform breccias of the Tri- fractured rock without rotation of individual clasts; (3) ore-
State and Upper Silesia districts. In plan view, some breccia- matrix breccias (Fig. 4B, E, K, M, P) that consist of host-rock
hosted deposits (East and Central Tennessee districts) appear fragments cemented by sparry calcite, dolomite, or sulfides;
as a reticulate pattern similar to karst cave systems. In con- (4) hydrothermally altered and replaced breccias (Fig. 4L-N)
trast, the Viburnum Trend contains breccia-hosted ores that that are produced by hydrothermal dissolution coeval with
occur in narrow breccia trends (less than several hundred me- ore formation, and the presence of rotated and transported
ters wide but up to 10 km long). clasts of previously deposited sulfides are diagnostic; and (5)
Fault-controlled deposits vary significantly in shape and internal sediments (Fig. B9) of stratified sand- to clay-size
size, reflecting the influence of a variety of structural and disaggregated wall rock may contain sulfide and rock clasts

0361-0128/98/000/000-00 $6.00 568


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 569

A) B) N
NW SE S Chebikat
Mennjel
Mine el Akhal Missouine Graben el Hamra
Horst

r
Rive
Horst
Shaft 1 Shaft 9 Shaft 12
Metaline Fm
Ledbetter Fm Josephine

Yellowhead

Yellowhead mineralization Slate Creek Fault


Josephine mineralization 500 m

C) Old open pit MEKTA BEDDIANE


SE ore zone ore zone
NW
~ 50 m
Rubble ~ 1 km

r VISEAN BASEMENT MESOZOIC COVER


Su
pa
Ca Upper Bajocian-Bathonian (marls
and carbonated sandstones)
Schists
rte Callovo-Oxfordian (marls and sandstones)
o
p aN Dacites
Ca Aalenobajocian (dolostones)

Marly limestones and calcarenites Pb-Cu (Zn) mineralization


Sandstones and marls
Dolostones, sandstones, and marls
E)
100 m Dolostones
SW SILVERMINES NE
0 500 1000 m Marly limestones
Ore zones and lens
Chadian carbonate turbidites
G Zone
D) Ballysteen Limestone Waulsortian Limestone
S WAULSORTIAN N Lower
Palaeozoic Old Red
rocks Sandstone B Zone
Pyrite
ABL Pyrite HG DS
200m

Oolite
S TYNAGH N
X XX Zinc - lead - (iron) ore zones
X XX ABL MS ABL X
X
ABL X Silica - hematite zone
Not to Scale (”Iron formation”)
Old Red Ballysteen Limestone Hydrothermal dolostone
Type 1 - Pyritic massive sulfide X X
Type 4 - Vein/Stringer sulfide Sandstone (ABL) breccia
Type 2 - Massive sulfide Type 5 - ABL hosted sulfide Interbedded shale and limestone
(Silvermines)
Type 3 - Disseminated sulfide Type 6 - Oolite hosted sulfide
Reef core (Tynagh only)
ABL MS - Massive sulfide in ABL HG DS - High grade disseminated Fault
mineralization
FIG. 3. Sections showing geology and mineralization in select MVT deposits. A. Cross section of the Metaline area, Wash-
ington State, United States. Modified from Hoy et al. (2000). B. Generalized section showing piano keylike horst/graben
structure of the Toussit-Bou Bou-Beker mining district, Morocco. From Bouabdellah et al. (1996). C. Cross section through
the central part of the Reocin deposit, Spain, showing the relationship of the ore lenses and the host dolostones. Modified
from Velasco et al. (2003). D. Conceptual model (not to scale) for mineralization at Lisheen, Ireland. From Fusciardi et al.
(2003). E. Cross sections through the Silvermines and Tynagh deposits, Ireland. From Hitzman and Beaty (1996).

(Fig. 4O). Fault and sedimentary breccias: Fault breccias clasts in open space (Fig. 4D, P); (2) pseudobreccia is a brec-
consist of clasts of adjacent local host rocks in contrast to cia-like fabric that is produced by the selective replacement
polylithic fragments in dissolution collapse breccias. Sedi- of a particular host-rock fabric; (3) zebra texture is usually
mentary breccias can show soft-sediment deformation and, in gangue dolomite that formed by a variety of processes (i.e.,
the case of the Boulder Conglomerate in the Navan mine, selective replacement of primary rock fabric, open-space fill-
provide critical constraints on timing of mineralization. Other ing of dilated bedding planes and fractures); (4) rhythmites
textures include: (1) snow on roof which typically consists of consist of rhythmically banded ore containing distinct gener-
sulfides preferentially coating the tops of crystals or breccia ations of sphalerite or gangue carbonate that usually develop

0361-0128/98/000/000-00 $6.00 569


570 LEACH ET AL.

A B

C D

10 cm
2 cm

E F

1 cm

G H

1 cm

FIG. 4. Samples from selected MVT deposits showing diversity of textures. A. Zebra texture from the Newfoundland Zinc
district consisting of interlayered dolomite, host rock, and sulfides. Photograph from Tom Lane. B. Sphalerite and hy-
drothermal dolomite cementing dolostone breccia clasts, Lucky Dog Mine, northern Arkansas district. C. Stratiform spha-
lerite and pyrite from the Reocin mine, Spain. This ore type extends for hundreds of meters in the lower part of the deposit.
D. Sample of spheroidal sphalerite with dendritic sphalerite, galena, and iron sulfides (pyrite plus marcasite) from the Po-
morzany mine. Note the “up direction” is indicated by “snow-on-the-roof” texture of galena and sphalerite along colloform
layers of light-colored sphalerite. Hydrothermal sulfide clasts in the left center of the sample are encrusted by sphalerite and
iron sulfides. Size of sample is 20 × 13 cm. E. Brecciated fragments of colloform sphalerite cemented by hydrothermal
dolomite, Pine Point district, Canada. F. Dendritic galena and colloform sphalerite from the Pomorzany mine, Upper Sile-
sia, Poland.

0361-0128/98/000/000-00 $6.00 570


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 571

I J

5 cm

K L

M N

3cm

O P

10 cm 2 cm

FIG. 4. (Cont.) G. Massive sulfide ore from the Nanisivik deposit, Northwest Territories, Canada, showing bedding-par-
allel sphalerite and pyrite with sparry dolomite. H. Selective replacement of fine-grained dolostone by pyrite, controlled by
fractures. Fabric of the host rock remains visible in the pyrite. From Salah Bouhlel. I. Selective sphalerite replacement of
thin, organic-rich layers in dolostone, Monte Cristo Mine, northern Arkansas district. J. Preore dissolution collapse breccia
in the Buick mine, Viburnum Trend. This breccia is the host for the ore matrix breccia in Figure 11N. K. Sample from the
Gayna River Proterozoic MVT deposit, Canada. L. Hydrothermally altered and replaced preore dissolution collapse breccia
in the Upper Silesia district, Poland. Stratigraphic top is to left in the image. Note the selective replacement of the fine rock
matrix by sulfides. M. Replaced preore dissolution breccia of the Yellowhead ore zone in the Metaline mine, Washington,
United States. Superimposed hydrothermal alteration has modified the original texture. Note rotated mineralized clasts.
From Jerry Zieg. N. Galena replacement of fine rock matrix in preore dissolution collapse breccia in the Buick mine, Vibur-
num Trend. O. Hydrothermal internal sediments for the Polaris mine, Northwest Territories, Canada. Note clastic sulfides
and the graded bedding. P. Cavity fill of sparry dolomite and sphalerite from the Newfoundland Zinc district. Note the spha-
lerite lining the bottoms of cavities and top of clasts. From Tom Lane.

0361-0128/98/000/000-00 $6.00 571


572 LEACH ET AL.

from cycles of dissolution and open-space filling; and (5) spe- serious conflict exists between Rb-Sr and paleomagnetic
oleothem-like sulfides consists of a variety of dripstone-like dates for three of the four districts in dispute: Pine Point, East
stalactites, stalagmites, curtains, and drape stones. Tennessee (Mascot-Jefferson City), and Upper Silesia. The
Heijlen et al. (2003) study for Upper Silesia was challenged
Age of ore deposition by Bradley et al. (2004) and Emsbo et al. (in prep.) because
For many years, the most significant obstacle to under- of clay included in the sphalerite from the Upper Silesia dis-
standing the genesis of MVT deposits was the lack of infor- trict. The pros and cons of Rb-Sr isotope and paleomagnetic
mation on the age of mineralization. Recent advances for dat- dating of MVT deposits are presented in Kesler and Carrigan
ing MVT deposits (Leach et al., 2001) have used (2002), Leach et al. (2002), Bradley et al (2004), and Kesler et
high-precision paleomagnetic dating and radiometric dating al. (2004). Resolution of this dispute is a critical area for fu-
by Re-Os, U-Pb, U-Th in calcite, Rb-Sr in sphalerite, Ar-Ar ture research.
and K-Ar on feldspar and clay minerals. The results of these
dating efforts are shown in Figure 5, including recent dates Tectonic setting of MVT deposits
for Re-Os on bitumen from the Polaris mine (Selby and The most important period for MVT genesis was the De-
Creaser, 2003), an Ar-Ar date for Nanisivik (Sherlock et al., vonian to the Permian, which corresponds to a series of in-
2004), and an Rb-Sr date for Upper Silesia (Heijlen et al., tense tectonic events during the assimilation of Pangea (Fig.
2003). Leach et al. (2001) did not include a few published 5). The second most important period for MVT deposit gen-
dates because they provided unusually broad age constraints esis was the Cretaceous to the Tertiary when microplate as-
(minimum or maximum ages) or they were geologically un- similation affected the western margin of North America and
reasonable (i.e., provided ages that were older than the host Africa-Eurasia. Based on the deposits that have been dated, it
rocks). appears that few MVT deposits were formed during periods
There is good agreement between the reported ages from of major plate dispersal represented by the Late Proterozoic
radiometric and paleomagnetic techniques. Nevertheless, the to early Paleozoic period and the rapid breakup of Pangea in
timing of ore deposition continues to be hotly debated for the Triassic through the Jurassic. Thus, the time of most MVT
some districts. Part of the debate is due to large uncertainties deposits broadly coincides with major contractional events in
in the ages for some studies (on the order of ±10–20 m.y.) and the Phanerozoic. This interpretation is consistent with a long-
debatable interpretations of the geologic constraints on the recognized connection between many MVT deposits and oro-
timing of ore deposition. Only four of the 19 districts or de- genic forelands (Garven, 1985; Mitchell, 1985; Leach and
posits that have been dated have conflicting ages of mineral- Rowan, 1986; Oliver, 1986; Leach et al., 2001). Bradley and
ization. One dispute is between the paleomagnetic age Leach (2003) examined the tectonic aspects of foreland evo-
(Symons et al., 2001) and Ar-Ar age for the Proterozoic- lution and concluded that the type of foreland is not a critical
hosted Nanisivik deposit (Sherlock et al., 2004). The most control on ore formation. Mississippi Valley-type deposits are
located in collisional (Ozark MVT province), Andean (west-
ern Canada basin deposits), and transpressional orogens
(Cévennes district; Fig. 6). Also associated with orogens are
Assimilation of Pangea Pangea Pangea Rapid Alpine-
MVT deposits that occur in fold and thrust belts. Some de-
Stable Unstable Breakup Laramide
District/Deposit of Pangea Assimilation posits formed in flat-lying rocks or were later caught up in the
age PC C O S D C P Tr J K T thrusting. Others formed synchronously with thrusting and
others formed after burial by the thrusts.
Ma 600 500 400 300 200 100
Some MVT deposits clearly formed in a large-scale exten-
Polaris
sional environment. The best examples of MVT formation in
Pine Point
Newfoundland an extensional environment are the Lennard Shelf deposits.
Zinc The age obtained for ore deposition (Christensen et al., 1995;
Gays River
Eastern Tennessee Brannon et al, 1996) coincides with crustal extension of the
Central Tennessee Fitzroy trough. There are no compressive tectonic events that
Southeastern
Missouri
affected the region since the time that the host rocks were
Central Missouri deposited.
N. Ark.Tri-State
Avecilla Extensional tectonic control
Upper Silesia Although MVT deposits typically formed during crustal-
U. Miss Valley scale contractional events, the single most important tectonic
Lennard Shelf
Cevènnes
controls at the deposit or district scale are extensional faults
Irish (normal, transtensional, and wrench faults) and associated
Robb Lake Legend fractures and dilatancy zones. This relationship applies to
Monarch- Host Rocks most MVT districts, including the Ozark MVT province (e.g.,
Kicking Horse
Paleomag. Clendenin and Duane, 1990). Bradley and Leach (2003) dis-
Nanisivik*
* Paleomagnetic age for Nanisivik 1100 Ma Radiometric cussed MVT deposits that formed within extensional domains
FIG. 5. Distribution of radiometric and paleomagnetic ages of MVT de-
that developed due to lithospheric flexure (Fig. 7) or in large
posits and their host rocks in the Phanerozoic. Modified from Leach et al. dilational zones within bounding strike-slip faults during
(2001). large-scale contractional events. Ordovician normal faults

0361-0128/98/000/000-00 $6.00 572


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 573

A) Collisional orogen A) Syncollision


continental migrating
arc crust crust forebulge axis of advancing
Ozark MVT drowning submarine
oceanic crust
foredeep Province emerging karst platform outer slope migrating
foredeep orogen
platform sea level
mantle lithosphere

asthenosphere SEDEX barite


backflow

B) Andean-type orogen foreland Western MVT


basin Canada flexurally
Basin MVT (Sedex?)
induced
normal fault

~250 km

asthenosphere
backflow B) Post collision
overfilled degraded
200 km forebulge mountain front
foreland basin

C) Inversion orogen
foreland foreland
Ave basin basin Cevennes
recharge

basinal
brines

MVT in MVT in fault


Ore deposit buried karst
FIG. 6. Comparison of collisional, Andean, and transpressional orogens. A.
Collisional-type orogen. Arc-passive margin collision, based on Neogene ex- ~250 km
amples from Timor, New Guinea, and Taiwan, and various older examples in- Ore deposit
cluding the Taconic and Ouachita orogenies. B. Andean-type orogen. Based FIG. 7. Block diagrams showing foreland evolution. A. During plate con-
on the present-day Andes and the Late Cretaceous-Paleocene Laramide sys- vergence, a submarine thrust belt loads the passive margin, thereby forming
tem of western North America. Convecting asthenosphere contributes to the foreland basin, extensional domain, and forebulge. Plate convergence
foreland subsidence on a broad, regional scale, which sets this type of fore- continually causes these features to migrate across the foreland plate. The
land system apart from others. C. Inversion-type orogen. Flanking both sides foreland basin remains underfilled because the depocenter migrates. Barite
are thrust-loaded foreland basins, based on the Pyrenees. From Bradley and mineralization along the foredeep axis is based on examples from the Oua-
Leach (2003). chitas. B. Plate convergence has ceased and the foreland basin has filled with
sediment, creating hydrologic conditions favorable for MVT mineralization.
This is the situation corresponding to mineralization in the Ozark region.
related to the Taconic collision controlled MVT mineraliza- From Bradley and Leach (2003).
tion in the Newfoundland Zinc district (Bradley, 1993). Far-
field tectonic effects in the host rocks for MVT deposits com-
monly reflect reactivation of preexisting faults and fractures ore deposition but drifted apart during the opening of the At-
in the basement (e.g., Ozark province, Irish Midlands, and lantic (Leach et al., 2001). In many districts (e.g. the deposits
Cèvennes). Far-field extension inboard of orogenic belts can in the Ozarks region in the United States midcontinent, Irish
extend several hundred kilometers into the foreland and may Midlands, Upper Silesia, and Cèvennes districts), the fluid in-
explain districts such as those in the Western Canada basin, clusion temperatures generally exceed values expected for ge-
Upper Mississippi Valley, and the Irish Midlands. Thus, the ologically reasonable thermal gradients and estimated strati-
endowment of orogenic forelands with MVT deposits reflects graphic burial temperatures at the time of mineralization.
the link between contractional events in orogenic zones, in- Therefore, the temperatures observed for many ore deposits
board extensional tectonic domains, and fluid flow. The ex- must be explained by unusually high geothermal gradients,
tensional domains provide fluid drains for large regional advective heat transport from deeper parts of the basin (i.e.,
aquifers (i.e., Ozark region) or focused pathways for the as- Ozark MVT province), or the ascent of deeply circulating flu-
cent of fluids in buoyancy driven hydrothermal systems. ids in basement rocks beneath the deposits (e.g., Upper Sile-
sia and Irish Midlands).
Nature of the ore fluids Salinity: The salinities of MVT fluids determined from
Temperature of the ore fluids: Fluid inclusion temperatures fluid inclusions are typically 10 to 30 wt percent NaCl equiv
in MVT deposits range from about 50° to 250°C; however, (Fig. 9). The compositional similarities between MVT inclu-
most of the temperatures are between 90° and 150°C (Fig. 8). sion fluids and oil-field brines are well established and have
These ranges are similar to values determined by Basuki and led to widespread acceptance of a basin-generated origin for
Spooner (2004). The highest fluid inclusion temperatures MVT fluids (e.g., White, 1958; Hanor, 1979). The high salin-
(~180° to >200°C) recorded are for the Irish Midlands and ity of sedimentary basins brines is explained by the dissolution
Gays River, ores that were relatively near each other during of evaporites, incorporation of connate bittern brines, or

0361-0128/98/000/000-00 $6.00 573


574 LEACH ET AL.

Deposit / District Deposit / District

Ozark region
Central Missouri Central Missouri
Ozark region

Tri-State

U.S.
Northern Arkansas
U.S.

Tri-State Viburnum Trend


Old Lead Belt Old Lead Belt

Mid-Cont.
Viburnum Trend

western U.S. Appalachian U.S.


Mid-cont.

Upper Mississippi Central Tennessee


U.S.

Valley
Central Tennessee
Austinville/Ivanhoe
Gays River
Appalachian

Austinville/Ivanhoe
East Tennessee
Gays River
East Tennessee

Canada -
Polaris
Pine Point Pine Point
western U.S.
Canada -

Robb Lake Robb Lake


Metaline Metaline
Nanisivik
Polaris Tynagh

Ireland
Silvermines
Tynagh Lisheen
Ireland

Silvermines Navan
Lisheen
Navan

U. Silesia
Alpine -
Upper Silesia
Africa Cevennes Upper Silesia

Bleiberg
Alpine -

Upper Silesia
Bleiberg
South Cevennes

Treves
Le Malines
Treves
Africa

Le Malines
Bushy Park
South

Bushy Park Morocco


Other

Reocin
Bou Grine
Reocin Lennard Shelf
Morocco
Other

3 6 9 12 15 18 21 24 27
Bou Grine
Salinity (NaCl wt.% eqv.)
Lennard Shelf
FIG. 9. Summary of fluid inclusion salinity data from sphalerite for se-
60 90 120 150 180 210 240 270 lected MVT deposits based on fluid inclusion studies. Vertical bar is the vi-
Homogenization Temperature (oC) sually estimated median value. U.S. = United States, mid-cont. = middle con-
tinental. A complete list of data sources is presented Appendix C (Fig. C2).
FIG. 8. Summary of homogenization temperatures for sphalerite for se-
lected MVT deposits based on fluid inclusion studies. Vertical bar is the vi-
sually estimated median value. U.S. = United States, mid-cont. = middle con-
tinental. A complete list of data sources is presented in Appendix C (Fig. C1). MVT districts are shown in Figure 10 (and Fig. C3A-E). De-
spite the fact that the data come from various laboratories
using several techniques for fluid inclusion extraction, they
through infiltration of evaporated surface waters (Hanor, are remarkably consistent and similar to present-day brines
1979). that are formed mainly from subaerially evaporated seawater
For some deposits, fluid inclusion data show distinct mixing (Kesler et al., 1996; Viets et al., 1996). Most of the data plot
trends between two different fluids (e.g., Irish Midlands, close to the seawater evaporation line (Fig. 10). Although
Upper Silesia). More commonly, the fluid inclusion data show these values do not entirely rule out the input of brines gen-
large ranges in salinity that could reflect the presence and erated by halite dissolution (Chi and Savard, 1997), they in-
variable mixing of multiple fluids during ore deposition (i.e., dicate that the contribution from such a fluid must be minor
Viburnum Trend: Leach, 1994, and references therein). except for a few districts. Fluid-rock reactions (dolomitiza-
Source of the fluids: The mole ratios of solutes measured in tion, and feldspar and clay mineral diagenesis) explain the
present-day basinal brines (e.g., Rittenhouse, 1967; Carpen- slight shift of the molar ratios away from the evaporated sea-
ter, 1978; Kharaka et al., 1987) have been used to determine water trend (Fig. C3A-E shows the Ca, Mg, and K systemat-
the source(s) of dissolved salts. These studies show that the ics for the fluid inclusion electrolyte data corresponding to
major solutes in basinal brines are derived from the subaerial Fig. 10).
evaporation of seawater or from the dissolution of evaporate Metal content of the ore fluids: The solubility of Zn and Pb
minerals (primarily halite) in the subsurface. The brine com- in chloride-rich brines at the conditions of MVT ore forma-
positions measured from fluid inclusions in sphalerite from tion are reasonably established. Although metal-bisulfide and

0361-0128/98/000/000-00 $6.00 574


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 575

1500 by the sediments in various minerals and or connate water


world MVT that was subsequently reduced by one or more processes
values outside (Sangster, 1990). The δ34S values for MVT sphalerite and
the main MVT field galena are typically less than those for seawater contempora-
neous in age with the host rocks (Sangster, 1990). Given that
1000 the isotopic compositions of oceanic sulfate varies consider-
ably with age, the range of δ34S values can be expanded de-
Cl/Br

pending on the mechanisms of sulfate reduction, and whether


Sea water the reduction occurred in an open or closed system, or
Gypsum whether the H2S was derived from organic material. Further-
500
more, the δ34S values for many districts reflect mixing of sul-
fur from different sources, isotopic fractionation as a function
Ksalts NaCl
of mineralogy, disequilibrium and kinetic effects between
MgSO4 mineral pairs, and chemical environment.
0 Given the complexities possible for sulfur isotope frac-
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 tionation, extreme variations in the δ34S values (Fig. 11) are
Na/Cl expected. Plotted in the diagram are the ranges of δ34S values
FIG. 10. Cl/Br- and Na/Br molar ratios of fluid inclusion solutes in spha-
lerite from world MVT deposits. The line is the change in the Cl/Br- and
Na/Br molar ratios of the residual fluid as seawater evaporates and precipi- Deposit / District
tates evaporite minerals. Details of the data, supplemental electrolyte data,

U.S. Ozark region


and data sources are given in Appendix C (Fig. C3). Viburnum Trend
Central Missouri
Old Lead Belt
organometallic complexes have been proposed, only metal- Tri-State
chloride complexes are considered likely base metal-transport- Appalachian Mid-Cont.
Central Tennessee
ing species for MVT ore fluids. Given the high salinity (~>10 Upper Mississippi Valley
wt %), the most important factors that control the solubility of
Pb and Zn in MVT brines are temperature, pH, and the activ-
ity of reduced sulfur. The effects of temperature and pH are Austinville/Ivanhoe
East Tennessee
considered less important than the activity of reduced sulfur Gays River
due to the low temperature of MVT ore fluids (~<200°C), high
chloride content, and limits imposed on the pH of the fluids by Pine Point
western U.S.
Canada -

the carbonate host rocks (e.g., Plumlee et al., 1994). This con- Rob Lake
Metaline
clusion is consistent with thermodynamic modeling (Fig. B10) Polaris
that shows that the dominant control on the content of Pb and Nanisivik
Zn in the MVT ore fluids is the activity of reduced sulfur (e.g.,
Anderson, 1983; Plumlee et al., 1994; Cooke et al., 2000; Tynagh
Ireland

Emsbo, 2000). Cooke et al. (2000) termed the two end-mem- Silvermines
Lisheen
ber fluid compositions (Emsbo, 2000) as either oxidized (sul- Navan
fate dominate) or reduced (reduced sulfur dominate), reflect-
ing the oxidation state of sulfur in basin brines.
U. Silesia
Alpine-

Mezica Upper Silesia


Present-day oil-field brines with significant metal content in- Bleiberg
clude brines in the Arkansas and Gulf Coast region in the United Raible
States (Carpenter et al., 1974; Kharaka et al., 1987; Moldovanyi
Africa Cevennes

and Walter, 1992; Hanor, 1996), northern Alberta (Billings et al., Villemagne
Treves
1969), and the Cheleken peninsula (Lebedev, 1973) regions. Le Malines
Present-day brines rarely have a pH below 4 and are more typi-
cally between 4.5 and 6 (e.g., Emsbo, 2000, and references
South

Pering
Bushy Park
therein). The Pb and Zn content of these brines can reach sev-
eral hundred parts per million and there is a pronounced inverse Reocin
Other

relationship with reduced sulfur content (Fig. B11). Morocco


Bou Grine
Sources of reduced sulfur: Sulfur isotope values of MVT
Lennard Shelf
sulfides throughout the world are consistent with sulfur being
-25 -20 -15 -10 -5 0 5 10 15 20 25 30
derived from a variety of crustal sources (Ohmoto and Rye,
δ 34S (‰ CDT)
1979; Sangster, 1990). Individual deposits or districts have
one or more possible sources of sulfur that may include sul- FIG. 11. Distribution of sulfur isotopes compositions for sphalerite from
fate-bearing evaporites, connate seawater, diagenetic sulfides, selected MVT deposits. If source included a numerical listing of data, the
median (vertical white line), 25th and 90th percentiles (extent of box) and total
sulfur-bearing organic material, H2S reservoir gas, and reduced range of data (horizontal bar) were calculated and plotted. If source listed
sulfur in anoxic waters of stratified basins. The ultimate only range of values, data are shown by horizontal bar only. A complete list of
source of the sulfur is most likely seawater sulfate contained data sources is presented in Appendix D (Fig. D1).

0361-0128/98/000/000-00 $6.00 575


576 LEACH ET AL.

for sphalerite from selected MVT deposits or districts and the Metal sources
visually estimated values for the median and ~ 90th percentile A fundamental but unanswered question is whether the
values of the data. Some of these distributions contain more metal-bearing fluids that formed the deposits required spe-
than one population of data representing multiple sulfur cific host-rock lithologies or rocks that contain high concen-
sources (e.g., Navan, Viburnum Trend, Upper Silesia). For trations of metals Pb and Zn. Information from Pb isotope
these districts with multiple sources of sulfur, sampling arti- data, discussed elsewhere in this paper, suggest a variety of
facts are possible (Fallick et al., 2001). Furthermore, varia- crustal sources for MVT deposits that include basement rocks
tions in isotopic values for different paragenetic stages are not of various compositions, weathered regolith, basal sand-
shown in Figure 11. For example, the δ34S values for spha- stones, and carbonate aquifers. If there is no favored lithology
lerite become more positive with younger paragenetic stages for sourcing the metals, then the temperature and composi-
for Navan (Anderson et al., 1998) but the contrary is true for tion of the fluids may have the critical controls on the extrac-
Upper Silesia (Leach et al., 1996). tion of metals from crustal materials.
Despite the variability of the data, some general observa- As noted above, the reduced sulfur content is the dominant
tions are warranted. Biogenic sulfate reduction (BSR) is com- control on the concentration of metals in MVT ore fluids,
monly invoked to explain the wide range in δ34S values in thus reduced, low sulfur basin brines will have greater poten-
MVT deposits and is the best explanation for the extreme tial to extract metals from a variety of lithologies. An impor-
negative values of δ34S observed in some deposits. Because tant control on reduced sulfur content in basin brines is the
the temperature of sulfide deposition in the deposits gener- presence of chemically reactive iron that can be sulfidized
ally exceeded conditions capable of sustaining efficient bacte- (Hunt, 1996; Emsbo, 2000). Rocks with high contents of re-
rial processes, BSR must have occurred at a different place active iron (i.e., hematitic sandstone aquifers) provide an ex-
and/or time relative to sulfide deposition. However, fluid in- cellent way to help form metalliferrous brines from a variety
clusion data suggest temperatures for some sphalerite depo- of host-rock lithologies.
sition as low as 50° to 70°C, which would permit BSR to occur
in some portion of the ore zones. Some MVT deposits had
steep thermal gradients away from ore (e.g., Upper Silesia) Precipitation processes
that would permit BSR proximal to the ore zone. Tri-State, Table B2 summarizes the more commonly proposed depo-
Bleiberg, Mezica, Les Malines, and the Irish Midlands dis- sitional processes for MVT ores. These models refer to
tricts all contain sphalerites with δ34S values characterized by whether reduced sulfur traveled with the metals or was in-
reduced sulfur produced from BSR. troduced at the depositional site. The reduced sulfur model
Reduced sulfur can be derived from the thermal decompo- requires metals and reduced sulfur to be transported to-
sition of sulfur bound in organic matter and which yields gether to the depositional site, with ore deposition by cool-
about 15 per mil fractionation of the original isotopic value of ing, mixing with dilute fluids, addition of H2S, or changes in
sulfur in the organic matter (Hunt, 1996). Reduced sulfur can pH produced by wall-rock alteration or loss of volatiles (e.g.,
be also be derived from the thermochemical reduction of sul- Anderson, 1975; Sverjensky, 1981). In order for reduced sul-
fate (TSR) by organic matter, which yields sulfur isotope frac- fur and metals to be transported together, pH values of about
tionations less than about 0 to 15 per mil (see references in 4.5 or lower are required, and high temperatures
Ohmoto and Rye, 1979) of the isotopic value of the sulfate (~200°–250°C) are also necessary (e.g., Sverjensky, 1984;
source. The rate of sulfate reduction by organic material is Cooke et al., 2000; Emsbo, 2000). The low pH required for
thought to be too slow at the relatively low temperatures of the reduced sulfur model limits MVT fluid migration to
about 80° to 150°C (Orr, 1982) to account for sufficient pro- rocks with low acid-buffering capability (i.e., siliciclastic
duction of reduced sulfur locally within the ore deposit. How- rocks) rather than carbonate aquifers and may be applicable
ever, reduced sulfur in many sedimentary basin fluids is com- to MVT deposits with a component of fluid flow in fractured
monly attributed to TSR at burial depths where temperatures siliciclastic basement rocks. The local sulfate reduction
are sufficient to overcome the kinetic constraints (Orr, 1974; model calls upon increasing the concentration of reduced
Hanor 1994; Hunt 1996). sulfur at the depositional site through sulfate reduction (Bar-
Mississippi Valley-type deposits with δ34S values lighter ton, 1967; Anderson, 1975; Beales, 1975). The most com-
than 15 per mil marine sulfate, corresponding to the ages of monly proposed mechanism involves bringing metals and
the host rocks, are commonly interpreted to have involved sulfate in the same fluid to the depositional site where
reduced sulfur derived by TSR. Such ores with large posi- methane or other organic matter reduces sulfate and precip-
tive δ34S values (e.g., Austinville, East Tennessee-Mascot itates sulfides. A variation of this process involves the ore
Jefferson City, Nanisivik, Pine Point, and the Lennard fluid carrying the reducing agent until it encounters a local
Shelf) are suggestive of TSR-derived reduced sulfur source of sulfate. The metal and reduced sulfur-mixing
(Ohmoto and Rye, 1979, p. 548). It is important to note, model calls upon the mixing of a metal-rich but reduced sul-
however, that positive isotopic values with restricted ranges fur-poor brine with a fluid rich in hydrogen sulfide at the de-
can also be produced by closed-system BSR, as well as positional site (e.g., Jackson, 1966; Anderson, 1975; Beales
through mixing of reduced sulfur from multiple sources in- and Kesler, 1996). The essential aspect of this model is that
volving more than one reduction process. Considering that any sulfur transported with the metals must be as sulfate.
both BSR and TSR processes can yield narrow distributions Corbella et al. (2004) showed that fluid mixing is an efficient
with positive isotopic values, proof of the reduction process way to make MVT ores and produce a variety of carbonate
is equivocal. dissolution features in MVT deposits.

0361-0128/98/000/000-00 $6.00 576


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 577

Ore controls produce carbonate dissolution and therefore destruction of


The ore controls recognized for selected MVT districts are possible speoleothems and other cave features.
shown in Figure 12. The ore controls can be viewed as fun- Faults and fractures: Faults and fractures are important
damental factors in fluid transmissivity, either at the district controls in most MVT districts. The faults are usually unmin-
or mine scale, that allow the focusing of fluid flow and create eralized; rather, ore is localized in dilatancy zones associated
opportunities for depositional processes to occur. Each ore- with the faults. Ores located in dilatancy zones of normal
controlling feature typically has an unmineralized counter- faults include the deposits in the Irish Midlands (e.g., Hitz-
part in each district, suggesting that the coincidence of sev- man, 1999) and Upper Silesia (Kibitlewski, 1991). Mineral-
eral controls must have been important in localizing ore. The ization is localized in dilatancy zones associated with wrench
various controls are generally interrelated (Fig. 12). For ex- faulting in the Viburnum Trend (Clendenin, 1993; Clendenin
ample, shale-depositional edges, limestone to dolostone tran- et al., 1994). Transtensional faults are important in the Ozark
sitions, and reef complexes are parts of sedimentary facies, all MVT province (Hudson, 2000) and dilatancy zones between
of which may be related to basement topography or faults. strike-slip faults are important ore controls in the Cévennes
Preore solution-collapse breccias: In nearly all MVT dis- district.
tricts, preexisting solution-collapse breccias and related car- Facies transitions: Shales and shaly carbonate units that act
bonate-dissolution features are important hosts for ore. These as aquitards within a stratigraphic sequence provide an im-
preore solution-collapse features are commonly located be- portant control on fluid migration. The relationship between
neath an unconformity or disconformity, suggesting they orig- a shale edge and location of MVT deposits in the Tri-State
inated by subaerial karstification. However, the absence of district is shown by the occurrence of ore only beyond the
speoleothems and other cave features could indicate that they subcrop edge of the Chattanooga and Northview shales, both
were simply destroyed prior to or during mineralization. For of which stratigraphically underlie the main ore-bearing car-
example, Corbella et al. (2004) proposed that the initial as- bonate units in the district (Siebenthal, 1916; Brockie et al.,
cent of mineralizing fluids into the ore zone, fluid mixing can 1968). In the Viburnum Trend, ore deposits are restricted to
a diagenetic dolostone located within a few kilometers of a fa-
cies transition to shaly limestone. Mississippi Valley-type ore
deposits typically occur near a limestone to dolostone transi-
tion (e.g., Upper Silesia, Viburnum Trend, and Old Lead Belt
S

HY

RE ON
UR D
NE

ES
S T ON E /
GE

of Southeast Missouri) that may simply reflect contrasting


CT AN
G RE N T
C OA R R IAEN D

SO ORE
B R L L A PI O N

AP
C I AS E

S
I
LEX
MP R
ED

FEA LUT
DO STO

permeability.
F R A LT S
TO SEM
CO UT

DIS E-
ALE

TU
B EF
LO

Reef and barrier complexes: The Viburnum Trend, Old


E

EC

PO

FA U

R
L
RE
LIM

DEPOSIT/
BA

P
SO

Lead Belt, Pine Point, Lennard Shelf districts, Gays River,


SH

DISTRICT
Viburnum Trend and the Gayna River deposits are examples of ore related to
Old Lead Belt carbonate reef complexes. In the Viburnum Trend, little ore
Tri-State actually occurs within reef rock; rather, ore is located in per-
North-Arkansas meable carbonate facies and sedimentary breccias that fringe
Cent. Tennessee
the original reef. The Pine Point orebodies occur in solution-
Mascot-Jefferson
City (East Tenn.) collapse breccias developed in a biohermal to bioclastic car-
Austinville bonate buildup (Rhodes et al., 1984). Both reef and carbon-
Gays River ate barrier complexes are parts of a sequence in which abrupt
Daniels Harbor changes in sedimentary facies produce dramatic permeability
Nanisivik
Polaris
contrasts.
Pine Point Basement topography: Some deposits are situated above or
Gayna near basement highs that controlled the development of sed-
Upper Mississippi imentary facies, brecciation, faulting, and pinchouts of sand-
Valley stone aquifers (e.g., Viburnum Trend and Old Lead Belt of
Alpine
the Southeast Missouri district, Pine Point, Upper Silesia, and
Robb Lake
Monarch-Kicking
Gays River).
Horse
Upper Silesia
The Irish Midlands Deposits
Lennard Shelf The carbonate-hosted Pb-Zn deposits of the Irish Midlands
Coxco have been the focus of intense debate regarding their origin,
Irish Midlands
Metaline
classification, and age of mineralization. They have been de-
Cevennes scribed as exhalative, SEDEX, synsedimentary, syndiage-
Bushy Park netic, diagenetic, epigenetic, MVT, and Irish-type deposits.
Pering “Irish-type” deposits emerged as a distinct ore classification in
Reocin the 1970s and early 1980s, at a time when many of the larger
Majo r Le ss Imp orta nt Mi nor deposits in the Irish Midlands were interpreted to be synsed-
Ore Co ntrol s i n Se l e c te d MV T D i stri c ts imentary or early diagenetic (syndiagenetic) in origin. This
FIG. 12. Summary of ore controls in selected MVT districts. Modified Irish-type classification was subsequently formalized into a
from Leach and Sangster (1993). genetic ore type by Hitzman and Large (1986), Russel and

0361-0128/98/000/000-00 $6.00 577


578 LEACH ET AL.

Skauli (1991), and Hitzman and Beaty (1996), recognizing district and the Conglomerate Group ore at the Navan de-
that some carbonate-hosted Pb-Zn deposits in the Irish Mid- posit). Perhaps the most striking feature of the Irish Midlands
lands were unequivocally typical MVT deposits (e.g., Harber- deposits is their dominant structural control (Fig. B13). Min-
ton Bridge in the Kildare subdistrict; Trude and Wilkinson, eralization is mainly concentrated along the downthrown
2001). In this paper, we use Irish Midlands for all the car- blocks of normal faults that in some areas (e.g., Lisheen and
bonate-hosted Pb-Zn ores in the region, whereas Irish type is Galmoy deposits) form a ramp-relay system (Hitzman and
reserved for a subset of deposits that some believe have a Beaty, 1996; Hitzman, 1999; Lowther et al., 1999; Fusciardi
unique origin. et al., 2003). The fault pattern in the host rocks was inherited
A detailed discussion of the Pb-Zn deposits of the Irish from Caledonian basement structures (Hitzman and Beaty,
Midlands is not presented because there is a rich literature 1996; Everett et al., 1999; Hitzman, 1999) that were reacti-
available for these deposits. The reader is referred to Andrew vated during Late Courceyan-Arundian extension as a result
et al. (1986), Hitzman and Large (1986), McArdle (1990), of the Hercynian orogeny to the south. Late Varisican-stage
Bowden et al. (1992), Andrew (1993), Anderson et al. (1995), strike-slip faulting offset the ore deposits, thus providing an
Hitzman and Beaty (1996), Johnston (1999), Blakeman et al. upper limit to the age of mineralization.
(2002), and a series of papers in Andrew et al. (1986) and Three sets of ore deposits that span the different ore types
Kelly et al. (2003). In recent years, a broad consensus has in the Irish Midlands are discussed below. These include (1)
been reached that the ores replaced carbonate rocks with the Silvermines and Tynagh deposits that have possible syn-
only a minor component interpreted to have formed on the genetic components, (2) deposits in the Rathdowney Trend
sea floor. Continuing debate mainly focuses on the precise (Lisheen and Galmoy) and Navan that are typical of the pro-
timing of mineralization, depth below the sea floor where the ductive deposits in the district, and (3) the prospects in the
ore precipitated, and the deposit attributes and genetic Kildare district that many consider typical MVT ores.
processes that some believe set them apart from MVT ores. Silvermines and Tynagh deposits: The Ballynoe barite de-
As previously noted, the authors of this paper are divided on posit at Silvermines (Fig. 3E) has long been considered the
the issue of classification of the Irish Midlands deposits. Are best example of mineral deposition by exhalative processes in
these simply MVT deposits located in the Irish Midlands or a the Irish Midlands (e.g., Wilkinson and Lee, 2003, and refer-
unique class of carbonate-hosted deposits Pb-Zn deposits? In ences therein). Features interpreted to be evidence for an ex-
Table A1 we indicate the deposits under the broad classifica- halative origin include massive collomorphic pyrite zones and
tion of MVT deposits, with an alternative classification as stratiform barite (Larter et al., 1981; Boyce et al., 1983; Mul-
“Irish-type.” lane and Kinnard, 1998), a pyritized worm tube (Boyce et al.,
Part of the controversy about the Irish deposits is rooted in 2003), and fossil vent biota and vents from the Tynagh deposit
the terminology commonly used to describe the ores (e.g., (Banks, 1986). The syngenetic environment is not universally
syndiagenetic, diagenetic, epigenetic, early and late diage- accepted, partly because the temporal and genetic link be-
netic) and the overlapping conditions inherent in these terms. tween the barite and collomorphic pyrite deposits at Ballynoe
Age determinations of the ore deposits are limited to one pa- with the replacive Pb-Zn ores at the adjacent Mogul deposits
leomagnetic study (Symons et al., 2002) and one Ar-Ar study is equivocal (e.g., Hitzman and Beaty, 1996). Some workers
of alteration of the Old Red Sandstone beneath the Silver- (Sevastupolo and Redmond, 1999) argue that the vent fauna
mines and Lisheen deposits (Hitzman et al., 1994). Because noted above are simply replacements of fossils in typical
of the limited geochronology, mineralization ages are mainly Waulsortian Limestone.
based on the relative timing of ore deposition with respect to Navan, Lisheen, and Galmoy deposits: The deposits in the
sequences of diagenetic features or events interpreted for the Rathdowney Trend, including Lisheen and Galmoy, are
host rocks. The difficulty is that these features are subject to broadly accepted as strata-bound (locally stratiform) replace-
different interpretations by different authors (e.g., the discus- ment ore deposits localized by normal faulting (Fig. 3D), with
sions of ore deposition relative to stylolite formation; Peace no evidence for synsedimentary ore deposition (e.g., Shearley
and Wallace, 2000; Reed and Wallace, 2001; Wilkinson and et al., 1996; Hitzman et al., 2003). Economic mineralization is
Lee, 2003). concentrated at the stratigraphic base of the Waulsortian
Limestone (Reed and Wallace, 2001, Wilkinson et al., 2005),
Geology of the Irish Midlands Pb-Zn deposits similar to mineralization at Silvermines and Tynagh. Fuscia-
The Pb-Zn ores of the Irish Midlands are hosted by Lower rdi et al. (2003) stated that the bulk of the mineralization at
Carboniferous (Dinantian) carbonate rocks that overlie a se- Lisheen postdates faulting, suggesting that mineralization is
quence of red beds, mainly sandstones, conglomerates, and younger than mid-Chadian. Hitzman et al. (1994) used 40Ar-
39
mudstones of the Old Red Sandstone (Fig. B12). The Old Ar dating of muscovite from hydrothermally altered Old
Red Sandstone unconformably overlies an early Paleozoic Red Sandstone and obtained an age of ~350 to 337 Ma (Cha-
basement of varied composition containing sedimentary and dian-Arundian) for the alteration that may be related to ore
volcanic rocks and intrusions. Mineralization is hosted mainly deposition.
by limestone of the Waulsortian Limestone and Navan The Navan mine (Ashton et al., 2003) is the world’s largest
Group, both of Courceyan age. The ores are mostly strata carbonate-hosted Pb-Zn deposit. The Navan deposit is hosted
bound and according to Wilkinson et al. (2003) about 98 per- in rocks of the Navan Group (Fig. B14), and ore deposition
cent of all production has been from Courceyan rocks. Cha- was localized by normal faults. Peace et al. (2003) described
dian-Arundian rocks are the youngest rocks that host miner- various ore textures at Navan that are typical of MVT deposits
alization (e.g., Harberton Bridge deposit in the Kildare (colloform, snow on the roof, and a variety of hydrothermal

0361-0128/98/000/000-00 $6.00 578


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 579

dissolution textures). Navan has the best constrained, but still deposit has temperatures near these values. However, the re-
controversial, estimates for the age of mineralization. Peace liability of the fluid inclusion temperatures for the Irish Mid-
and Wallace (2000) proposed a Holkerian age (or younger) lands deposits has been questioned (e.g., Hitzman, 1995;
based on carbonate petrology, and Symons (2002) obtained Peace, 1999; Reed and Wallace, 2001).
an Arundian/Holkerian age for mineralization using paleo- 7. Ore formation is considered to be at or close to the age
magnetic dating. The presence of mineralized clasts in the of the host rocks (~<5 m.y.), and deposition was at or near the
Chadian-Arundian Boulder Conglomerate together with min- sea floor. Although some MVT deposits are believed to have
eralization that postdate the Boulder Conglomerate are inter- formed within a few hundred meters of the sea floor (e.g., tri-
preted (e.g., Blakeman et al., 2002) to show that mineraliza- State, Upper Silesia), only the Lennard Shelf MVT deposits
tion spanned the Chadian to the Holkerian (e.g., McArdle, have such a small difference between the age of the ore and
1990; Ashton et al., 1992). The mineralized clasts in the the host rocks.
Boulder Conglomerate are consistent with some mineraliza-
tion near the sea floor. Comments on the classification of the Irish-type ores
Prospects in the Kildare district: Small low-grade deposits The controversial age of mineralization ranges from late
in the Kildare district of the Irish Midlands occur as tabular Courceyan (~352–355 Ma) for the proposed SEDEX or syn-
and strata-bound replacement ores within domal or pipelike depositional ages for Silvermines, to Chadian-Arundian-
breccia bodies and are generally considered to be MVT de- Holkerian for Navan. Mineralization is as young as Abian for
posits (e.g., Hitzman, 1986; Andrew, 1993; Hitzman and the Abbeytown deposit. To reconcile the various proposed
Beaty, 1996). The deposits are hosted in rocks of Courceyan ages, it is possible that mineralization was a relatively pro-
to Arundian age and have isotopic characteristics that are con- tracted process, spanning Chadian to Holkerian, as suggested
sidered distinct from other deposits in the district (Gallagher by Ashton et al. (2003). In addition to the age of mineraliza-
et al., 1992; Trude and Wilkinson, 2001). tion relative to the host rocks, the anomalously high fluid in-
clusion temperatures stand out as distinctly different from
Comparison of Irish-type to MVT deposit features typical MVT ores. Other than these two distinctions, the de-
The features considered most characteristic of Irish-type bate regarding the classification of the Pb-Zn deposits in the
deposits (Wilkinson, 2003) are listed below, with examples Irish Midlands as Irish type can be viewed as a philosophical
of MVT deposits bearing similar characteristics noted as choice between splitting a broad family of MVT ores into dif-
appropriate. ferent types based on the diversity of deposit attributes or
treating diversity as a feature of MVT deposits. The authors
1. The sulfides are dominated by bacteria-derived reduced defer this decision to the readers.
sulfur, considered to be the most characteristic feature of
Irish-type deposits by Wilkinson (2003, p. 980). Bacteria-de- SEDEX Deposits
rived reduced sulfur is characteristic of some MVT deposits SEDEX deposits are widely distributed but are most com-
(e.g., Tri-State, Les Malines, Mezica, Bleiberg, Raible). These mon in North America, Australia, and Asia (Fig. B15). The
MVT examples have sulfur isotope values lighter than –25 per SEDEX term was originally applied to laminated sulfide de-
mil and median values of about –8 to –15 per mil (Fig. 11), posits that were believed to have formed by the exhalation of
clearly derived from BSR. Therefore, the dominance of bac- the ore fluids onto the sea floor. However, SEDEX ore fluids
teria-derived reduced sulfur in the ores do not set the Irish are now considered to be basin brines that either exhaled
deposits apart from MVT deposits, as stated by Wilkinson et onto a basin floor (e.g., Lydon, 1983; Sangster, 1990, 2002)
al. (2005, p. 78). and/or replaced basin sediments in the shallow subsurface.
2. Ore is localized by steeply dipping normal faults that Resource data for all SEDEX deposits are included in Table
were the pathways for metal-bearing fluids. Ore control by A1. The geologic and geochemical characteristics of selected
steeply dipping faults is also characteristic of many MVT de- SEDEX deposits (excluding Broken Hill-type) are listed in
posits such as the Reocin, Upper Silesia, Cévennes, Lennard Table B3.
Shelf, Alpine deposits, Metaline, and Tri-State MVT ores.
3. Syndepositional faults are important ore controls. Corre- Genetic based subclassifications
sponding MVT examples are Reocin, Bleiberg, Moro Agudo, Although we use descriptive classifications for our data
and Pine Point. compilations, some widely used classifications include genetic
4. Ore deposition occurred by the mixing of a metal-bear- aspects. For example, a two-fold division was proposed by
ing fluid with a fluid containing reduced sulfur. This same Sangster (2002) that is based on the presence or absence of a
process has been proposed for many MVT ore deposits (e.g., discordant feeder pipe and associated alteration underneath
Southeast Missouri, Pine Point, East Tennessee). the stratiform deposits. Vent-proximal deposits have a discor-
5. Basement rocks and or detritus from the basement are dant feeder pipe and alteration complex that lies directly un-
the principle source for Pb and other metals. This relation- derneath the stratiform body, whereas vent-distal deposits
ship is also proposed for many MVT ores, including those in lack such a feature and are considered to have been spatially
the Upper Silesia, Cévennes, and Alpine districts (see discus- disconnected from hydrothermal vent zones. Vent-distal de-
sions on Pb isotopes below). posits are stratiform, concordant with enclosing strata, and
6. Temperatures of the Irish deposits are unusually hot (up display well-developed internal layering. Sangster and Hillary
to 250°C). Although some MVT deposits have a few fluid in- (1998) suggested that more than 80 percent of deposits con-
clusion temperatures that exceed 150°C, only the Gays River sidered to be SEDEX type are vent-distal deposits.

0361-0128/98/000/000-00 $6.00 579


580 LEACH ET AL.

Cooke et al. (2000) proposed that the composition of the wide variety of rock types that include shale, carbonate, and
sedimentary basin fill controls the redox nature of the basinal carbonate- and organic-rich clastic rocks such as siltstone and
brines and suggested a subclassification based on brine type: less commonly sandstone- and conglomerate-rich sequences
cool, oxidized, acidic to near-neutral pH brines, which formed (Sangster, 1990). In Table A1, SEDEX deposits are divided
what they termed McArthur-type deposits, and hotter, re- into carbonate-hosted (carb-hst), shale-hosted (sh-hst), and
duced, acidic brines which formed Selwyn-type deposits. Di- coarse clastic-hosted (cc-hst) varieties, based on the dominant
rect evidence of the chemistry of the brines responsible for lithology that hosts the ore. Figure B16 compares resource
deposition of most SEDEX deposits is lacking, so inferences data for the three dominant lithologies listed above. As a
about brine chemistry are derived from the composition of group, the subclass shale-hosted deposits have the largest
the sedimentary basin fill and/or the mineralogy of the de- tonnage of Zn + Pb deposits. The relatively large mean size of
posit. Cooke et al. (2000) proposed that most vent-proximal the shale-hosted deposits is greatly influenced by a few very
deposits formed from reduced brines (Selwyn type), whereas large deposits (Red Dog, Kholodninskoye, and Howards
most vent-distal deposits formed from oxidized brines Pass). These large shale-hosted deposits are carbonaceous
(McArthur type). (Table B3), suggesting that the presence of organic matter
Another classification was proposed by Emsbo (2000), may have been a factor in their size. As shown in Figure B16,
based largely on studies of deposits in Nevada that are inter- there appears to be no distinction in Pb or Zn grade with re-
preted to be gold-rich SEDEX deposits. According to Emsbo spect to host-rock type. However, coarse clastic-hosted de-
(2000), depending on size and relative proportions of base posits have higher mean concentrations of Ag, and shale-
metals, barite, and gold, three general deposit types exist: (1) hosted deposits contain more Cu.
very large Zn-Pb deposits representing the base metal-rich Regardless of the dominant lithology of the host rocks, lo-
end of the spectrum (HYC, Century, Mt. Isa, Sullivan, and cally derived fragmental sedimentary rocks are a common
Red Dog); (2) intermediate Zn-Pb-Ba deposits that are gen- feature of most SEDEX deposits (Lydon, 2004a). These com-
erally smaller (Rammelsberg, Tom, Jason, and deposits in the prise debris flows, stratiform and discordant breccias, con-
Qinling belt of China) that typically contain high gold con- glomerates, and mud flows and are usually composed of rock
tents; and (3) barite deposits that have a low content of base types which occur at the same or deeper stratigraphic levels
metal sulfides but are Au rich, such as those in Nevada as the SEDEX deposit itself.
(Emsbo, 2000). Historically, gold has been considered to be
unimportant in SEDEX deposits, but the concept of gold-rich Grade and tonnage
end members is intriguing and warrants further study and Globally, SEDEX deposits constitute a very important re-
consideration. source, comprising more than 50 percent of the world’s Zn
Although any of the above classifications may be broadly and Pb reserves and more than 25 percent of the world’s Zn
applicable to most SEDEX deposits, there are always signifi- and Pb production (Goodfellow, 2004; Table A1). Based on
cant exceptions. For example, the classification by Cooke et data compiled and included in Appendix A, there are 25
al. (2000) suggested that the brines responsible for producing SEDEX deposits worldwide (out of 130 deposits in the data-
Selwyn-type deposits with abundant barite (e.g., Selwyn basin base, excluding Broken Hill-type deposits) that are 50 Mt or
deposits, Red Dog) are hot, reduced, and acidic, based on greater in size; 13 of these contain 10 Mt or more of com-
thermodynamic modeling and the assumption that Ba, Pb, bined Zn + Pb (Figs. 13, 14).
and Zn are carried together in the same fluid. However, re- Figure 14 shows an excellent correlation between total
cent data from the Red Dog deposit show most barite was de- tonnage and contained metal (Pb + Zn), with a fairly consis-
posited before the main sulfide stage, suggesting that the tent ratio of about 10:1, or average grade of 10 percent, re-
fluid that deposited barite was not necessarily the same as gardless of the size of the deposit or district. This is consis-
that responsible for sulfide deposition (Kelley et al., 2004b). tent with trends shown by MVT deposits and may indicate
Difficulty also arises in applying the vent-proximal versus that the ore fluids for both ore types were saturated (or near
vent-distal classification due to the common inability or diffi- saturated) with respect to both Pb and Zn, as indicated from
culty in identifying a vent zone. Some vent zones may not be solubility relationships shown in Figure B10. However, the
exposed due to postmineralization structural disruption controls on the amount of metal in the deposit could reflect
and/or overprinting by later hydrothermal or metamorphic the total volume of ore fluid that produced the deposit or the
events. Second, the evidence cited for the presence of a vent metal content of the ore fluid, the latter dependent mainly
zone is commonly ambiguous. For example, the presence of on the temperature, salinity, and reduced sulfur content of
veins is commonly cited as evidence for a vent zone in many the ore fluid. Given that SEDEX and MVT ore fluids are
deposits. However, as Sangster and Hillary (1998) pointed probably similar, insights into the correlation of total tonnage
out, footwall alteration, in conjunction with the veins, is re- and contained metal for SEDEX and MVT deposits noted
quired in order to classify deposits as vent proximal. And last, above are possible from the the fluid inclusion summary by
the vent-distal classification does not consider deposits that Basuki and Spooner (2002). They noted that the tempera-
formed from subsea-floor replacement of preexisting sedi- tures of MVT ore fluids do not correlate with deposit grade
ments that do not require the venting of ore fluids. or tonnage (Basuki and Spooner, 2002), whereas salinities
seem to have a more complex relationship. Perhaps the
Host-rock lithology amount of reduced sulfur in the ore fluids may be the most
Originally regarded as being more or less confined to black important control on the metal content, as suggested by
shales, SEDEX deposits are recognized to be hosted by a Hanor (1996).

0361-0128/98/000/000-00 $6.00 580


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 581

35 aspect ratio of 20 or more comprise sheets of stratiform sul-


fides (Fig. 15E) up to a few tens of meters in thickness and
32.5 hundreds to thousands of meters in lateral extent (e.g., Cen-
30
tury: Broadbent et al., 1998; HYC: Large et al., 2005). In
some cases, the deposits contain more than ten stacked lenses
27.5 (Mt. Isa, George Fisher, Aguilar, Santa Lucia; Table B3; Fig.
15F), but in such cases, most of the ore is contained in only a
25 few lenses.
Some deposits have been interpreted as having a well-de-
22.5
veloped feeder zone that underlies or is adjacent to bedded
20 ores; these include Sullivan, Rammelsberg, Tom, and Jason
Pb + Zn (Mt)

(Table B3; Fig. 15C, D). Typically the feeder zones are rooted
17.5 in a synsedimentary fault (Large, 1983; Goodfellow et al.,
1993) and consist of brecciated and altered sedimentary rocks
15
overprinted by veins and vein networks composed principally
12.5 of quartz, Fe-Mn carbonates, and sulfides. Other vein and al-
teration minerals reported in feeder zones are barite, mus-
10 covite, chlorite, and tourmaline (Table B3). Many of these de-
posits have a complex sulfide-altered zone, interpreted as a
7.5 “vent complex” overlying the feeder zone (Goodfellow et al.,
5
1993). Unlike volcanic-hosted massive sulfide (VHMS) de-
posits, the feeder zones in SEDEX deposits contain fewer
2.5 sulfides, more carbonate, and are rarely chalcopyrite rich. Ex-
ceptions to this are the Cu-bearing stockwork zones at Ram-
0 melsberg (termed the Kniest) and Mt. Isa (silica-dolomite).
Sullivan
Mount Isa

Dongshengmiao

Cirque
Hilton

Changba-Lijiagou

Zinkgruvan
Kholodninskoye

Arditurri

Meggen
Faro

Balmat
Lik
Howards Pass

Dairi
HYC

Rampura-Agucha

Anarraaq
Red Dog

Rammelsberg

Lady Loretta
Gorevsk

Filizchai

Tekeli
Dugald River
George Fisher

Zawarmala
Century

Aguilar
Rajpura-Dariba

In the latter case, the Cu mineralization is suggested to be


significantly later than the stratiform Zn-Pb mineralization
(Large et al., 2005).
The large deposits in north Australia (Century, HYC, Lady
Loretta, Hilton, and George Fisher) occur as a series of thin
FIG. 13. Bar diagram showing the largest 30 SEDEX deposits based on stacked sulfide-rich sheets with intervening unaltered sedi-
total Pb + Zn content (Mt). Data are listed in Table A1. mentary rocks (Fig. 15E, F) and are close to major faults in-
terpreted to be the conduits for ore fluids. The absence of
footwall alteration zones in the faults is interpreted as evi-
dence that the deposits formed distal from the vent complex
Deposit morphology (Sangster and Hillary, 1998; Large et al., 2002, 2005).
SEDEX deposits are commonly described in terms of their The deposit morphology largely reflects the depositional
aspect ratio—i.e., the ratio of the lateral extent of the body to environment (Fig. 16). For example, mound or lens-shaped
their maximum stratigraphic thickness (Lydon, 1996). Some deposits with low aspect ratios are those that are classified as
SEDEX deposits have a low aspect ratio (e.g., Red Dog, vent proximal (Fig. 16A) or they may form partly by replace-
Anarraaq) and are lens or wedge shaped with a relatively flat ment of a preexisting mound- or lens-shaped barite body, as
top and a convex base (Fig. 15A-C). SEDEX deposits with an in the case of the Red Dog deposits (Fig. 16E). Other de-
posits that typically comprise stacked elongate lenses are
thought to have formed in stratified brine pools, (e.g., HYC
100 deposit: Large et al., 2001, 2002) or in a preexisting hydro-
carbon source reservoir (e.g., Century: Broadbent et al., 1998,
2002) (Fig. 16B, C). Elongate orebodies that are layered or
lens shaped may also result from replacement of carbonate
Pb+Zn (Mt)

10 layers within shale or mudstone (Fig. 16D).


Mineralogy and zoning
The most important primary minerals in SEDEX deposits
1 are sulfides, carbonates, barite, and quartz (including chert
and siliceous shale), although proportions of these vary signif-
icantly (Table B3). Pyrite is generally the iron sulfide mineral
present in the deposits and its abundance can vary from rare
.1 or minor (e.g., Red Dog, Howards Pass, Century) to compris-
1 10 100 1000
Size (Mt) ing the most abundant sulfide (e.g., Cirque, Meggen, HYC).
FIG. 14. Size (Mt) of SEDEX deposits vs. total metal (Pb + Zn) content In some Proterozoic deposits (Mt. Isa, Sullivan), pyrrhotite is
(Mt). Data are listed in Table A1. common and can exceed pyrite abundance in certain parts of

0361-0128/98/000/000-00 $6.00 581


582 LEACH ET AL.

A) Aqqaluk deposit (Red Dog) B) Anarraaq deposit


W E NW SE

?
?
?

PERMIAN-CRETACEOUS MISSISSIPPIAN
Hangingwall mudstone, Barite Ore Melange
PERMIAN-TRIASSIC MISSISSIPPIAN
siltstone, shale, and chert
Massive Ore Thrust fault Hangingwall shale Carbonaceous shale; Fe-rich sulfides
MISSISSIPPIAN Vein Ore and chert unmineralized

60 m
Barite Thrust fault
100 m MISSISSIPPIAN Massive sulfides

60 m
Carbonaceous shale; Footwall
unmineralized calcareous shale Barite Banded sulfides

C) Sullivan deposit D) Jason deposit


W E
SURFACE
VENT COMPLEX

TRANSITION ZONE
100 m

BEDDED ORE
1000 m

Jason fault
100 m
MIDDLE/LATE DEVONIAN SILURIAN
Carbonaceous shale Sandstone/conglomerate Silty limestone
Bedded sulfides Gabbro Turbidites Sedimentary breccia Zn-Pb mineralized zone Fault
Transition zone Chaotic breccia Stratigraphic Siltstone turbidites EARLY/MIDDLE DEVONIAN
Massive sulfides Marker units
Unconformity
Conglomerate Conglomerate Chert
Fault

E) Century deposit F) George Fisher deposit


W E

W E Undifferentiated Urquhart
Shale
Siltstone
Mineralized interval: Zn +/-
Pb-bearing rhythmically lamina-
Pandora fault ted pyritic siltstones intercala-
ted with banded mudstones

Cambrian limestone Footwall shale-siltstone Mineralized Barren mudstone


Proterozoic weathering surface Interval
Unit 1 Major post-D3 fault
Hangingwall sandstone Footwall carbonaceous
Unit 2
shale
Hangingwall siltstone-shale Unit 3
Unit 4

0 500 m Fault 0 100 m

FIG. 15. Generalized cross sections showing the diversity in morphology of SEDEX deposits. A. Aqqaluk deposit, one of
four deposits at Red Dog (from Kelley et al., 2004b). B. Anarraaq deposit (from Kelley et al., 2004a). C. Sullivan deposit
(from Lydon, 2004a). D. Jason deposit (from Goodfellow et al., 1993). E. Century deposit (from Broadbent et al., 2002). F.
George Fisher deposit (from Chapman, 2004).

the deposit (Large et al., 2005). At the Duddar deposit in Pak- (Lydon, 2004b; Large et al., 2005), most SEDEX deposits
istan, marcasite is the dominant iron sulfide (Jankovic, 1986). are Zn rich relative to Pb and have Zn/(Zn + Pb) ratios that
The main Pb-Zn minerals are sphalerite and galena. Al- average about 0.7 (Table A1). Silver is reported for 66 de-
though the Mt. Isa and Sullivan deposits contain near-equal posits (median value is 46 g/t). Sulfosalts such as tetrahedrite,
proportions of Pb and Zn in some parts of the deposits freibergite, and boulangerite are typically concentrated

0361-0128/98/000/000-00 $6.00 582


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 583

A. Vent-proximal B. Vent-distal
(e.g. Sullivan, Lydon 1996; Goodfellow et al. 1993) (e.g. HYC, Large et al. 2001)

SL SL
oxic sw oxic sw
plume mixed layer
anoxic sw
stratified brine pool, anoxic

massive hydrothermal
ore products
feeder
bedded
zone
ore

C. Replacement of oil trap D. Carbonate replacement


(e.g. Century, Broadbent et al. 1998) (e.g. Anarraaq; Kelley et al., 2004a)

dolomite breccia
siderite alteration alteration

Zn-Pb
lenses

E. Barite replacement
(e.g. Red Dog. Kelley et al., 2004b)

S sulfides: disseminated and interstitial to coarse barite barite N

Veins

Black carbonaceous and siliceous mudstone

Calcareous shale

FIG. 16. Depositional models for stratiform Zn-Pb-Ag deposits. A. Vent-proximal deposits (i.e., Sullivan, Canada). B. Vent-
distal deposits (i.e., HYC, Australia). C. Replacement of oil trap (i.e., Century, Australia). D. Replacement of carbonate units
Anarraaq, Alaska). E. Replacement of barite. Modified from Large et al. (2002).

near-vent complexes and are the main host of silver in some Meggen, Anarraaq). Barite may be peripheral to or strati-
deposits (Lydon, 1996). Chalcopyrite can be a minor (0.1–0.5 graphically above the deposit (e.g., Anarraaq), or it may form
wt %) constituent of some deposits (Red Dog, HYC, Meggen) crudely segregated mixtures with sulfide minerals (Red Dog,
but is rarely a significant component (>2 wt %; Rammelsberg, Rammelsberg, Jason), but many deposits have no associated
Rajpura-Dariba). barite (Table B3).
Carbonate minerals include calcite, siderite, dolomite, and Quartz (or silica as chert or siliceous shale) is listed as a
ankerite. Manganese-bearing carbonates are common in common gangue mineral in many deposits (e.g., Rammels-
many deposits (e.g., Century, Rammelsberg, Lady Loretta, berg: Red Dog; Moore et al., 1986; Large and Walcher, 1999).

0361-0128/98/000/000-00 $6.00 583


584 LEACH ET AL.

It occurs as irregular veins and patches in what is interpreted 1 2 3


as the vent complex in many of the Selwyn basin deposits Massive, brecciated, crudely
bedded pyrite, galena,
Laminated, and
occasionally fragmented
Chert, barite +/- minor
sphalerite, pyrite, Ca-
(Goodfellow et al., 1993; Goodfellow, 2004). Numerous de- sphalerite, Fe-and Fe-Mg-Ca pyrite, sphalerite, galena, Mg carbonates,
chert, barite, Fe-Mg magnetite,hematite,
carbonates, barite (at top)+/-
posits contain quartz, as chert, in the layered or bedded ore pyrrhotite, sulfosalts carbonates +/- beds of
host rock types
Mn enrichment
zones (Geer, 1988; Turner, 1990; Goodfellow, 2004). The Ore grade decreases
Fe:Zn increases
timing of quartz deposition is assumed to be coincident with Pb:Zn decreases Fe:Pb increases
Zn:Pb increases
sulfides, but some quartz may form earlier or later than the Ba:Zn increases Zn:Ba decreases
Zn:Mn decreases
ore-stage sulfides. For example, veins along the perimeter of Thickness decreases
the main deposit at Century are paragenetically late, as indi-
cated by Pb isotope data (Broadbent et al., 2002), and Ans-
1 2 3
dell et al. (1989) suggested that late quartz veins that cut the SEDIMENTARY
main ore zones in the Tom deposit formed during postore FRAGMENTAL
LITHOLOGIES
deformation and metamorphism. Quartz, including chert (Breccia flows, average scale:
talus breccias,
and siliceous shale, are unusually abundant in the Red Dog conglomerates, 100 m
etc.)
deposits compared to other SEDEX deposits. Some silica SYNSEDIMENTARY Veins and replacement
20 m
deposition predates sulfide deposition, some formed concur- FAULT ZONE galena, sphalerite, Fe- and Fe-Mg-Ca
carbonates, pyrite, pyrrhotite,
rent with sulfide deposition (Slack et al., 2004b), and some minor chalcopyrite
formed during burial 150 m.y. after sulfide deposition (Leach FIG. 17. Cross section through an idealized SEDEX deposit showing geo-
et al., 2004). chemical zoning (modified from Lydon, 2004a).
Apatite (fluorapatite) is a common constituent in some de-
posits (Anarraaq, Red Dog, Howards Pass). In the Howards
Pass deposit, there is an inverse relationship between calcite of Fe/Pb and Fe/Zn also generally increase with increasing
and apatite, with calcite decreasing and apatite increasing distance from the vent complex (Fig. 17).
from the base to the top of the mineralized zone (Goodfellow, Vertical metal zonation is also common. Large et al. (2005)
2004). In the Red Dog and Anarraaq deposits, the rare earth show that on a deposit scale, the average Zn concentration
element (REE) patterns of apatite assumed to be temporally (and generally Zn/Pb ratio) decreases upward in many of the
associated with the sulfides differ from those of sedimentary Australian deposits (e.g., HYC, Century, George Fisher). This
apatite in host rocks away from known mineralization. The is consistent with the Red Dog deposits where the Zn/Pb ratio
former have positive Eu anomalies compared to the latter decreases from about 4.5 at the base of the deposit to about 3
(Slack et. al., 2004b). Positive Eu anomalies are also charac- at the top, accompanied by a generally increasing Zn/Fe ratio
teristic of apatite in the Gamsberg deposit (Stalder and (Kelley et al., 2004b). The vertical geochemical variations in
Rozendaal, 2004). Fluorite has been reported as a rare trace SEDEX deposits may reflect changes in the overall abun-
mineral at the Drenchwater deposit in Alaska (Werdon, 1996) dances of ore minerals (relative proportions of sphalerite and
and the Gunga deposit in Pakistan (Jankovic, 1986), and galena), or alternatively, they could reflect the chemical com-
molybdenite occurs in trace amounts in the Howards Pass de- position of the sulfides. For example, the Fe content in spha-
posits (Goodfellow, 2004). lerite in some deposits (e.g., Howards Pass; Goodfellow,
Iron sulfide or sulfosalts are the main mineralogical hosts 2004) generally decreases upward.
for As, Sb, and Tl (Kelley et al., 2004b), although arsenopyrite
is present in some deposits (e.g., Tom; Goodfellow, 2004). Alteration
Sphalerite is the mineral host for Cd and Hg, although trace The style of alteration and halo development related to
amounts of cinnabar have been reported in a few deposits SEDEX deposits depends on the composition and perme-
(Large, 1983). Selenium is enriched in some ores and/or al- ability plus porosity characteristics of the host sediments. Al-
tered wall rocks (hundreds of ppm), but the mineralogical though the intensity of alteration in SEDEX systems is com-
residence is poorly known. monly much less than VMS systems, the extent of the halos
Mineralogical zoning is common in many deposits (Fig. 17). along the favorable stratigraphy may be far greater. Based on
Lydon (2004a) described a zonation from reduced mineral fa- detailed studies of selected deposits, two broad types of alter-
cies (sulfides, ferroan carbonates) within the upflow zone to ation patterns are apparent.
more oxidized facies (barite, iron oxides, calcic carbonates) at Fe-Mn carbonate alteration halos: This type of alteration is
the periphery. This is particularly well demonstrated for some best developed in the dolomitic siltstone-hosted deposits of
of the Selwyn basin deposits (e.g., Jason; Goodfellow, 2004). northern Australia (Lady Loretta: Large and McGoldrick,
This trend is reflected in decreasing ratios of Zn/Ba and 1998; HYC: Large et al., 2000; Century: Broadbent et al.,
Zn/Mn (Fig. 17; Lydon, 2004a) away from the upflow zone. 2002; George Fisher: Chapman, 2004). The halos are thickest
There is also a common increase in Zn/Pb ratios outward surrounding the ore and extend along stratigraphy for dis-
from the vent complex and this has been well documented for tances of a few hundred meters to tens of kilometers (Fig.
the Tom (Goodfellow and Rhodes, 1990), Jason (Turner, 18). The Fe and Mn contents of sedimentary and hydrother-
1990), Cirque (Jefferson et al., 1983), and Sullivan (Hamilton mal carbonates commonly decrease away from ore. At Lady
et al., 1982) deposits. Other element ratios including Pb/Ag, Loretta, the carbonate zonation away from the sulfide deposit
Cu/(Zn + Pb), and SiO2/Zn ratios decrease away from the is Mn siderite to Mn ankerite to ferroan dolomite to dolomite.
vent complex (Goodfellow, 2004) in the Selwyn deposits. In Century shows a similar Mn siderite and ankerite envelope
deposits in which most of the Fe is contained in pyrite, ratios (Large et al., 2005; Fig. 7). Siderite is missing from the

0361-0128/98/000/000-00 $6.00 584


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 585

ore
lenses

Mn
weak
Mn, TI Mn
Carbonate zones
hydrothermal
siderite + pyrite
100 m

contact between overlying shale fluids


Mn,Tl

ankerite + pyrite
and underlying dolomite
Mn-Fe-dolomite fault
1km
dolomite

FIG. 18. Schematic Fe-Mn carbonate alteration halos (from Large et al., 2002b).

extensive halo at HYC and the highest concentrations of Mn silicification and sulfide deposition is controversial because
in dolomitic siltstone are in the immediate footwall of the de- textural relations between quartz and sulfides are ambiguous
posit, where values of up to 1.7 wt percent MnO in whole owing to the commonly fine-grained nature of the silica min-
rock, equivalent to 12 wt percent MnO in dolomite, have eral. Paragenetic studies by Broadbent et al. (1998) suggest
been measured (Large et al., 2000). Chemical changes in sed- that deposition of base metal sulfides postdates the silica
iments within the carbonate alteration halo involve addition event at the Century deposit. Slack et al. (2004b) suggest that
of FeO and MnO and depletion of MgO, CaO, and rarely the silicification precedes or overlaps with the early main ore
Na2O (Large and McGoldrick, 1998). stage at Red Dog, whereas Leach et al. (2004) provide an al-
Other SEDEX deposits hosted by carbonate-bearing sedi- ternative postore explanation for the majority of quartz in the
ments show some aspects of the Fe-Mn halo zones described deposit.
above. At Meggen, Gwosdz and Krebs (1977) defined a Mn Trace element halos: Within the siderite and ankerite halos
halo in the ore-equivalent limestone that extends for over 5 of some deposits, metals are enriched to values of 1,000 ppm
km from the edges of the orebody. Turner (1990) and Good- to 1 wt percent Zn and 100 to 1,000 ppm Pb (Lambert and
fellow (2004) recorded the common occurrence of Fe-rich Scott, 1973; Large and McGoldrick, 1998). The metal enrich-
carbonate alteration at the Tom and Jason deposits, particu- ment occurs in the immediate hanging-wall sediments, up to
larly in the footwall calcareous sediments adjacent to the several hundred meters, and along the favorable horizon for
feeder zones and extending laterally away from these zones hundreds to thousands of meters. At HYC, Zn-Pb enrichment
along permeable units. In the Red Dog district, some shales extends for at least 15 km from ore along the favorable hori-
interbedded with barite in the hangingwall of the deposits zon (Large et al., 2000). This compares with Zn-Pb dispersion
contain sparse to abundant MnO (up to 8.7 wt %) but it is un- of 500 m from ore at Rammelsberg (Walcher, 1986).
certain if the Mn-rich shales are alteration products of the In addition to Mn, Tl is an important geochemical indicator
ore-forming event (Slack et al., 2004b). and vector to mineralization. In the Australian deposits, Tl oc-
Silicate alteration halos: Silicate alteration has been curs at values of 100 to 1,000 ppm through the Zn-Pb ores
recorded in some SEDEX deposits hosted by siliciclastic sed- and decreases to values of less than 1 ppm at distances of 1 to
iments. At Sullivan, tourmaline alteration is concentrated in a 20 km along the favorable horizon (Large et al., 2000). Thal-
pipe-like zone below the thickest part of the deposit (Fig. lium can be a geochemical indicator for exploration of
B17). Minor chlorite-pyrrhotite alteration occurs lateral to SEDEX deposits in Alaska (Slack et al., 2004b), although due
the tourmaline pipe, and the bedded ores are underlain by to postore structural complexities, it is not possible to track
muscovite-altered sediments. Hanging-wall alteration is char- the anomalies for great distances along favorable horizons.
acterized by a zone of chlorite-albite-pyrite alteration extend- On a regional scale, the concentrations of Tl in black shale
ing up to 200 m above the deposit, surrounded by an exten- away from known mineralization (>1 km) range from <0.1 to
sive zone of muscovite alteration (Fig. B17). Near the 1.1 ppm (Slack et al., 2004a), whereas the ores contain as
transition between bedded sulfides and the vent complex, much as 315 ppm, and siliceous black shale in and around the
there is abundant carbonate alteration. Similar envelopes of deposits in the Red Dog district contains from 60 to 110 ppm
muscovite alteration occur within the non-calcareous sedi- Tl (Slack et al., 2004b). In situ trace element analyses (LA-
ments that host the deposits in the Anvil camp (Carne and ICP-MS) of pyrite show similar enrichments, with pyrite near
Cathro, 1982; Shanks et al., 1987). the deposits having significantly greater Tl (up to 12,220
Silicification in the form of quartz is reported as an alter- ppm) and Ag (up to 900 ppm) than pyrite in black shale dis-
ation phase for some SEDEX deposits. For example, silicified tal (>1 km) from known mineralization (<3.3 ppm Tl and <12
wall rocks are known at the Tom, Jason, Vangorda, Rammels- ppm Ag; Slack et al., 2004b).
berg and Century deposits (Broadbent et al., 1998) and sev- Other elements that may show dispersion in sediments sur-
eral other deposits (Slack et al., 2004b, and references therein). rounding SEDEX deposits include As, Ba, Bi, Ge, Hg, Ni, P,
For the deposits that contain added silica, the timing between and Sb (Table B3). Although REE data are available for only

0361-0128/98/000/000-00 $6.00 585


586 LEACH ET AL.

a few deposits, studies to date show that generally positive Eu include fine-grained layered and banded textures (Fig. 19A-
anomalies characterize ore and adjacent wall rocks in some H), with or without coarser grained brecciated, veined, frag-
deposits (Red Dog, Anarraaq: Slack et al., 2004b; Dengjiasha mental, or chaotic textures (Fig. 19H-P). The highly variable
deposit in the Oinling belt of China: Ma et al., 2004). Other grain size is generally controlled by primary sulfide mineral-
deposits exhibiting similar trace element anomalies include ogy and the extent of hydrothermal coarsening and/or recrys-
the Drenchwater prospect in the central Brooks Range (Slack tallization during diagenesis and burial metamorphism.
et al., 2004b), bedded and massive sulfide ores of the Sullivan A common texture in many deposits is bedding-parallel lay-
deposit in British Columbia (Slack et al., 2000), the Meggen ers of base metal sulfides or other hydrothermal products that
and Rammelsberg deposits in Germany (Geer, 1988; Large alternate with beds of the host rock types (Fig. 19A-H). Indi-
and Walcher, 1999), and the Howards Pass and Tom deposits vidual layers of sulfides range in thickness from millimeters to
in Yukon Territory (Goodfellow et al., 1983; Goodfellow and several tens of centimeters. Bedding contacts are commonly
Rhodes, 1990). sharp, particularly between layers of sulfide minerals and host
Isotopic halos: Recent research at HYC and Lady Loretta rock types. In some unmetamorphosed examples (e.g., Cen-
(McGoldrick et al., 1998; Large et al., 2001) indicates that ex- tury: Fig. 19F; HYC: Fig. 19G, H; Anarraaq: Fig. 19D), the
tensive variations in carbon and oxygen isotopes exist around sulfides are very fine grained, with individual sphalerite grains
the Mn-Tl and Fe carbonate lithogeochemical halos. less than 0.5 µm. In contrast to the regularly layered textures
Dolomite within the siltstones for at least 15 km along strike common in some deposits, many deposits contain textures
from the HYC deposit exhibits 18O-enriched (δ18O = that are much more chaotic and varied. Large zones in some
23–26‰) and 13C-depleted (δ13C = –2 to –3.5‰) isotopic deposits are composed of massive and/or replacement zones,
signatures compared to typical sedimentary dolomites of the breccias, irregular veins, nodular textures and/or dissemi-
Barney Creek Formation from the region (δ18O = 20–23‰ nated sulfides, barite, or carbonates (Fig. 19I-P). Such zones
and δ13C = 0 to –2‰). A similar isotopic trend has been are usually interpreted as having formed by the reaction of
recorded in carbonates from the Sullivan deposit, where Tay- upflowing hydrothermal fluids with earlier hydrothermal
lor et al. (2000) reported a 3 per mil increase in δ18O and 2 minerals and host sediment (Large et al., 2002; Goodfellow,
per mil decrease in δ13C of calcite from the footwall slates to 2004).
the “Main band” bedded ore. Goodfellow (2004) reported Synsedimentary versus replacement textures: Textural evi-
negative δ13C values (–3 to –9‰) for altered siliceous lime- dence commonly used for synsedimentary sulfides includes
stone, with a trend to less negative values from the base to the the following: (1) sulfide-rich bands that are intricately inter-
top of the mineralized zone in the Howards Pass deposit. layered with host sediments and which have sharp boundaries
Large et al. (2001) suggested two possible mechanisms for between sulfides and unmineralized host-sediment bands
producing the isotopic patterns: lateral flow of hydrothermal (e.g., Large et al., 1998); however, these textures may be pro-
fluids through porous strata in the subsea floor (replacement duced as a result of sulfide replacement of carbonate layers
model), and regionally extensive, warm, stratified brine pools within mudstone, for example, at the Anarraaq deposit (Fig.
(brine pool or exhalative model). Large et al. (2001) suggests 19D; Kelley et al., 2004a); (2) compaction load-cast features
the brine-pool model is more compatible with the great lat- that are described by Large et al. (1998) for the HYC deposit
eral extent of the halo and the uniformity of the isotopic com- (Fig. 19H; (3) microbreccia layers and rounded and sub-
positions within the halo. rounded sulfide clasts in graded sediments interpreted as re-
Preliminary studies on the strontium isotope compositions worked sulfide layers deposited on the basin floor (Moore et
of barite and carbonates associated with SEDEX deposits al., 1986; Large et al., 1998; Fig. 19H, N); and (4) beds that
commonly have radiogenic Sr compositions (Fig. B18). Good- are composed of multiple sulfide layers suggesting a sedi-
fellow (2000) showed that 87Sr/86Sr in barite and Zn-Pb de- mentary origin and deposition from a density flow (e.g., Jason:
posits from the Selwyn basin have strongly radiogenic values, Goodfellow, 2004; Sullivan: Lydon, 2004b).
up to 0.718 compared to seawater values of around 0.708 Evidence cited to indicate subsea-floor replacement in-
(Fig. B18). At HYC, the orebody and surrounding ankeritic cludes (1) carbonate or barite layers or nodules that have
shales display an Sr isotope halo of similar dimension to the been partly or wholly replaced by pyrite or base metal sulfides
carbon and oxygen isotope halo. Initial 87Sr/86Sr ratios range (e.g., part of the HYC deposit: Large et al., 1998; Anarraaq
from about 0.720 to 0.750, compared to normal marine values deposit: Kelley et al., 2004a; Fig. 19O), (2) base metal sulfides
in dolomites beyond the halo of 0.706 (Large et al., 2001). replacing earlier pyrite (e.g., Meggen: Geer, 1988; Anarraaq:
Barite from deposits in the Red Dog district has a total range Kelley et al., 2004a), (3) base metal sulfides replacing fossils
in 87Sr/86Sr from 0.709 to 0.710 (Ayuso et al., 2004; Fig. B18). in carbonate (Anarraaq: Kelley et al., 2004a), (4) pseudo-
Carbonate (87Sr/86Sr = 0.710–0.714) and witherite (87Sr/86Sr = morphs of barite or other minerals, now filled with sulfides
0.711) are more radiogenic than barite. The radiogenic Sr iso- (e.g., Red Dog: Kelley et al., 2004b; Fig. 19K), or (5) sulfides
tope ratios of barite and carbonate minerals in the deposits replacing glass shards in mass flow and tuff units (HYC: El-
suggest the ore fluid pathway involved extensive interaction dridge et al., 1993). However, all of these features can also
with evolved rocks deep in the basin or within the basement. occur during subsea-floor hydrothermal replacement below
an exhalative orebody.
Ore textures
Photographs of hand samples from selected deposits (Fig. Tectonic setting
19) illustrate the highly variable textures of SEDEX deposits. SEDEX deposits occur in two broad settings: intraconti-
Nearly all SEDEX deposits described in the literature nental or intracratonic failed rifts, and rifted Atlantic-type

0361-0128/98/000/000-00 $6.00 586


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 587

A B C

2cm 2cm

D E F

1cm
1cm

G H I

4cm 1cm 1 cm

FIG. 19. Hand samples of selected SEDEX deposits showing diversity of textures. A. Intensely folded delicate lamina-
tions of sphalerite, galena, and pyrite with pale gray cherty limestone, and dark gray carbonaceous chert in a sample from
the Active Member, lower unit, Howards Pass (XY) deposit, Yukon Territory, Canada. From W.D. Goodfellow. B. Layered
ore typical of the upper two thirds of the Main Band and overlying A, B, C, and D bands, Sullivan deposit, British Colum-
bia, Canada. Layers of galena, sphalerite, pyrrhotite, and argillite are intercalated. From K.R. McClay. C. Interbedded semi-
layered white barite, cream sphalerite, and dark gray mudstone from the Tom deposit. Galena occurs finely disseminated in
sphalerite layers. From W.D. Goodfellow. D. Banded sulfides from the Anarraaq deposit, northern Alaska, United States.
Light tan bands are sphalerite that alternate with iron sulfides. Thin section petrography reveals pyrite and sphalerite re-
placed preexisting radiolarian-bearing carbonate turbidites (Kelley et al., 2004a). E. Layered sulfides alternating with
siliceous shale/mudstone from the Rammelsberg deposit, Germany. F. Very fine layering of sulfides (tan) in calcareous and
carbonaceous shale from the Century mine, Australia. Based partly on detailed thin section analyses, Broadbent et al. (1998,
2002) interpret the textures to be replacement of carbonate (siderite) by sulfides. G. Fine layering of sulfides with carbon-
ate (dolomitic siltstone) from the HYC deposit, McArthur basin, Australia. H. Sedimentary breccia and laminated sphalerite-
galena ore from the HYC deposit. Note the lode cast contact relationship between the base of the graded breccia bed and
the laminated Zn-Pb siltstone layer, indicative of synsedimentary deposition of the ores. I. Sulfide-bearing barite from the
Aqqaluk deposit (Red Dog), northern Alaska, United States. White coarse barite and interstitial brown sphalerite and pyrite.
Kelley et al. (2004b) interpreted the textures to indicate coarsening of preexisting fine-grained barite by the introduction of
fluids that deposited sphalerite.

0361-0128/98/000/000-00 $6.00 587


588 LEACH ET AL.

J K L

psuedomorph
of barite

1cm 1cm 1cm

M N

1cm 1cm

O P

sphalerite

pyrite

1cm 1cm

FIG. 19. (Cont.) J. Massive sulfide ore from the Main deposit (Red Dog). Sphalerite, galena, and pyrite are disseminated
in black siliceous matrix. Note fragmental character of ore. K. Massive sulfide ore from the Main deposit (Red Dog), show-
ing psuedomorphs of barite, now filled with pyrite and sphalerite. L. Planar veins from the Main deposit (Red Dog). Note
the multiple episodes of veining. M. Black silicified argillite cut by veins of ferroan carbonate containing pyrite, pyrrhotite,
galena, and arsenopyrite from the Tom deposit, Yukon Territory, Canada. Later tectonic veins also present. Goodfellow
(2004) interpreted the veins to represent the stringer zone deposit. From W.D. Goodfellow. N. Rounded clasts of fine-
grained pyrite and argillite in the Main band, Sullivan deposit. Lydon (2004a) interpreted the larger and elongated clasts at
the bottom of the sample, which is the base of the Main Band, to reflect proximity to their source in the immediate footwall.
Upward and farther from the base, rounding of pyrite and comminution of argillite clasts reflect a greater distance of trans-
port and a greater amount of autogenous milling during sulfide flow (Lydon, 2004a). From J.W. Lydon. O. Nodular carbon-
ate in black siliceous mudstone from the Anarraaq deposit. Note pyrite and sphalerite partially replacing the carbonate. P.
Highly mineralized sample from massive sulfide zone at Anarraaq, showing chaotically textured iron sulfide, dark red-brown
sphalerite and minor galena.

continental margins (Fig. 20). The only deposits that are Because many of the continental margin-hosted deposits
clearly formed in failed rifts are those of the northern Aus- are Paleozoic in age, they can be reasonably located on plate
tralian Proterozoic (Table A1), which are described by Large reconstructions (e.g., Goodfellow, 2004). However, Protero-
et al. (2005). SEDEX deposits occurring in continental mar- zoic tectonic configurations are poorly constrained. Thus, the
gin settings have a greater temporal and geographic range. paleogeographic placement of the northern Australian

0361-0128/98/000/000-00 $6.00 588


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 589

A) continental back-arc extension However, recent isotopic studies show that the abundant mi-
accretion & sedimentary basins crobial carbonates clearly record C isotope signatures consis-
arc magmatism hosting SEDEX deposits tent with precipitation from Proterozoic seawater (e.g.,
crust Braisier and Lindsay, 1998; Large et al., 2001). Thus, al-
mantle though they were intracontinental and were developed on
upwelling thinned continental crust, these basins were embayments in
asthenosphere the sense that they were connected to the open ocean for
440 km much of their history. As such, they were natural factories for
the generation of the basinwide shallow water to emergent
570 km carbonate-platform and evaporite strata that characterize the
500 km upper mineralized parts of these basins.
A summary of the understanding of the intracontinental
B) AXIAL
RIFT
RIFTED MARGIN PLATFORM
rift-hosted deposits in the Proterozoic basins of northern Aus-
platform sediments
Sea Level
alkaline volcanism tralia is presented by Large et al. (2005). In terms of the rela-
tholeiitic sills tionships between tectonic setting and mineralization, basin-
scale architecture and hydrologic models exist that arguably
crust explain both the tectonostratigraphic location of the deposits
and the overall fertility of the Australian basins (Large et al.,
extensive 2002). These are reliant on long-lived, intracontinental (but
partial melting
alkaline
uppervolcanism
mantle
marine-linked) basin systems that incorporate regionally ex-
tensive (>500,000 km2) clastic aquifer systems overlain by
equally extensive hydrologic aquitards, provided in this case
upwelling by platform carbonates. Large et al. (2002) suggested that
asthenosphere basins with these configurations are rare in Paleozoic conti-
nental settings, perhaps explaining the apparent restriction of
RIFTED MARGIN PLATFORM
C) alkaline volcanism
large intracontinental rift-hosted deposits to the Proterozoic.
Sea Level In contrast to the intracontinental rift-hosted deposits,
where it can be argued that the requisite basin settings are
rare, rifted passive margins are common in time and space
oceanic crust
throughout the Paleozoic. However, most of these systems do
continental not host stratiform Zn deposits. Some of the known deposits
crust and deposit clusters (e.g., Selwyn basin, Red Dog district)
upper have been studied in various degrees of detail, and general-
mantle
ized basin-scale models for their formation have been pro-
mantle posed (e.g., Large, 1983; Lydon, 1983; Goodfellow et al.,
1993; Dumoulin et al., 2004; Young, 2004; Fig. 20B,C).
Sedimentary and basin setting: Regardless of the tectonic
FIG. 20. Cartoons showing tectonic models for the development of sedi- setting, most basins that contain SEDEX deposits have the
mentary basins hosting SEDEX deposits. A. Intracontinental or failed rift same basic stratigraphic elements comprising a basal clastic
setting where extensional basins are developed in the overriding plate related
to a north-dipping subduction zone along the southern margin of the craton
and/or volcanic dominated succession, overlain by shales and
(e.g., northern Australia; Large et al., 2005). B. Continental rift floored by carbonates. Within this broad basin architecture, the deposits
oeanic crust and filled with a thick sequence of clastic sediments (e.g., Sel- occur in the upper succession where they are linked by two
wyn basin; Goodfellow, 2004). C. Passive continental margin rift with oceanic common factors. First, they are hosted by reduced, fine-
crust outboard from continental crust and sedimentary basin (e.g., northern grained siltstones-shales-mudstones and/or carbonate units
Alaska; Young, 2004).
within these reduced sediments. With one exception (Lady
Loretta: Large et al., 2005), where sedimentological analysis
SEDEX basins is limited only within a reconstruction of the is available, these host rocks reflect deposition in relatively
northern Australian terrane. The overall setting for these deep-water subwave base settings. In the intracontinental rift
basins has recently been related to extension in the overriding systems that are dominated by shallow to emergent clastics,
plate associated with a northerly dipping subduction system carbonates, and evaporites, these subwave-base facies may re-
in central Australia (Betts et al., 2003). In contrast to the open flect anything from relatively abrupt and short-lived periods
marine setting that prevailed for the continental margin- of localized structurally controlled increases in accommoda-
hosted deposits, the northern Australian basins are generally tion space (e.g., HYC: Large et al., 1998), to longer lived pe-
portrayed as extensional (e.g., Plumb et al., 1980; Betts et al., riods of more aerially extensive marine transgressions (e.g.,
2003) or transtensional (e.g., Muir, 1983) rifts developed en- Century: Andrews, 1996). In the case of the continental mar-
tirely within continental crust (Fig. 20A) in the manner of the gin systems, deep-water reduced sedimentation rates are gen-
landlocked East African rift system. Consequently, much of erally the rule rather that the exception (e.g., in the Belt and
the carbonate- and evaporite-dominated part of the basin fill Selwyn basins, the shale host for SEDEX deposits highlights
for these deposits has been interpreted in terms of lacustrine the importance of restricted sedimentation rates in preserv-
facies models (e.g., Muir, 1983; Donnelly and Jackson, 1988). ing organic matter). This is borne out in the Selwyn basin

0361-0128/98/000/000-00 $6.00 589


590 LEACH ET AL.

itself, where the subbasins that host the Howards Pass de- approximately 400°C and high salinity. However, the Rosh
posits were almost entirely starved of siliciclastic sediment Pinah deposit may be a transitional shale-hosted to VMS de-
(Goodfellow and Jonasson, 1987). It is also the case for the posit. Sphalerite-bearing veins near the Century deposit are
Kuna basin, where Red Dog and Anarraaq are hosted by 45 believed to represent a late paragenetic stage of the Century
to 250 m of sediments that accumulated over ~30 m.y. (Du- mineralizing event (Broadbent et al., 1998). Fluid inclusions
moulin et al., 2004; Young, 2004). In both cases, adjacent car- studies (Bresser, 1992) yielded homogenization temperatures
bonate platforms that rim the basin are interpreted to have of 98° to 180°C and 8.9 to 21.5 wt percent NaCl equiv. Polito
restricted siliciclastic input, preserving organic carbon and et al. (2005) report preliminary fluid inclusion temperatures
creating conditions for anoxic bottom water with locally high of 70° to 160°C, with salinities of 15 to 23 wt percent NaCl
productivity. equiv from coarse-grained sphalerite in the ore zone.
The second factor that links all SEDEX deposits is that they Fluid inclusion data have been reported for the Red Dog
occur in areas affected to some degree by synsedimentary deposit (Forrest, 1983; Edgerton, 1997a, b; Leach et al.,
faulting and/or subbasin formation. This can be the most 2004). Homogenization temperatures of fluid inclusions from
striking feature of the deposit area geology, reflected in quartz in vein and massive sulfide ore at Red Dog are be-
abrupt lateral facies and thickness changes and interfingering tween 175° and 329°C and salinities were 0 to 8 wt percent
of coarse-grained fault scarp-derived debris with the fine- NaCl equiv (Forrest, 1983). Homogenization temperatures of
grained mineralized host rocks such as at HYC (Large et al., between 255° and 302°C and salinities of 4 to 5 wt percent
1998) and many of the Selwyn basin deposits (e.g., Jason and NaCl equiv were reported for five fluid inclusions of uncer-
Tom: Goodfellow, 2004), or it can be subtle changes in the fa- tain trapping origin in vein sphalerite. Edgerton (1997a) pro-
cies and thickness of the host strata such as at Century vided a large data set that contains more than 300 measure-
(Broadbent et al., 1998) or Anarraaq (Dumoulin et al., 2004). ments on barite, quartz, calcite, witherite, and sphalerite that
There are two reasons for this association. In all cases, synsed- span an extremely wide range in temperatures
imentary faults are an obvious pathway for the focused ascent (100°–>350°C). Subsequent work on the Red Dog deposits
of metal-bearing brines from deeper basin aquifers. In addi- determined that a later hydrothermal fluid deposited much of
tion, for the deposits that are considered to have formed in the quartz. Primary fluid inclusions in sphalerite yielded ho-
brine pools, some form of depression or subbasin is necessary mogenization temperatures of between 100° to 200°C and ice
to trap the exhaled brines. melting determinations indicated salinities of about 14 to 19
wt percent NaCl equiv (Leach et al., 2004).
Nature of the ore fluids Source of the ore fluids: In spite of the fact that direct fluid
Temperature and salinity: Definitive fluid inclusion evi- inclusion data are limited, it is generally assumed that the ore
dence on the temperature, composition, and source of the fluids were principally hot metalliferous basinal brines (e.g.,
ore-forming fluids for SEDEX deposits is extremely limited. Badham, 1981; Lydon, 1983). A generalized model is shown in
Although the temperatures of SEDEX ore fluids are com- Figure B19 (from Lydon, 2004a). Elevated temperatures of
monly considered to be higher than MVT ore fluids, there are brines that formed SEDEX deposits could be due to unusually
few data that support this assumption. The most widely high geothermal gradients associated with extensional tecton-
quoted fluid inclusion data for SEDEX deposits are from the ism or from the circulation of the brines to sufficiently large
Jason (Gardner and Hutcheon, 1985) and Tom (Ansdell et al., crustal depths. High heat flow could be linked to contempora-
1989) deposits in the Selwyn basin. Ansdell et al. (1989) re- neous, usually deep, magmatic activity in some cases, particu-
ported fluid inclusion data for ankerite in the alteration zone larly for deposits that have spatially and temporally associated
and quartz in veins that cut the main ore zones in the Tom de- igneous rocks (i.e., Sullivan), although the volume of igneous
posit. They discounted the data from quartz because it prob- rocks for most deposits is usually minor. In most deposits,
ably formed during postore deformation and metamorphism. however, there is generally little to no evidence to support di-
The salinity data for ankerite could be valid but possible pos- rect heating, or contribution of metals, by a magmatic body to
tore modification of the fluid inclusion densities may have af- the hydrothermal systems (Leach et al., 2004; Lydon, 2004a).
fected the homogenization temperatures. The salinities for High salinities have been attributed to either subsurface
ankerite are 2 to 18.3 wt percent NaCl equiv (Ansdell et al., dissolution of evaporites (Land and Macpherson, 1992) or the
1989). Gardner and Hutcheon (1985) reported fluid inclusion downward migration of dense brines. However, except for the
data for the Jason deposit which also experienced postore de- northern Australian Proterozoic basins, sequences of evapor-
formation and low-grade metamorphism. Temperatures re- ites have not been reported in most sedimentary basins host-
ported for the Jason deposit (in ankerite) range from 234° to ing SEDEX deposits, and therefore the dissolution of evap-
274°C, with salinities of 8.1 to 15.2 wt percent NaCl equiv orites in the subsurface is unlikely to be the dominant source
(Gardner and Hutcheon, 1985). of the high salinities in most SEDEX deposits. The case for
A wide range of salinities (<1 wt % to > 45 wt % NaCl gravitational settling of evaporitic brines into the lower part of
equiv) and temperatures (<100°–>400°C) are reported for the rift-fill sequence is much more supportable (Leach et al.,
the Sullivan deposit (Leitch and Lydon, 2000). However, un- 2004; Lydon, 2004a). Direct evidence comes from fluid in-
certainties as to the geologic context of the veins and a history clusion electrolyte data from the Red Dog deposits that show
of concurrent fluid flow make interpretation of the fluid in- that the ore fluids derived their salinity from the evaporation
clusion data difficult (Lydon, 2004b). Frimmel (2001) reported of seawater (Leach et al., 2004). Other more indirect evi-
that a few primary fluid inclusions in the silicified footwall at dence in support of the evaporitic brine model are that many
the Rosh Pinah deposit in Namibia yielded temperatures of SEDEX deposits formed in the high evaporation latitudes of

0361-0128/98/000/000-00 $6.00 590


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 591

the contemporaneous equator (see discussion below). In ad- Deposit/District


dition, indications of evaporitic conditions have been found in

North Australia
HYC
the host sequences or their lateral stratigraphic equivalents Century
for several sedimentary basins that host SEDEX deposits, Lady Loretta
Mt. Isa
such as gypsum nodules and salt casts in the McArthur basin
George Fisher
(e.g., Goodfellow, 1983; Muir, 1983; Cooke et al., 2000), gyp-
Sullivan
sum nodules on the margins of the Belt-Purcell basin (Chan-

Canada
Howards
dler and Gregoire, 2000), and anhydrite-bearing red beds and Pass Tom
supratidal dolomite in shallow platform carbonate rocks adja- Jason
cent to the Kuna basin in the Brooks Range, Alaska (Du-

Argentina India U.S.A. Germany


Meggen
moulin et al., 2004).
Rammelsberg
The sinking of basin-hugging brines is a process that has
been described for brine-pool models of ore deposition Red Dog
(Sangster, 2002). Deposits that display evidence for shallow Balmat
subsea-floor replacement could form by the infiltration of Rampura Agucha
brines into sediments underlying a brine pool. A critical fac-
tor regarding the feasibility of such a process is the perme-
ability and porosity of the sediments. The brines will not sink Aguilar

if the underlying sediments have low permeability (e.g., mud-


-10 -5 0 5 10 15 20 25 30
stone). However, if the sediments are coarse clastics, the per-
δ 34 S (‰CDT)
meability may be sufficiently high to allow the overlying brine
to sink by displacing the less dense pore water. The sinking FIG. 21. Distribution of sulfur isotope compositions for sphalerite from
brine process was mentioned by Williams (1978) with refer- selected SEDEX deposits. If source included a numerical listing of data, the
median (vertical white line), 25th and 90th percentiles (extent of box), and
ence to the origin of the HYC deposit in Australia but re- total range of data (horizontal bar) were calculated and plotted. If source
cently challenged by Ireland et al. (2004). listed only range of values, data are shown by horizontal bar only. A complete
Metal content of the ore fluids: Considering that both MVT list of data sources is presented in Appendix D (Fig. D2).
and SEDEX deposits formed from basinal brines at similar
temperatures and brine compositions, the previous discussion
on the transport of metals for MVT ore fluids are applicable deposits with sufficient data, this trend is well established
to SEDEX deposits. However, Cooke et al. (2000) empha- (e.g., HYC, Century deposits: Large et. al., 2005; Red Dog
sized control of the metal content of the ore fluid by the redox deposits: Kelley et al., 2004b). This isotopic progression may
character of the rocks in the basin that hosts SEDEX de- indicate that the mechanism for deriving reduced sulfur pro-
posits. Cooke et al. (2000) proposed two types of SEDEX ore gressed from BSR (producing light values) to TSR (heavier
fluids dependent on the nature of the dominant sulfur species values) through time and with increasing temperature (Kelley
present in solution (H2S or SO): (1) reduced, acidic, and mod- et al., 2004b). Alternatively, it is possible that the reduction of
erate- to high-temperature (200°–300°C) brines similar to sulfate occurred in an isotopically closed system that became
those that form VMS deposits, which contain barite ± Au in progressively enriched in heavy sulfur over the life of the min-
the ore assemblage; and (2) oxidized, neutral to mildly alka- eralization (Broadbent et al., 1998; Large et al., 2005).
line pH, low-temperature (100°–250°C) brines that do not In some deposits, there is a lateral or vertical zoning of sul-
contain appreciable barite or Au. Cooke et al. (2000) sug- fur isotope values. For example, in the Tom deposit, the δ34S
gested that the H2S-rich brines are generated in reduced values of galena increase from about 6 to 8 per mil in distal
shale and siliciclastic turbidite-bearing sedimentary basins parts of the orebody to greater than 10 per mil in the vent
and commonly form vent-proximal SEDEX deposits, whereas complex, a pattern that has been interpreted to represent in-
the SO4-rich brines are generated in oxidized sedimentary creased replacement of barite by galena (Goodfellow, 2004).
basins containing red-bed rift fill and a carbonate sag phase Similar increasing δ34S values have been reported for the Sul-
and commonly form vent-distal Zn-Pb-Ag deposits. livan deposit (Lydon, 2004a) with bedded ore sulfides gener-
ally lighter (–10 to +5‰) than sulfide minerals in the vent
Sources of sulfur complex (–5 to +5‰). Stratigraphic variations have also been
The sulfur isotope values of galena and sphalerite in recorded. In the Howards Pass deposit, δ34S values of 10 to
SEDEX deposits show a spread of data from about –10 (Sul- 20 per mil display a general decrease upward in the XY de-
livan) to nearly 30 per mil (Aguilar), but the bulk of the val- posit. This is consistent with banded ore in the Sullivan de-
ues fall within the –5 to 15 per mil range (Fig. 21). In most posit which displays a general trend for decreasing sulfur iso-
cases, the data are consistent with the ultimate source of the tope composition stratigraphically upward but with
sulfur as marine sulfate (either as seawater, porewater, or in superimposed steps to heavier values at the base of successive
preexisting sulfate minerals such as barite), with reduction of sulfide bands higher in the sequence (Lydon, 2004a). How-
sulfate to sulfide usually involving biogenic sulfate reduction ever, this contrasts with the HYC deposit in Australia, which
(BSR), thermochemical sulfate reduction (TSR), or both, de- generally shows an increase in δ34S values higher up the sec-
pending on the temperature and the availability of reductant. tion (Large et al., 2005). Taylor and Beaudoin (2000) attrib-
In some deposits, there is an isotopic fractionation trend to uted the stratigraphic trends toward lighter values to an in-
heavier δ34S values in the later stages of the paragenesis. For creasing proportion of sulfide being derived by BSR and

0361-0128/98/000/000-00 $6.00 591


592 LEACH ET AL.

trends toward heavier values to an increase in the proportion of recent research at the Cannington deposit (Bodon, 1998;
of hydrothermally supplied sulfide. Trends toward heavier Chapman and Williams, 1998; Walters and Bailey, 1998).
values upward may reflect sulfate reduction under restricted Some authors have suggested that Broken Hill-type deposits
or closed conditions or an evolution toward a dominance of are essentially metamorphogenic, with mineralization con-
hydrothermal sulfur with later and higher ore lenses. trolled by fluid migration along thrusts and shear zones influ-
enced by localized rheological and chemical contrasts (Hobbs
Broken Hill-Type Deposits et al., 1998; Fig. B21). This interpretation is difficult to rec-
We regard Broken Hill-type deposits as a category of oncile with other lines of evidence such as Pb isotope signa-
SEDEX deposits (Table A1). This reflects the dominant tures (Large et al., 2005), sulfide textures (Bodon, 1998), or
strata-bound, metasedimentary rock-hosted nature of miner- REE systematics (Bodon, 1996). A more realistic hybrid
alization (Goodfellow, 2004). However, a number of authors model involves structurally influenced hydrothermal zone re-
have argued that Broken Hill-type deposits are sufficiently fining of a premetamorphic protolith (Walters and Bailey,
different compared to other recognized SEDEX deposits so 1998; Marshall and Spry, 2000). In this model, refining does
as to warrant a separate classification (Parr and Plimer, 1993; not involve significant external introduction of metals but
Walters, 1998). A comparison of SEDEX and Broken Hill- does result in significant modification of metal zonation (Wal-
type deposits shows that Broken Hill-type deposits contain ters and Bailey, 1998).
significantly higher contents of Ag and Pb relative to the Given the extent and complexity of metamorphic, struc-
SEDEX category (Fig. B1, Table 1). In terms of median val- tural, and metasomatic overprinting, effective exploration
ues, Broken Hill-type deposits are almost three times more strategies for Broken Hill-type mineralization are generally
enriched in Ag and one and a half times more enriched in Pb based on empirical associations (Willis, 1996; Walters et al.,
compared to other SEDEX deposits. 2002). This is reflected in the separate listing of exploration
Characteristic features of Broken Hill-type deposits (Wal- characteristics for Broken Hill-type deposits in Appendix E
ters et al., 2002) include: (1) amphibolite-granulite facies of Table E1. Some of the exploration strategies utilize charac-
Paleo- to Mesoproterozoic clastic metasedimentary host se- teristics related to metamorphism and metasomatism that are
quences, (2) spatial associations with thin amphibolites and specific to Broken Hill-type deposits. Examples include use
felsic gneisses interpreted as probable metavolcanic rocks, (3) of geochemical signatures in detrital resistate indicator min-
relationships with thin, laterally extensive marker units such erals such as gahnite and garnet formed during amphibolite-
as quartz-gahnites and iron formations (Fig. B20), (4) no sig- to granulite-facies metamorphism and high-temperature
nificant association with reduced graphitic or pyritic stratigra- metasomatism (Walters, 2001).
phy, (5) stacked, low-aspect ore lenses modified by complex However, it is important to stress that not all SEDEX de-
structural overprints, (6) association with large-scale strata- posits that have undergone metamorphism are included in
bound alteration halos characterized by high K/Na ratios the Broken Hill-type category. Examples of SEDEX deposits
(>10) and Fe-(Mn) garnet spotting, (7) lack of an associated metamorphosed to amphibolite facies that are not regarded
stringer zone or evidence for Mg-rich alteration assemblages, as Broken Hill type (Table A1) includes the Rampura-
(8) association with Fe-Mn-Ca-F–rich skarnlike gangue as- Agucha, Rajpura-Dariba, and Sindesar-Khurd deposits in the
semblages, (9) extreme zonation between Pb-Ag– versus Zn- Rajasthan district of northwest India (Deb and Pal, 2004; Roy
dominant ore lenses, (10) a distinctive element association in- et al., 2004; Shah, 2004).
cluding Sb, Cu, As, Bi, and Au, and (11) a general lack of Given that metamorphism is a characteristic feature but not
pyrite within ore lenses. a prerequisite for inclusion in the Broken Hill-type category,
The extent that these differences reflect metamorphic and this raises the question of why all known Broken Hill-type ex-
structural overprinting and reworking of a SEDEX protolith amples appear to occur in Paleo-Mesoproterozoic terranes
versus fundamentally different genetic models remains a con- and what are the apparent premetamorphic differences com-
troversial and poorly resolved issue (Marshall and Spry, 2000; pared to other SEDEX categories. Although premetamorphic
Walters et al., 2002). Large et al. (1996) argued that a signifi- interpretation is a contentious process, significant differences
cant difference between Broken Hill-type and SEDEX de- that appear to represent mainly premetamorphic associations
posits relates to the redox difference between ore fluid and in Broken Hill-type deposits include the following: (1) nature
the sedimentary environment. In this model, Broken Hill- of the host sequence, characterized by dominant immature
type deposits are interpreted as having formed from reduced clastic metasedimentary sequences ranging from pelites to
H2S-bearing metalliferous brines that were exhaled or infil- psammopelites and psammites, an absence of calcareous
trated into oxic sedimentary sequences. This is supported by units or carbonates, and very minor graphite or pyrite in the
the relatively sulphur poor nature of Broken Hill-type miner- immediate host sequences; (2) a spatial association with
alization, general absence of pyritic host sequences, associa- minor bimodal metavolcanic units; (3) well-developed lateral
tions with magnetite-bearing Fe formations, occurrence of marker horizons; (4) high concentrations of Cu, As, Sb, Bi,
well-developed Mn halos, and the restricted range of δ34S val- Au, F, and Cl; (5) relative sulfur deficient assemblages with
ues in sulfides. only localized pyrrhotite or pyrite; (6) a small range of δ34S
However, classification schemes based on interpretation of values in sulfides; (7) strong zonation between Fe-Mn-Ca-Pb-
premetamorphic relationships are fraught with difficulty. Ag– and Si-Zn–dominant ore lenses; and (8) extensive strata-
There is increasing awareness of the role played by postpeak bound alteration halos involving Na, Ca, Mg, and Sr deple-
metamorphic hydrothermal and structural processes associ- tion, and K, Rb, Pb, and Mn enrichment with no evidence for
ated with Broken Hill-type deposits. This is mainly the result footwall feeder zones or breccia pipes.

0361-0128/98/000/000-00 $6.00 592


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 593

A significant factor in the ongoing debate for a separate clas- more clearly evident and dominant in the main Bergslagen
sification of Broken Hill-type deposits reflects an Australian district (Walters, 1998). While it is acknowledged that the
perspective based on the Paleo-Mesoproterozoic basins of Broken Hill-type deposits remain an enigmatic grouping,
northern Australia and the Broken Hill block. The Broken Hill there is sufficient evidence to support its inclusion within a
and Cannington deposits are broadly coeval in terms of host separate category of SEDEX deposits.
stratigraphic ages, with SEDEX deposits in lower metamor-
phic grade sequences in the Mount Isa Western Succession Lead Isotopes
and McArthur basin (Large et al., 2005). The quartz- In compiling data for a review and comparison of Pb iso-
feldspar–dominant clastic rocks with minor bimodal igneous tope compositions in MVT and SEDEX deposits, only those
units that characterize the host sequences at Broken Hill and for which an arbitrary minimum of five published analyses
Cannington contrast markedly with the reduced dolomitic silt- were available of ore-stage galena or sphalerite were selected.
stones and shales that typically host Australian SEDEX de- Data for gangue dolomite, barite, calcite, etc., were not in-
posits. This is further emphasized by differences in Pb isotope cluded. By these criteria, approximately 340 analyses for 18
ratios, with lower 207Pb/204Pb ratios in Broken Hill-type de- individual MVT deposits and 560 analyses for 28 SEDEX de-
posits that lie on a different growth curve to Australian posits were compiled. An additional 230 analyses for 12 MVT
SEDEX deposits, suggesting derivation of Pb from mantle-re- districts, for which insufficient data were available for indi-
lated, rather than upper crustal source rocks (Large et al., vidual deposits, have been included. Data are presented on
2005). However, differences are less marked in comparison to standard 207Pb/204Pb versus 206Pb/204Pb diagrams (Fig. 22)
SEDEX deposits elsewhere in the world that are hosted by along with Pb isotope growth curves of Zartman (1984).
more clastic dominated sequences (e.g., Sullivan) or in com- Expanded scales in some figures resulted in steep north-
parison to Broken Hill-type deposits outside of Australia that east-southwest linear arrays for several deposits (see App. D).
show more transitional SEDEX features (e.g., Gamsberg). Such arrays are analytical in nature and therefore have no ge-
The Sullivan district and deposit (Leitch et al., 2000; Turner ologic meaning. The most common cause of these arrays is
et al., 2000) are particularly relevant to improving our under- mass discrimination errors caused by slight variations in fila-
standing of premetamorphic associations in the Broken Hill- ment temperatures and sample loads and produces arrays
type category. The quartz wacke and quartz arenite turbidite with a constant slope of ~1.4. For modern analyses performed
sequences that characterize the host Lower and Middle under conditions of near-constant run temperature and load-
Aldridge Formations at Sullivan are similar to the interpreted ing techniques, errors due to mass fractionation effects
host sequences for Broken Hill-type deposits, although the should be ~0.15 percent for 207Pb/204Pb and ~0.1 percent for
206
anoxic nature of the Aldridge Formation (Goodfellow, 2000) is Pb/204Pb. Error bars reflecting these average values are in-
less comparable. The spatial association of mineralization at cluded in the Pb isotope diagrams.
Sullivan with the Moyie sills is also reminiscent of the associa- This overview of Pb isotope compositions of sediment-
tion with amphibolite units in Broken Hill-type settings. While hosted Pb-Zn deposits does not attempt to present either a
the Moyie sills postdate the main phase of mineralization they detailed analysis or interpretation of the available data.
are associated with extensive albite alteration and may have in- Rather, the data are graphically presented in what the authors
fluenced fluid migration during subsequent phases of over- regarded as logical groupings and general observations are of-
printing hydrothermal events. There are indications that these fered regarding the nature and distribution of the data.
later hydrothermal events were long lived and resulted in sig- Figure 22 illustrates the extreme range of data represented
nificant textural and mineralogical refining of the Sullivan de- by these deposits and districts. 206Pb/204Pb values, for exam-
posit, with massive pyrrhotite replacing sphalerite and galena ple, range from about 13.5 to more than 24, a spread which is
in the vent complex and bedded ores. Sericitic alteration entirely accounted for by MVT deposits and districts. Data
forms an extensive strata-bound halo around the Sullivan de- points beyond the 0 Ma isochron represent samples from
posit (Leitch at al., 2000), involving enrichment of K and de- MVT deposits in the midcontinent region of the United
pletion of Ca, Na, and Mg. This is similar to the alteration States and Canada and are discussed in more detail in the Ap-
trends around Broken Hill-type deposits (Stanley et al., 2001), pendix D. Although both data sets illustrate that some de-
which in turn are comparable to sericitic alteration associated posits contain Pb of mainly mantle derivation, the large ma-
with VMS deposits, and contrasts with alteration trends asso- jority of deposits lie above the orogene curve indicating, not
ciated with other SEDEX deposits (Large et al., 2000). surprisingly, the predominance of continental crustal sources
Some Broken Hill-type deposits show transitional trends for these sediment-hosted deposits.
that suggest affinities with other SEDEX categories or VMS SEDEX deposits, as a global group (Fig. 22A), are bimodal in
styles. The Gamsberg deposit in South Africa has undergone terms of model Pb ages and display a hiatus between ~800 and
less extensive metamorphic and hydrothermal overprinting ~1100 Ma. Mississippi Valley-type deposits (Fig. 22B), although
compared to other Broken Hill-type deposits and shows tran- almost entirely comprising a younger model Pb age grouping,
sitional associations with other SEDEX categories (Rozen- present a hiatus between ~800 and ~1400 Ma. Mississippi Val-
daal, 1980; Rozendaal and Stumpfl, 1984). This includes fine- ley-type deposits also exhibit an abundance of so-called future
scale banding, zones of banded pyrite, graphitic host model Pb ages, a feature lacking in the SEDEX data.
sequences, and distal stratiform barite. The Zinkgruvan de-
posit in the southern Bergslagen district of Sweden is gener- MVT deposits and districts
ally considered to be a Broken Hill-type deposit but shows A basement source for Pb is indicated for many MVT
transitional relationships to VMS deposit styles that become districts or deposits, including several of the world’s major

0361-0128/98/000/000-00 $6.00 593


594 LEACH ET AL.

16.2 SEDEX deposits


A) Pb isotope data for a majority of SEDEX deposits plot
above the orogene curve, indicating a predominance of conti-
15.8
800
UC nental crust sources for these deposits. Furthermore, a few
1600 ORG even plot above the upper crust curve (i.e., high micrometer),
0 Ma
UM indicating a large component of a very radiogenic source.
Pb

15.4 2000 LC SEDEX dating: Traditionally, SEDEX deposits have been


204

400
800 assumed to be essentially contemporaneous with host-rock
Pb /

2400
1200 sedimentation. The Red Dog deposit, which is the only one
15.0
directly dated to test this assumption, yielded an Re-Os age
207

1600
2800 which coincides with that of the host rock age (Morelli et al.,
2000
2004).
14.6
More commonly, global growth curves are used to calculate
SEDEX only model Pb ages for SEDEX deposits. To test this, model Pb
ages of a selection of eight SEDEX deposits were compared
14.2 with their host-rock ages. All gave model ages that ranged
13 15 17 19 21 23
206 204
from slightly older to much older than their respective host
Pb / Pb rocks (see App. D), indicating that global U-Pb growth curves
16.2 should not be used to calculate model Pb ages. A better ap-
B) proach involves application of terrane-specific Pb growth
curves, as was done for the northern Australia deposits (Large
15.8 et al., 2005).
800 UC
1200 MVT versus SEDEX comparison
1600 ORG
0 Ma
UM The degree of within-deposit Pb isotope homogeneity is
Pb

15.4 2000 LC
frequently cited in the literature as a distinguishing feature
204

400
800 between SEDEX and MVT deposits. Comparison of the data
Pb /

2400
1200
15.0
compiled for the present review demonstrates the similarity
207

1600
in homogeneity between the two deposit subtypes, thus pre-
2800
2000 cluding the general use of within-deposit variation to distin-
14.6
guish between MVT and SEDEX deposits (Fig. D3A-O).
2400 Other observations include: (1) Pb in both subtypes ap-
MVT only pears to have been derived from crustal sources, in accord
with their close association with sedimentary basins; (2) MVT
14.2
13 15 17 19 21 23 deposits in the United States midcontinent region contain
206 204 very high abundances of radiogenic Pb, whereas no SEDEX
Pb / Pb deposits analyzed to date reveal such abundant radiogenic Pb;
FIG. 22. Pb isotope plots for (A) MVT and (B) SEDEX deposits. All but a (3) model Pb ages in SEDEX deposits are generally older
very few analyses represent ore-stage galena; the others are sphalerite. Non- than the ages of their host rocks. MVT model ages, in con-
sulfide phases are not included. Growth curves from Zartman (1984). Values trast, show little or no correlation with mineralization-age de-
of mm (238U/204Pb) for each growth curve are as follows: mantle = 8.37; lower
crust = 6.21; orogene = 11.31; upper crust = 13.22. Small dots on the growth terminations; some MVT deposits yield future model ages
curves denote positions of isochrons at 400-m.y. intervals. Lack of space pre- while others show model ages significantly older than the host
cluded consideration of 208Pb/204Pb vs. 206Pb/204Pb data plots. rocks.
A complete list of data sources is presented in Appendix D (Fig. D3).

Fluid Flow, Metal Transport, and


Modeling of Pb-Zn Ore Genesis
Hydrogeologic studies of sediment-hosted ores have ad-
districts. Lead isotope data for Bushy Park, Pering, Gays dressed fundamental research issues concerning the mecha-
River, Irish Midlands, Lennard Shelf, Upper Mississippi Val- nisms of flow, controls on rates and flow patterns, effects of
ley, Southeast Missouri, and Tri-State (see App. D) have all faults, and geochemistry of transport and mechanisms of ore
been interpreted in terms of a significant component from deposition. Addressing these issues in a quantitative fashion,
basement-derived Pb. High micrometer values for several numerical models have been developed which are based on
districts also confirm a supracrustal origin for Pb. In contrast, the equations governing fluid flow, heat transfer, and metal
Newfoundland Zinc is alone among the MVT deposits to in- transport in sedimentary rocks. Although both theory and
dicate an anomalously low-micrometer source for Pb. computer technology exist to model three-dimensional fields,
MVT dating: A comparison of MVT model Pb ages shows a most studies have focused on two-dimensional calculations
very poor correlation with measured mineralization ages (see because they are easier to build and analyze, and because in
App. D), suggesting strongly that the former should not be re- many cases geologic and geochemical data are not well
lied on to accurately date MVT mineralization. enough constrained or three-dimensional hydrologic results

0361-0128/98/000/000-00 $6.00 594


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 595

too cumbersome to render at the basin scale (Garven, 1995; Only one Ph.D. dissertation deals with the hydrology for
Raffensperger, 1996). Pb-Zn ore formation in the Irish Midlands/Carboniferous
MVT ores in foreland basins: These deposits have received basin (Hazlett, 1997). This thesis presents finite element
the greatest amount of hydrogeologic study, motivated mostly models for both thermal convection and topography-driven
by curiosity and scientific debates in the 1970s and 1980s flow in the Carboniferous platform. Hazlett (1997) was not
about the hydrologic mechanisms driving large-scale brine able to resolve which fluid-flow mechanism best represented
migration (Garven and Freeze, 1984; Person and Garven, the ore system for the Midlands, as neither hydrologic model
1994) and the geochemistry of metal transport and ore depo- could replicate fluid inclusion data under geologically realis-
sition (Sverjensky, 1986; Plumlee et al., 1994). The methods tic boundary conditions. Overall, however, the fault-con-
used for modeling reactive transport with heat and fluid flow trolled free convection simulations were more encouraging.
are reviewed by Garven (1985, 1995). With these results in hand, Garven et al. (1999) presented re-
Garvin and Freeze (1984) suggested that topography-dri- active flow simulations at the deposit scale, to test a concep-
ven flow appears to be the most robust hydrologic mechanism tual model whereby metal-bearing fluids migrate up normal
for MVT ores, although Cathles and Adams (2005) disagree. faults to mix with sulfide-bearing fluids within platform car-
The primary argument for topography-driven flow is that bonate strata at Lisheen, Ireland. The reactive flow model
MVT deposits are commonly associated with the uplifted shows how the fluids mix on time scales of 104 to 105 yr and
margins of foreland and intracratonic basins. Furthermore, how ore mineralization is concentrated in the deepest porous
the flow patterns, rates of flow, and thermal effects generated carbonate beds adjacent to the faults, provided adequate
by topography-driven hydrologic systems seem to provide op- sources of metals and H2S are sustained.
timal conditions for ore genesis, conditions that are not simu- SEDEX ores in extensional basins: For the McArthur basin,
lated in models of other fluid-flow processes, such as basin Garven et al. (2001) built and compared hydrologic models
subsidence and compaction (Cathles and Smith, 1983; for topography-driven flow, thermally-driven flow (free con-
Bethke, 1985; Swenson et al., 2004), rupturing of overpres- vection), and transient flows associated with overpressuring
sured sediments and basement by normal faults (Sharp and and fault rupture. These models were designed to test a vari-
Domenico, 1976; Roberts et al., 1996), convection near high ety of hypotheses concerning deep fluid migration and the
heat-producing granites (Fehn et al., 1978; Solomon and origin of base metal ores at HYC (Large et al., 1998). The
Heinrich, 1992), or the postseismic flow associated with numerical results for thermal convection display a strong
thrust faulting (Oliver, 1986; Garven et al., 1993). Some geo- structural control on fluid flow caused by multiple north-
physicists argued that the models of Garven and Freeze trending fault systems, which characterize the Batten trough.
(1984) require permeabilities that are too large and/or flow The basin-wide flow pattern suggests that Na-Ca-Cl brines
rates too high, which would lead to premature flushing of acquired base metals in the deepest levels of the basin stratig-
brines from the foreland basin (Deming and Nunn, 1991; raphy as fluids migrated eastward through the aquifer system.
Cathles and Adams, 2005). Whatever the numerical limita- Upward flow was relatively rapid along the Emu Fault zone,
tions of modeling topography-driven flow with respect to where fluid-discharge temperatures are predicted to have
MVT ore genesis, it is difficult to ignore the close association been 130°C in the muddy sediments near the sea floor at
with tectonic uplift, orogenesis, and other geochemical HYC. Transient pulses of flow probably characterized later
processes that appear to be affiliated with continental-scale periods of transpressional stress, thrusting, and strike-slip
brine migrations documented in North America (Leach and faulting that punctuated the basin history. Further hydro-
Rowan, 1986; Oliver, 1986; Bethke and Marshak, 1990; geologic studies of thermohaline convection by Simms and
Arnold et al., 1996; Appold and Garven, 1999; Rowan and Garven (2004) and Yang et al. (2004) explored these topics
Marsily, 2001; Bradley and Leach, 2003). further.
Multicomponent chemical species transport and reactive A similar hydrologic scenario may apply for stratiform ore
flow modeling have been applied to carbonate-hosted ores in mineralization in the extensional late Paleozoic Kuna basin of
the MVT districts and Irish Midlands (Garven and Raf- northwestern Alaska. Garven et al. (2003) modeled hy-
fensperger, 1997; Appold and Garven, 1999, 2000; Garven et drothermal flow patterns in the Kuna basin, but reactive flow
al., 1999; Corbella et al., 2004). These numerical studies modeling needs to be integrated with the heat and fluid-flow
model various geochemical scenarios for ore deposition and modeling to evaluate geochemical mechanisms of metal
show that replacement of preexisting iron sulfides and/or transport, fluid mixing, and ore deposition. Unfortunately, the
local mixing of fluids rich in H2S and metals are the most ef- stratigraphic sequence and ore deposits at Red Dog have
fective way to deposit sphalerite and galena ores, whereas been severely deformed, thrust faulted, and silicified during
processes of cooling, pH shift, or dilution are less effective hy- the Middle Jurassic to Early Cretaceous Brooks orogeny, and
drologically. For example, Appold and Garven (2000) pre- therefore it is difficult to identify the late Paleozoic normal
sented two-dimensional numerical models of mixing that re- faults that probably served as conduits for hydrothermal con-
produce generally the metal zonation pattern observed in the vection in this extensional basin.
Viburnum Trend.
Irish Midlands basin ores: Consistent with the debate on The Secular Distribution of MVT and SEDEX Ores
the origin and classification of the Irish Midlands ores, there The uneven distributions of MVT and SEDEX deposits
is uncertainty regarding the fluid-flow mechanisms for the (Fig. 23) in the rock record (Meyer, 1981, 1988; Sangster,
Irish ores, dating back to classic models of thermal convection 1990; Barley and Groves, 1992; Goodfellow et al., 1993;
for seawater (Russell, 1978). Titley, 1993; Leach et al., 2001; Lydon, 2004b) are generally

0361-0128/98/000/000-00 $6.00 595


596 LEACH ET AL.

A) MVT deposits (Fig. 23A). This pattern is more pronounced when the ages of
Phanerozoic Proterozoic Archean
MVT deposits are plotted against tons of contained metal
NP MP PP (Fig. 24). The paucity of MVT ores in the Proterozoic con-
50 trasts with the abundance of SEDEX ores (Fig. 23B). The
45 NPa K J Tr P C D S O C
contrast in abundance between MVT and SEDEX ores in the
50 Precambrian is unlikely related to preservation because
45
40
40
SEDEX deposits that formed mainly along passive margins
35 would be more prone to erosion and destruction relative to
Pb + Zn (Mt)

35
MVT deposits that formed mainly inboard of orogenic belts.
Pb + Zn (Mt)

30
30 25 The different distributions can also not be attributed to the
20
25
abundance of carbonate rocks required for MVT deposits, be-
15
10 cause carbonate rocks are abundant in the Precambrian
20 5 (Condie, 2000). Considering the possible implications of this
15
0 observation, a more detailed discussion of the Precambrian
0 100 200 300 400 500 600
MVT deposits is presented.
Age (Ma)
10
Precambrian MVT deposits
5
The significant Precambrian MVT deposits are listed in
0 Table 2. Stromatolitic dolostone of the Campbellrand Sub-
0 500 1000 1500 2000 2500 3000
group (2.52–2.65 Ga), Transvaal Supergroup, is the only
Age (Ma)
known Archean carbonate succession that hosts Pb-Zn min-
eralization similar to MVT deposits. Among numerous
B) SEDEX deposits known occurrences, at least two, the Bushy Park prospect
Phanerozoic Proterozoic Archean (Schaefer, 2002) and the Pering deposit (Wheatley et al.,
NP MP PP 1986) are of economic size. A small number of dolostone suc-
80
cessions of late Paleoproterozoic and Mesoproterozoic age
70 host subeconomic strata-bound Pb-Zn mineralization. Min-
eralization ages have been determined for only two deposits
60 (Table 2). A Paleoproterozoic mineralization age (2.0–2.1
Ga) has been firmly established for the deposits hosted by
Pb + Zn (Mt)

50 the Neoarchean Transvaal Supergroup (Schaefer, 2002;


Duane et al., 2004). Attempts to date the Nanisivik deposit
40 have yielded ambiguous results. Based on paleomagnetic
studies, Symons et al. (2001) suggested an age of 1095 Ma,
30 whereas an Ar-Ar age of 461 Ma was obtained on alteration
orthoclase (Sherlock et al., 2004). There are no radiometric
20

10 Phanerozoic Proterozoic
NPa K J Tr P C D S O C NP MP
45
0
0 500 1000 1500 2000 2500 3000
40
Age (Ma)

FIG. 23. Age distribution of mean host-rock age vs. Pb + Zn (Mt) contents 35
for (A) MVT and (B) SEDEX deposits. Bars with light gray shading include
deposits with extreme uncertainty (~ ±300 m.y.) for the ages of the ores. Data 30
Pb + Zn (Mt)

from Table A1.


25

20
attributed to secular variations in geologic processes that con-
trolled the formation and destruction of these deposits. Al- 15
though these secular distributions are poorly understood,
10
they have the potential to focus exploration to favorable re-
gions or time slices in the rock record. The deposit endow- 5
ment data in Table A1 provide an opportunity to quantita-
tively examine the distribution of Pb and Zn ores during 0
0 200 400 600 800 1000 1200
Earth history. Age (Ma)
MVT deposits Mascot-Jefferson City Pine Point
The relatively few MVT deposits hosted in Precambrian Nanisivik Upper Silesia
rocks contrasts with their abundance in Phanerozoic rocks FIG. 24. Age of MVT ores vs. Pb + Zn (Mt). Data from Table A1.

0361-0128/98/000/000-00 $6.00 596


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 597

TABLE 2. Secular Data Showing Precambrian Deposits and Host-Rock Ages

Name Location Formation/Supergroup Host-rock age (Ga) Reference

Late Archean carbonate-hosted Pb-Zn mineralization


Bushy Park South Africa Campbellrand /Transvaal SG 2.52–2.651 Duane (2002)
Pering deposit South Africa Campbellrand/Transvaal SG 2.52–2.65 Wheatley et al. (1986); Greyling (2000)

Paleoproterozoic dolostone-hosted Pb-Zn mineralization


Esker prospect Canada Rocknest Fm /Coronation SG 1.89 Gummer et al. (1996)
Blende deposit Canada Gillespie Lake/Wernecke Mts SG 1.71–7.84 Robinson and Godwin (1995)
Coxco Australia McArthur Group 1.6–1.8 Walker et al. (1983)
Various occurrences Australia McArthur Group 1.6–1.8 Williams (1972); Solomon and Groves (2000)
Kamarga deposit Australia Lawn Hill Platform/Mt Isa inlier 1.6-1.8 Jones et al. (1999)

Neoproterozoic dolostone-hosted Pb-Zn mineralization


Nanisivik deposit Canada Society Cliffs Fm, Bylot SG 1.2 Sherlock et al. (2004)
Morro Agudo deposit Brazil Bambui SG 0.75 Oliveira et al. (undated)
Tianbaoshan Distict China Dengying Fm ?? Xiaochun et al. (2000)
Gayna River District Canada Little Dal Group ?? Gibbins (1983)
Goz Creek Prospect Canada Backbone Ranges Fm ?? Gibbins (1983)

Notes: Fm = Formation; SG = Supergroup; ?? = host-rock age unknown


1 Mineralization age of for the Archean deposits of the Transvaal SG have been established as 2.0 to 2.1 Ga (Schaefer, 2002; Duane et al., 2004)

or paleomagnetic age constraints available for any of the questionable. However, Pb-Zn ores hosted by Neoprotero-
other examples. Thus, it is uncertain for most examples zoic dolostones are clearly MVT.
whether or not the epigenetic mineralization is indeed of Pre- The absence of MVT deposits in the Archean may be attrib-
cambrian age. uted to elevated heat flow through the early Earth’s crust (Nis-
However, it is apparent that all the examples listed above bet and Fowler, 1983), the increased rates of tectonic recycling
have many characteristics common to Phanerozoic MVT ores. (de Wit, 1998), and the dominance of small crustal blocks.
Strata-bound breccia bodies host epigenetic sphalerite and These conditions were unsuitable to establish large carbonate
galena mineralization, with variable amounts of pyrite and/or platforms. With the stabilization of the first large cratonic
marcasite, and minor chalcopyrite. Important gangue miner- blocks in the Neoarchean and Paleoproterozoic, extensive car-
als are sparry dolomite, calcite, quartz, and minor solid hy- bonate platform environments developed for the first time.
drocarbon residues. Stable isotope and fluid inclusion data for These ancient carbonate platforms, though as vast and as com-
some deposits (Ghazban et al., 1990; Arne et al., 1991; Robin- positionally diverse as their Phanerozoic counterparts
son and Godwin, 1995; Wachowiak et al., 1997; Greyling, (Grotzinger and James, 2000), consisted primarily of stromato-
2000; Solomon et al., 2000; Schaefer, 2002) show that the ore lites and sea-floor precipitates. Synsedimentary lithification,
fluids were high-salinity brines at temperatures up to 300°C. early diagenetic dolomitization and silicification, lack of biotur-
Marine sulfate, reduced either by biogenic and/or thermo- bation, and coarse skeletal carbonate debris render these plat-
chemical reduction mechanisms, was evidently the most im- form carbonate successions impermeable (Grotzinger, 1989).
portant source of sulfide (Robinson and Goodwin, 1995; Sufficent permeability for MVT ore fluids may have been lim-
Greyling, 2000; Solomon et al., 2000). Limited Pb and Sr iso- ited to fractures or karst systems that tapped suitable aquifers.
tope data identify cratonic basement rocks as metal sources This may explain the strong structural control on most Pre-
(Schaefer, 2002; Duane et al., 2004). cambrian carbonate-hosted Pb-Zn deposits.
Although the characteristics described above are similar Neoproterozoic platform carbonate successions reveal the
to Phanerozoic MVT deposits, the genetic status of these shift from in situ precipitation of carbonate to trapping and
deposits is uncertain due to the spatial association of some binding of loose sediment in Neoproterozoic stromatolites
with igneous activity and the paucity of dates for mineral- plus reefs with increased complexity and porosity for the first
ization. The mineralization age for the deposits hosted by time in Earth’s history (Grotzinger and James, 2000). There is
the Transvaal Supergroup, for example, coincides with the also no doubt that modern plate tectonic processes operated
intrusion of the giant Bushveld Igneous Complex, and a at this time. It is during the Neoproterozoic that all the fun-
magmatic origin of these deposits may be invoked. The damental requirements to form MVT deposits approached
Blende deposit may also have an igneous-related origin the conditions encountered in the Phanerozoic.
since it is closely associated with a dioritic sill (Robinson and
Goodwin, 1995) and spatially associated with the enigmatic SEDEX deposits
Wernecke breccia bodies (Thorkelson et al., 2001). Little is SEDEX ores are mainly restricted to two time periods (Fig.
known about the Esker prospect, but its size and extensive 23); one in the Proterozoic and another in the Phanerozoic
thin, tabular nature (Gummer et al., 1997) is exceptional for (e.g., Goodfellow et al., 1993; Goodfellow, 2004; Lydon,
a MVT deposit. Thus, the classification of some of the Pb- 2004a). The Proterozoic group (~1.4–1.8 Ga) includes the
Zn ores hosted by ancient carbonate successions remains large deposits in the Mt. Isa-McArthur basin of Australia

0361-0128/98/000/000-00 $6.00 597


598 LEACH ET AL.

(Large et al. 2005), the Sullivan deposit in Canada, and the the extreme deposit attributes of Broken Hill-type deposits,
deposits in the highly metamorphosed rocks of the Aravalli- they are excluded from this comparison. We focus on the
Delhi belt of northwestern India. The Phanerozoic ores in- broad features of SEDEX and MVT deposits that may be
cludes the large Red Dog deposits and the deposits in the Sel- used to compare the similarities and contrast the differences
wyn basin. The older age group corresponds to SEDEX between these two subtypes of ores.
deposits formed in failed continental rifts, whereas deposits
of the younger age group formed along passive continental Metal character
margins (e.g., Large et al., 2005, and references therein). Nine of the 10 largest sediment-hosted Pb-Zn deposits are
SEDEX deposits are not known to have formed prior to about SEDEX. Of the deposits that contain at least 2.5 Mt total
2.2 Ga. The specific tectonic factors controlling the emer- metals, there are 35 SEDEX and 15 MVT deposits. However,
gence of SEDEX deposits are elusive, because the two set- these comparisons do not consider some giant MVT districts
tings in which they are known—rifts and passive margins— (e.g., Viburnum Trend) that resource data are available only
appear to have existed back as old as 3 Ga (e.g., Kusky and for district totals, despite the fact that the mines are simply
Hudleston, 1999). However, the emergence of SEDEX de- production units of a continuous mineralized zone. Although
posits (~1.8–2.2 Ga) correspond to the time when 80 percent the size distribution of Pb-Zn deposits is skewed by large
of the preserved continental crust had formed, global radi- SEDEX deposits, the two deposit types have an excellent cor-
ogenic heat production was about 1.5 times as great as today, relation between total tonnage and contained metal (Pb +
and plates, accordingly, may have been smaller (Leach et al., Zn), with a fairly consistent ratio of about 10:1, or average
2005, and references therein). grade of 10 percent, regardless of the size of the deposit or
SEDEX Pb-Zn deposits form mainly from oxidized fluids district. Zinc grades are approximately the same for both,
which contain low concentrations of reduced sulfur. Any sul- whereas Pb and Ag grades are about 25 percent greater for
fur that is present in these fluids must be in an oxidized state SEDEX deposits. Three times as many SEDEX deposits have
(e.g., sulfate). Therefore, the appearance of SEDEX Pb-Zn reported Cu contents and the median Cu value of SEDEX
deposits is likely a consequence of the evolution of the oxy- deposits is nearly double that of MVT deposits.
genation of the hydrosphere and atmosphere (Huston and Higher Cu contents of SEDEX deposits could reflect
Logan, 2004, and references therein). Between 2.4 and 1.8 higher fluid temperatures and sulfur redox conditions of the
Ga, the atmosphere became oxygenated, with the hydros- ore fluids. However, the greater Cu content of SEDEX ores
phere also progressively becoming oxidized, sulfide poor, and could be related to the silicliclastic nature of the host rocks
sulfate rich (e.g., Huston and Logan, 2004). Huston and that may also account for their greater Pb grades. These com-
Logan (2004) suggested that changes in oceanic geochemistry parisons suggest that the ore fluids and ore-forming processes
probably led to the flowering of SEDEX deposits, beginning are similar for MVT and SEDEX ores.
at ~1.8 Ga in India and between 1655 and 1575 Ma in Aus-
tralia. The evolution of sulfate-reducing bacteria by this time Nature of the ore fluids: Temperature, composition,
may have been critical to the production of H2S at the site of and source of the ore fluids
deposition, allowing deposition of ore metals (Goodfellow, The wealth of fluid inclusion data for MVT deposits shows
2004; Huston and Logan, 2004). Goodfellow and Jonasson they formed from basin brines that were mainly derived from
(1987) and Turner (1992) suggested that periods of SEDEX the evaporation of seawater. Temperatures of ore deposition
ore deposition were related to global periods of oceanic were usually between 90° and 150°C; a few deposits have
anoxia that provided ideal conditions for the accumulation of fluid inclusion homogenization temperatures in excess of
biogenically reduced sulfur in restricted basins, as well as 200°C. Metals were transported as chloride complexes in
preservation in these redox environments. brines containing low amounts of reduced sulfur at near-neu-
Goodfellow et al. (1993) noted that Proterozoic and Paleo- tral pH to slightly acidic conditions and buffered by carbon-
zoic SEDEX deposits formed at high-evaporation latitudes, ate rocks. However, some MVT deposits may have formed
consistent with the suggestion that some SEDEX ore fluids from metal-bearing fluids circulating in basements rocks that
derived their salinity from the evaporation of seawater (Leach mixed with fluids rich in reduced sulfur in the sedimentary
et al., 2004). Furthermore, indicators of evaporitic conditions cover rocks.
are present in the sedimentary sequences for several sedi- Despite the paucity of direct information on the nature of
mentary basins that host SEDEX deposits (e.g., McArthur SEDEX ore fluids, isotopic studies and deposit modeling
and Belt-Purcell basins). These observations suggest that suggest that they also formed from basin brines with similar
both passive margin and intracratonic basins that evolve in temperatures and salinities. Supporting evidence for these
evaporative zones are fertile sites for the genesis of SEDEX conclusions are the few fluid inclusion studies on sphalerite
ores. Support for this is the observation that more than 85 that show the ore fluids were seawater evaporative brines
percent of the Phanerozoic ores were deposited at paleolati- at temperatures less than 200°C. Other studies (e.g., Selwyn
tudes that mirror present-day high-evaporative global regions basin) indicate that temperatures could have been as high as
(Leach et al., 2005). ~275°C. Modeling studies, together with observed mineral
assemblages, are consistent with transport of metals in flu-
Comparison of MVT and SEDEX Deposits and ids at near-neutral pH and low contents of reduced sulfur.
Their Ore-Forming Processes However, hot (>200°C) acidic fluids in equilibrium with re-
Having examined each of the MVT and SEDEX subtypes duced siliciclastic-dominated basin sediments may have
internally, we now compare them against one another. Given formed some SEDEX deposits. There is no evidence that

0361-0128/98/000/000-00 $6.00 598


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 599

there were discernible differences between MVT and generally constrained by relationships to sedimentary or dia-
SEDEX ore fluids. genetic features in the rocks. These studies suggest that de-
position of SEDEX ores was coeval with sedimentation or
Lead and reduced sulfur sources early diagenesis, whereas some deposits such as Century
Lead isotope compositions for MVT and SEDEX deposits formed by replacement processes about 20 m.y. after sedi-
indicate that Pb was derived from a variety of crustal sources. mentation. Thus, MVT and SEDEX ores generally formed at
Neither MVT nor SEDEX deposits have Pb isotope compo- distinctly different times with respect to the age of the rocks,
sitions that are more homogeneous than the other. However, the latter closer to the host-rock age; however, there is some
only MVT deposits in the United States midcontinent overlap in the range of timing.
Lennard Shelf and Robb Lake region uniquely contain very
high abundances of radiogenic Pb. Thus, Pb isotope data for Tectonic setting
MVT and SEDEX do not indicate fundamental differences in Mississippi Valley-type deposits dominantly formed in plat-
the source of Pb. form carbonate sequences, usually localized within exten-
There is an apparent difference between the sulfur isotope sional zones inboard of orogenic belts, whereas SEDEX de-
composition of SEDEX and MVT ores. Both SEDEX and posits formed in intracontinental or failed rifts and rifted
MVT isotopic compositions extend over a large range; how- continental margins. Sedimentary packages that contain
ever, most SEDEX deposits have isotopic compositions that SEDEX and MVT deposits have similar stratigraphic ele-
are largely positive but range between –5 to 15 per mil. Val- ments, consisting of a basal clastic and/or volcanic-dominated
ues for MVT ores also extend over a wider range but include succession and overlain by shales and/or carbonates. SEDEX
relatively more deposits or districts with negative values. deposits are hosted mainly by reduced, fine-grained silt-
Processes that could account for the isotopically heavier re- stones-shales-mudstones and/or carbonate units within clastic
duced sulfur in SEDEX ores include closed-system BSR of sediments, whereas MVT deposits are located in carbonate-
seawater or sulfate in pore water in restricted basins, or TSR dominated sequences. Although most SEDEX and many
processes at higher temperatures; both pathways for sulfate MVT deposits were localized by extensional faults, SEDEX
reduction may have been enhanced by the abundance of or- deposits are mainly related to growth faults active at the time
ganic matter in the SEDEX environment. of mineralization.
Given the similar metal character of MVT and SEDEX de-
Deposit morphology and laminated ores in carbonate rocks posits, and the fluids that deposited them, perhaps the most
Mississippi Valley-type and SEDEX deposits typically have significant difference between the two subtypes is the depo-
distinct morphologies that set them apart. Most MVT de- sitional environment that was controlled by their respective
posits were localized by the interplay between extensional tectonic settings. Mississippi Valley-type ores formed mainly
faults and extensive carbonate dissolution features that are inboard of orogenic belts in platform carbonates, whereas
usually strata bound. Consequently, MVT deposits are usually SEDEX ores were deposited in passive margins and intra-
strata bound on a district scale but commonly discordant at a continental failed rifts. These contrasting depositional envi-
deposit scale. SEDEX ores are usually stratiform and have ronments could account for the differences in sulfur isotope
laminated and bedding-parallel sulfides that commonly com- compositions, metal endowments, deposit morphology, and
prise a series of stacked ores lenses. SEDEX deposits with their temporal distributions in the rock record.
vent-proximal ore facies have both discordant and stratiform
aspects. The distinction between SEDEX from MVT ores Concluding Remarks
based on deposit morphology is equivocal. For example, the This review points out many advances in our understanding
common occurrence of stratiform and laminated carbonate of sediment-hosted Pb-Zn deposits since the 75th Society of
replacement ores in some MVT deposits has generated de- Economic Geologists Anniversary. However, it is abundantly
bate regarding their classification as SEDEX ores. clear that we still debate many fundamental issues. As
Mississippi Valley-type ores are, by definition, hosted in Charles Behre wrote in the foreword to Economic Geology
carbonate rocks, yet some SEDEX ores are also contained in Monograph 3 on carbonate-hosted Pb-Zn deposits (Brown,
carbonates. Therefore, the presence of ore in carbonate rocks 1967, p. vii): “For at least two centuries and a corresponding
does not always provide a clear distinction. For example, the number of generations of geologists and miners there have
Anarraaq deposit in the Red Dog district formed by the sub- been active arguments concerning the origin of a certain type
sea-floor replacement of carbonate rocks. Furthermore, the of mineral deposit. How to characterize the type itself is de-
classification of some SEDEX deposits in metamorphosed batable and, indeed, objections have been raised to grouping
rocks is based solely on the stratiform nature of the ores. It is several possible examples under one heading because each
odd that there are no MVT deposits in Table A1 that are has its individually distinctive features.” The failure of the au-
metamorphosed. thors to agree on a classification scheme that adequately de-
scribes and characterizes sediment-hosted Pb-Zn ores under-
Age of ore deposition scores an issue more fundamental than simply being able to
The ages of MVT ores are generally tens of millions of years ascribe a “type” to some deposits. The most important need is
younger than their host rocks. However, a few MVT deposits to have the ability to constrain the age of the ore deposition
(e.g., Irish Midlands, Lennard Shelf, and Gays River ores) are processes using high-precision geochronology. If we were
close (<~5 m.y.) to the age of their host rocks. In the absence able to precisely date ore deposition, we could settle the is-
of direct dates for SEDEX deposits, their age of formation is sues of whether the deposit is Irish or Broken Hill type or for

0361-0128/98/000/000-00 $6.00 599


600 LEACH ET AL.

that matter, SEDEX or MVT. We must move away from ——2000, Reactive flow models of ore formation in the Southeast Missouri
model-driven debates based on textures or deposit morphol- district: ECONOMIC GEOLOGY, v. 95, p. 1605–1626.
Arne, D.C., Curtis, L.W., and Kissin, S.A., 1991, Internal zonation in a car-
ogy and toward processes constrained by the age and duration bonate-hosted Zn-Pb-Ag deposit, Nanisivik, Baffin Island, Canada: ECO-
of ore-forming events and their association with tectonic NOMIC GEOLOGY, v. 86, p. 699–717.
events. Arnold, B.Z., Bahr, J.M., and Fantucci, R., 1996, Paleohydrology of the
Upper Mississippi Valley zinc-lead district: Society of Economic Geologists
Acknowledgments Special Publication 4, p. 378–389.
Ashton, J.H., Black, A., Geraghty, J., Holdstock, M., and Hyland, E., 1992,
We especially thank Wayne Premo for his exceptional input The geological setting and metal distribution patterns of Zn-Pb-Fe miner-
to the interpretations and discussion on Pb isotopes, alization in the Navan boulder conglomerate, in Bowden, A.A., Earls, G.,
Cameron Rombach for his invaluable compilations of data, O’Connor, P.G., and Pyne, J.F., eds., The Irish Minerals Industry
Beth Hillary for her careful management of the deposit data 1980–1990: Dublin, Irish Association for Economic Geology, p. 171–210.
Ashton, J.H., Holdstock, M.P., Geraghty, J.F., O’Keefe, W.G., Martinez, N.,
base, and Stuart Bull for his contribution on the tectonic set- Peace, W., and Philcox, M.E., 2003, The Navan orebody—discovery and
ting of SEDEX deposits. Garth Graham, Erin Marsh, Sara geology of the South West Extension, in Kelly, J.C., Andrew, C.J., Ashton,
Felling, and Phillip Zelenak helped with various aspects of J.H., Boland, M.B., Earls, G., Fusciardi, L., and Stanley, G., eds., Europe’s
the manuscript. The authors gratefully acknowledge the fol- major base metal deposits: Dublin, Irish Association for Economic Geol-
lowing individuals who provided information for the compila- ogy, p. 405–430.
Ayuso, R.A., Kelley, K.D., Leach, D.L., Young, L.E., Slack, J.F., Wandless,
tion of Table A1: Arne Bjørlykke and Leif Furuhaug (Geolog- G., Lyon, A.M., and Dillingham, J.L., 2004, Origin of the Red Dog Zn-Pb-
ical Survey of Norway), Jan-Anders Perdahl (Geological Ag deposits, Brooks Range, Alaska: Evidence from regional Pb and Sr iso-
Survey of Sweden), Tony Page and Elizabeth Haynes (U.S. tope sources: ECONOMIC GEOLOGY, v. 99, p. 1533–1554.
Geological Survey), Gerry Stanley (Geological Survey of Ire- Badham, J.P.N., 1981, Shale-hosted Pb-Zn deposits: Products of exhalation of
formation waters?: Transactions of the Institution of Mining and Metal-
land), Ian Jonasson (Geological Survey of Canada), Mats lurgy, sec. B, v. 90, p. B70–B76.
Willden (Boliden Mining Co.), Richard Kyle (Univ. of Texas Bak, B., 1986, Weglanowa mineralizacja Zn, Pb, i Fe w slasko-krakowskich
at Austin), and G.M. Derrick. zlozach rud cynku i olowiu: Cracow, Poland, Archives of the Academy of
The authors appreciate the variety of contributions made to Mining and Metallurgy, unpublished report.
this manuscript by the following individuals: Colin Andrew, Banks, D.A., 1986, Hydrothermal chimneys and fossil worms from the Ty-
naugh Zn-Pb deposit, Ireland in Andrew, C.J., Crowe, R.W.A., Finlay, S.,
John Ashton, Maria Boni, Mohammed Bouabdellah, Salah Pennell, W.M., and Pyne, J.F., eds., Geology and genesis of mineral de-
Bouhlel, Dwight Bradley, Terry Briggs, Stuart Bull, Lucy posits in Ireland: Dublin, Irish Association for Economic Geology, p.
Chapman, Poul Emsbo, Leo Fusciardi, Murray Hitzman, 441–448.
Craig Johnson, Tom Lane, Joel Leventhal, Michael Lewchuk, Barley, M.E., and Groves, D.I., 1992, Supercontinent cycles and the distrib-
ution of metal deposits through time: Geology, v. 20, p. 291–294.
Ghislain de Marsily, Jean-Claude Macquar, Craig McClung, Barton, P.B., Jr., 1967, Possible role of organic matter in the precipitation of
Derrick Rhodes, Henri Rouvier, David Sinclair, David Symons, the Mississippi Valley ores: Genesis of strata-bound lead-zinc-barite-fluo-
Jacques Thibieroz, Francisco Velasco, and Jerry Zieg. Reviews rite deposits in carbonate rocks: ECONOMIC GEOLOGY MONOGRAPH 3, p.
and comments by Wayne Goodfellow, Jeff Hedenquist, John 371–377.
Thompson, and Jamie Wilkinson improved the manuscript. Basuki, N.I., and Spooner, E.T.C., 2004, A review of fluid inclusion temper-
atures and salinities in Mississippi Valley-type Zn-Pb deposits: Identifying
thresholds for metal transport: Exploration and Mining Geology, v. 11, p.
TEXT REFERENCES 1–17
Anderson, G.M., 1975, Precipitation of Mississippi Valley-type ores: ECO- Beales, F.W., 1975, Precipitation mechanisms for Mississippi Valley-type ore
NOMIC GEOLOGY, v. 70, p. 937–942. deposits: ECONOMIC GEOLOGY, v. 70, p. 943–948.
——1983, Some geochemical aspects of sulfide precipitation in carbonate Beales, F.W., and Jackson, S.A., 1966, Precipitation of lead-zinc ores in car-
rocks, in Kisvarsanyi, E., et al., eds., Proceedings of International Confer- bonate reservoirs as illustrated by Pine Point ore field, Canada: Transac-
ence on Mississippi Valley-type lead-zinc deposits: Rolla, Missouri, Univer- tions of the Institution of Mining and Metallurgy, sec. B, v. 75, p.
sity of Missouri-Rolla Press, p. 61–76. B8278–B8285.
Anderson, G.M., and Macqueen, R.W., 1982, Ore deposit models-6. Missis- Bechtel, A., Shieh, Y-N., Pervaz, M., and Puttman, W., 1996, Biodegradation
sippi Valley-type lead-zinc deposits: Geoscience Canada, v. 9, p. 107–117. of hydrocarbons and biogeochemical sulfur cycling in the salt dome envi-
Anderson, I.K., Earls, G., Hitzman, M.W., and Tear, S., eds., 1995, Irish car- ronment: Inferences from sulfur isotope and organic geochemical investi-
bonate-hosted Zn-Pb deposits: Society of Economic Geologists Guidebook gations of the Bahloul Formation at the Bou Grine Zn/Pb ore deposit,
Series, v. 21, 296 p. Tunisia: Geochimica et Cosmochimica Acta, v. 60, p. 2833–2855.
Anderson, I.K., Ashton, J.H., Boyce, A.J., Fallick, A.E., and Russell, M.J., Bethke, C.M., 1985, A numerical model of compaction-driven groundwater
1998, Ore depositional processes in the Navan Zn-Pb deposit, Ireland: flow and heat transfer and its application to the paleohydrology of intracra-
ECONOMIC GEOLOGY, v. 93, p. 535–563. tonic sedimentary basins: Journal of Geophysical Research, v. 90, p.
Andrew, C.J., 1993, Mineralization in the Irish Midlands, in Pattrick, R.A.D., 6817–6828.
and Poyla, D.A., eds., Mineralization in the British Isles: London, Chap- Bethke, C.M., and Marshak, S., 1990, Brine migrations across North Amer-
man and Hall, p. 208–269. ica—the plate tectonics of groundwater: Annual Review of Earth and Plan-
Andrew, C.J., Crowe, R.W.A., Finlay, S., Pennell, W.M., and Pyne, J, eds., etary Science, v. 18, p. 287–315.
1986, Geology and genesis of mineral deposits in Ireland: Dublin, Irish As- Betts, P.G., Giles, D., and Lister, G. S., 2003, Tectonic environment of shale-
sociation for Economic Geology, 711 p. hosted massive sulfide Pb-Zn-Ag deposits of Proterozoic northeastern Aus-
Andrews, S., 1996, Stratigraphy and depositional setting of the upper McNa- tralia: ECONOMIC GEOLOGY, v. 98, p. 557–576.
mara Group, Lawn Hill region: James Cook University, Economic Geology Billings, G.K., Kesler, S.W., and Jackson, S.E., 1969, Relation of zinc-rich for-
Research Unit Contributions 55, p. 5–9. mation waters, northern Alberta, to the Pine Point ore deposit: ECONOMIC
Ansdell, K.M., Nesbitt, B.E., and Longstaffe, F.J., 1989, A fluid inclusion and GEOLOGY, v. 64, p. 385–391.
stable isotope study of the Tom Ba-Pb-Zn deposit, Yukon Territory, Bjørlykke, A., and Sangster, D.F., 1981, An overview of sandstone lead de-
Canada: ECONOMIC GEOLOGY, v. 84, p. 841–856. posits and their relation to red-bed copper and carbonate-hosted lead-zinc
Appold, M.S., and Garven, G., 1999, The hydrology of ore formation in the deposits: ECONOMIC GEOLOGY 75TH ANNIVERSARY VOLUME, p. 179–213.
Southeast Missouri district: Numerical models of topography-driven fluid Blakeman, R.J., Ashton, J.H., Boyce, A.J., Fallick, A.E., and Russel, M.J.,
flow during the Ouachita orogeny: ECONOMIC GEOLOGY, v. 94, p. 913–936. 2002, Timing of interplay between hydrothermal and surface fluids in the

0361-0128/98/000/000-00 $6.00 600


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 601

Navan Zn + Pb orebody, Ireland: Evidence from metal distribution trends, Cathles, L.M., and Smith, A.T., 1983, Thermal constraints on the formation
mineral textures, and ?34S analyses: ECONOMIC GEOLOGY, v. 97, p. 73–91. of Mississippi Valley-type lead-zinc deposits, and their implications for
Bodon, S.B., 1996, Genetic implications of the paragenesis and rare-earth el- episodic dewatering and deposit genesis: ECONOMIC GEOLOGY, v. 78, p.
ement geochemistry at the Cannington Ag-Pb-Zn deposit, Mt. Isa inlier, 948–956.
northwest Queensland: University of Tasmania, Centre for Ore Deposit Chandler, F.W., and Grégoire, D.C., 2000, Sulfur, strontium, and boron iso-
and Exploration Studies (CODES) Special Publication 1, p. 133–144. topes from replaced sulphate evaporate nodules in the Altyn Formation,
——1998, Paragenetic relationships and their implications for ore genesis at Lower Belt Supergroup, Montana: Clues to the sedimentary environment
the Cannington Ag-Pb-Zn deposit, Mount Isa inlier, Queensland, Australia: of the Sullivan deposit: Geological Association of Canada, Mineral Deposit
ECONOMIC GEOLOGY, v. 93, p. 1463–1488. Division Special Publication 1, p. 251–258.
Bouabdellah, M., Brown, A.C., and Sangster, D.F., 1996, Mechanisms of for- Chapman, L.H., 2004, Geology and mineralization styles of the George
mation of internal sediments at the Beddiane lead-zinc deposit, Toussit Fisher Zn-Pb-Ag deposit, Mount Isa, Australia: ECONOMIC GEOLOGY, v. 99,
mining district, northeastern Morocco: Society of Economic Geologists p. 233–256.
Special Publication 4, p. 112–143. Chapman, L.H., and Williams, P.J., 1998, Evolution of pyroxene-pyroxenoid-
Bowden, A.A., Earls, G., O’Connor, P.G., and Pyne, J.F., eds., 1992, The Irish garnet alteration at the Cannington Ag-Pb-Zn deposit, Cloncurry district,
Minerals Industry 1980–1990: Dublin, Irish Association for Economic Queensland, Australia: ECONOMIC GEOLOGY, v. 93, p. 1390–1405.
Geology. Chi, G., and Savard, M., 1997, Sources of basinal and Mississippi Valley-type
Boyce, A.J., Coleman, M.L., and Russel, M.J., 1983, Formation of fossil hy- mineralizing brines; mixing of evaporated seawater and halite-dissolution
drothermal chimneys and mounds from Silvermines, Ireland: Nature, v. brine: Chemical Geology, v. 143, no. 3–4, p. 121–125.
306, p. 545–550. Christensen, J.N., Halliday, A.N., Vearncombe, J., and Kesler, S.E., 1995,
Boyce, A.J., Little, C.T.S., and Russell, M.J., 2003, A new fossil vent biota in Testing models of large-scale crustal fluid flow using direct dating sulfides:
the Ballynoe barite deposit, Silvermines, Ireland: Evidence for intracra- Rb-Sr evidence for early dewatering and formation of MVT deposits, Can-
tonic sea-floor hydrothermal activity about 352 Ma: ECONOMIC GEOLOGY, ning basin, Australia: Geology, v. 90, p. 877–884.
v. 98, p. 649–656. Clendenin, C.W., 1993, Faults, fluids, and Southeast Missouri MVT deposits
Bradley, D.C., 1993, Role of lithospheric flexure and plate convergence in [abs.]: Geological Society of America Abstracts with Programs, v. 25, no. 3,
the genesis of some Appalachian zinc deposits: U.S. Geological Survey Bul- p. 12.
letin, v. 2039, p. 35–43. Clendenin, C.W., and Duane, M.J., 1990, Focused fluid flow and Ozark Mis-
Bradley, D.C., and Leach, D.L., 2003, Tectonic controls of Mississippi Val- sissippi Valley-type deposits: Geology, v. 18, p. 116–119.
ley-type lead-zinc mineralization in orogenic forelands: Mineralium De- Clendenin, C.W., Niewendrop, C.A., Duane, M.J., and Lowell, G.R., 1994,
posita v. 38 p. 652–667. The paleohydrology of Southeast Missouri Mississippi valley-type deposits,
Bradley, D.C., Leach, D.L., Symons, D., Emsbo, P., Premo, W., Breit, G., interplay of faults, fluids, and adjoining lithologies: ECONOMIC GEOLOGY, v.
and Sangster, D.F., 2004, Reply to Discussion on “Tectonic controls of Mis- 89, p. 322–332.
sissippi Valley-type lead-zinc mineralization in orogenic forelands” by S.E. Condie, K.C., 2000, Episodic continental growth models: Afterthoughts and
Kesler, J.T. Chesley, J.N. Christenson, R.D. Hagni, W. Heijlen, J.R. Kyle, P. extensions: Techtonophysics, v. 322, p. 153–162.
Munchez, K.C. Misra, and R. van der Voo: Mineralium Deposita, v. 39, p. Cooke, D.R., Bull, S.W., Large, R.R., and McGoldrick, P.J., 2000, The im-
515–519. portance of oxidized brines for the formation of Australian Proterozoic
Braisier, M.D., and Lindsay, J.F., 1998, A billion years of environmental sta- stratiform sediment-hosted Pb-Zn (Sedex) deposits: ECONOMIC GEOLOGY,
bility and emergence of eukaryotes, new data from northern Australia: Ge- v. 95, p. 1–18.
ology, v. 26, p. 555–558. Corbella, M., Ayora, C., and Cardellach, E., 2004, Hydrothermal mixing, car-
Brannon, J.C., Podosek, F.A., and Cole, S.C., 1996, Radiometric dating of bonate dissolution and sulfide precipitation in Mississippi Valley-type de-
Mississippi Valley-type ore deposits: Society of Economic Geologists Spe- posits: Mineralium Deposita, v. 39, p. 344–357.
cial Publication 4, p. 546–554. Deb, M., and Pal, T., 2004, Geology and genesis of the base metal sulphide
Bresser, H.A., 1992, Origin of base metal vein mineralization in the Lawn deposits in the Dariba-Rajpura-Bethumni belt, Rajasthan, India, in the
Hill mineral field, northwestern Queensland: Unpublished B.Sc. Honors light of basin evolution, in Deb, M., and Goodfellow, W.D., eds., Sediment
thesis, Townsville, James Cook University, 115 p. hosted lead-zinc sulphide deposits: Attributes and models of some major
Broadbent, G.C., Myers, R.E., and Wright, J.V., 1998, Geology and origin of deposits in India, Australia and Canada: New Delhi, India, Narosa Pub-
shale-hosted Zn-Pb-Ag mineralization at the Century deposit, northwest lishing House, p. 304–327.
Queensland, Australia: ECONOMIC GEOLOGY, v. 93, p. 1264–1294. Deloule, E., Allewgre, C., and Doe, B., 1986, Lead and sulfur isotope mi-
Broadbent, G.C., Andrews, S.J., and Kelso, I.J., 2002, A decade of new ideas: crostratigraphy in galena crystals from Mississippi Valley-style deposits:
Geology and exploration history of the Century Zn-Pb-Ag deposits, north- ECONOMIC GEOLOGY, v. 81, p. 1307–1321.
eastern Queensland, Australia: Society of Economic Geologists Special Deming, D., and Nunn, J.A., 1991, Numerical simulations of brine migration
Publication 9, p. 119–140. by topographically driven recharge: Journal of Geophysical Research, v. 96,
Brockie, D.C., Hare, E.H., Jr., and Dingess, P.R., 1968, The geology and ore de- p. 2485–2499.
posits of the Tri-State district of Missouri, Kansas, and Oklahoma, in Ridge, Diehl, S.F., Goldhaber, M.B., and Mosier, E.L., 1989, Regions of feldspar
J.D., ed., Ore deposits of the United States, 1933–1967: New York, American precipitation and dissolution in the Lamotte Sandstone, Missouri—impli-
Institute of Mining, Metallurgical, and Petroleum Engineers, p. 400–430. cations for MVT ore genesis [abs.]: U.S. Geological SurveyOpen-File Re-
Brown, J.S., 1967, Isotopic zoning of lead and sulfur in southeast Missouri: port 89–169, p. 5–7.
ECONOMIC GEOLOGY MONOGRAPH 3, p. 410–426. Donnelly, T.H., and Jackson, M.J., 1988, Sedimentology and geochemistry of
——1970, Mississippi Valley-type lead-zinc ores: A review and sequel to the a mid-Proterozoic lacustrine unit from northern Australia: Sedimentary
“Behre Symposium:” Mineralogical Deposita, v. 5, p. 103–119. Geology, v. 58, p.145–169.
Callahan, J.S., 1964, Paleogeographic premises for prospecting for strata- Doyle, E., and Bowden, A.A., 1995, Field guide to the Galmoy zinc-lead de-
bound base metal deposits in carbonate rocks: CENTO Symposium on posit: Society for Economic Geologists Guidebook Series, v. 21, p.
mining geology and base metals, Ankara, Turkey, Proceedings, p. 191–248. 139–145.
Carne, R.C., and Cathro, R.J., 1982, Sedimentary exhalative (sedex) zinc- Duane, M.J., Kruger, F.J., Turner, A.M., Whitelaw, H.T., Coetzee, H., and
lead-silver deposits, northern Canadian Cordillera: Canadian Mining and Verhagen, B., 2004, The timing and isotopic character of regional hy-
Metallurgical Bulletin, v. 75, no. 840, p. 66–78. drothermal alteration and associated epigenetic mineralization in the west-
Carpenter, A.B., 1978, Origin and chemical evolution of brines in sedimen- ern sector of the Kaapvaal craton (South Africa): Journal of African Earth
tary basins: Oklahoma Geological Survey Circular 79, p. 60–77. Sciences, v. 38, p. 461–476.
Carpenter, A.B., Trout, M.L. and Pickett, E.E., 1974, Preliminary report on Dumoulin, J.A., Harris, A.G., Blome, C.D., and Young, L.E, 2004, Deposi-
the origin and evolution of lead- and zinc-rich oil field brines in Central tional settings, correlation, and age of Carboniferous rocks in the western
Mississippi: ECONOMIC GEOLOGY, v. 69, p. 1191–1206. Brooks Range, Alaska: ECONOMIC GEOLOGY, v. 99, p. 1355–1384.
Cathles, L.M., III, and Adams, J.J., 2005, Fluid flow and petroleum and min- Dunham, K.C., 1966, Role of juvenile solutions, connate waters and evapor-
eral resources in the upper (<20-km) continental crust: ECONOMIC GEOL- itic brines in the genesis of lead zinc-fluorine-barium deposits: Transac-
OGY 100TH ANNIVERSARY VOLUME, p. 000. tions, Institution of Mining and Metallurgy, sec. B, v. 75, p. B226–B229.

0361-0128/98/000/000-00 $6.00 601


602 LEACH ET AL.

Edgerton, D., 1997a, Geologic models of sediment-buffered hydrothermal Garven, G., Bull, S.W., and Large, R.R., 2001, Hydrothermal fluid flow mod-
vents: A case study of the Red Dog Zn-Pb-Ag orebody, western Brooks els of stratiform ore genesis in the McArthur basin, Northern Territory,
Range, Alaska: Unpublished Ph.D. dissertation, Austin, University of Texas, Australia: Geofluids, v. 1, p. 289–311.
209 p. Garven, G., Raffensperger, J.P., Dumoulin, J.A., Bradley, D.A., Young, L.E.,
——1997b, Reconstruction of the Red Dog Zn-Pb-Ba orebody, Alaska: Im- Kelley, K.D., and Leach, D.L., 2003, Coupled heat and fluid flow model-
plications for the vent environment during the mineralizing event: Cana- ing of the Carboniferous Kuna basin, Alaska: Implications for the genesis
dian Journal of Earth Sciences, v. 34, p. 1581–1602. of the Red Dog Pb-Zn-Ag-Ba ore district [abs.]: Journal of Geochemical
Eldridge, C.S., Williams, N., and Walshe, J.L., 1993, Sulfur isotope variabil- Exploration, v. 78–79, p. 215–219.
ity in sediment-hosted massive sulfide deposits as determined using the ion Geer, K.A., 1988, Geochemistry of the stratiform zinc-lead-barite mineral-
microprobe SHRIMP II: A study of the HYC deposit at McArthur River, ization at the Meggen mine, Federal Republic of Germany: Unpublished
Northern Territory, Australia: ECONOMIC GEOLOGY, v. 88, p. 1–26. Ph.D. thesis, University Park, PA, Pennsylvania State University, 176 p.
Emsbo, P., 2000, Gold in Sedex deposits: Society of Economic Geologists Gerdemann, P.E., and Meyers, H.E., 1972, Relationships of carbonate facies
Reviews, v. 13, p. 427–437. patterns to ore distribution and to ore genesis in the Southeast Missouri
Erickson, R.L., Mosier, E.L., Viets, J.G., Odland, S.K., and Erickson, M.S., lead district: ECONOMIC GEOLOGY, v. 67, p. 426–433.
1983, Subsurface geochemical exploration in carbonate terrane-midconti- Ghazban, F., Schjwarcz, H.P., and Ford, D.C., 1990, Carbon and sulfur iso-
nent United States, in Kisvarsanyi, G., et al., eds., Proceedings of Interna- tope evidence for in situ reduction of sulfate, Nanisivik lead-zinc deposits,
tional Conference on Mississippi Valley-type lead-zinc deposits: Rolla, Uni- Northwest Territories, Baffin Island, Canada: ECONOMIC GEOLOGY, v. 85,
versity of Missouri, Rolla Press, p. 575–583. p. 360–375.
Everett, C.E., Wilkinson, J.J., and Rye, D.M., 1999, Fracture-controlled fluid Goodfellow, W.D., 2000, Anoxic conditions in the Aldridge basin during for-
in the Lower Paleozoic basement rocks of Ireland: Implications for the mation of the Sullivan Zn-Pb deposit: Implications for the genesis of mas-
genesis of Irish-type Zn-Pb deposits: Geological Society of London Special sive sulphides and distal hydrothermal sediments: Geological Association of
Publications, v. 155, p. 247–276. Canada, Mineral Deposits Division Special Publication 1, p. 218–250.
Fallick, A.E., Ashton, J.H., Boyce, A.J., Ellam, R.M., and Russell, M.J., 2001, ——2004, Geology, genesis and exploration of SEDEX deposits, with em-
Bacteria were responsible for the magnitude of the world-class hydrother- phasis on the Selwyn basin, Canada., in Deb, M., and Goodfellow, W.D.,
mal base metal orebody at Navan, Ireland: ECONOMIC GEOLOGY, v. 96, p. ed., Sediment hosted lead-zinc sulphide deposits: Attributes and models of
883–846. some major deposits in India, Australia and Canada: New Delhi, India,
Fehn, U., Cathles, L.M., and Holland, H.D., 1978, Hydrothermal convection Narosa Publishing House, p. 24–99.
and uranium deposits in abnormally radioactive plutons: ECONOMIC GEOL- Goodfellow, W.D., and Jonasson, I.R., 1987, Environment of formation of the
OGY, v. 73, p. 1556–1566. Howards Pass (XY) Zn-Pb deposit, Selwyn basin, Yukon: Canadian Institute
Fontboté, L., and Boni, M., eds., 1994, Sediment-hosted Zn-Pb ores: Ger- of Mining and Metallurgy Special Volume 37, p. 19–50.
many, Springer Verlag, Society for Geology Applied to Mineral Deposits, Goodfellow, W.D., and Rhodes, D., 1990, Geological setting, geochemistry
Special Publication 10, 480 p. and origin of the Tom stratiform Zn-Pb-Ag-barite deposits: Geological Sur-
Forrest, K., 1983, Geologic and isotopic studies of the Lik deposit and the vey of Canada Open File 2169, p. 177–241.
surrounding mineral district, DeLong Mountains, western Brooks Range, Goodfellow, W.D., Jonasson, I.R., and Morganti, J.M., 1983, Zonation of chal-
Alaska: Unpublished Ph.D. dissertation, Minneapolis, University of Min- cophile elements about the Howard’s Pass (XY) Zn-Pb deposit, Selwyn
nesota, 161 p. basin, Yukon, in Parslow, G.R., ed., Geochemical exploration: Journal of
Frimmel, H.E., 2001, Geodynamic and paleoclimate setting of the Neopro- Geochemical Exploration 1982: Elsevier, Amsterdam-New York, p. 503–542
terozoic Rosh Pinah Zn-Pb province, southwestern Namibia, in Piestrzyn- Goodfellow, W.D., Lydon, J.W., and Turner, R.J.W., 1993, Geology and gen-
ski, ed., Mineral deposits at the beginning of the 21st century, Proceedings esis of stratiform sediment-hosted (SEDEX) zinc-lead-silver sulphide de-
of the Joint Sixth Biennial SGA-SEG Meeting, Krakow, Poland, August posits: Mineral deposit modeling: Geological Association of Canada Special
26–29: Lisse, The Nederlands, Swets and Zeitlinger Publishing, p. 129–132. Paper 40, p. 201–251.
Fusciardi, L.P., Güven, J.F., Stewart, D.R.A., Carboni, V., and Walsh, J.J., Gregg, J.M., 1985, Regional epigenetic dolomitization in the Bonneterre
2003, The geology and genesis of the Lisheen Zn-Pb deposit, Co. Tipper- Dolomite (Cambrian), southeastern Missouri: Geology, v. 13, p. 503–506.
ary, Ireland, in Kelly, J.G., Andrew, J.H., Boland, M.B., Earls, G., Fuscia- Greyling, L.N., 2000, The Paleoproterozoic carbonate-hosted Pering lead-
rdi, L., and Stanley, G., eds., Europe’s major base metal deposits: Dublin, zinc deposit, South Africa: Unpublished M.Sc. thesis, Johannesburg, Rand
Irish Association for Economic Geology, p. 455–476. Afrikaans University, 129 p.
Gallagher, V, Boyce, A.J., Fallick, A.E., and Mohr, P.J., 1992, An isotopic Grotzinger, J.P., 1989, Facies and evolution of Precambrian depositional sys-
study of the Harberton Bridge Fe-Zn-Pb deposit, County Kildare, and its tems: Emergence of the modern platform archetype: Society for Sedimen-
implication for metallogenesis in Ireland, in Bowden, A. A., Earls, G., O’- tary Geology Special Publication 44, p. 79–106.
Connor, P. G., and Pyne, J. F., eds., The Irish Minerals Industry 1980–1990: Grotzinger, J.P., and James, N.P., 2000, Precambrian carbonates: Evolution
Dublin, Irish Association for Economic Geology, p. 261–272. of understanding: Society for Sedimentary Geology Special Publication 67,
Gardner, H.D., and Hutcheon, I., 1985, Geochemistry, mineralogy, and ge- p. 3–20.
ology of the Jason Pb-Zn deposits, Macmillan Pass, Yukon, Canada: ECO- Gummer, P.K., Plint, H.E., and Rainbird, R.H., 1997, The Esker Lake
NOMIC GEOLOGY, v. 80, p. 1257–1276. prospect: Stratabound Pb-Zn-Cu-Ag in emergent inner shelf carbonates,
Garven, G., 1985, The role of regional fluid flow in the genesis of the Pine Rocknest Formation, Coronation Supergroup, NWT: Yellowknife, NWT
Point deposit, Western Canada sedimentary basin: ECONOMIC GEOLOGY, v. Geological Mapping Division, Exploration Overview, v. 1996, p. 3.18–3.19.
80, p. 307–324. Gustafson, L.B., and Williams, N., 1981, Sediment-hosted stratiform de-
——1995, Continental-scale groundwater flow and geologic processes: An- posits of copper, lead, and zinc: ECONOMIC GEOLOGY 75TH ANNIVERSARY
nual Review of Earth and Planetary Science, v. 23, p. 89–117. VOLUME, p. 139–178.
Garven, G., and Freeze, R.A., 1984, Theoretical analysis of the role of Gwosdz, W., and Krebs, W., 1977, Manganese halo surrounding Meggen ore
groundwater flow in the genesis of strata-bound ore deposits. 1. Mathe- deposit, Germany: Transactions of the Institution of Mining and Metal-
matical and numerical model, 2. Quantitative results: American Journal of lurgy, sec. B, v. 86, p. B73–B77.
Science, v. 284, p. 1085–1174. Hagni, R.D., 1983, Minor elements in Mississippi Valley-type ore deposits, in
Garven, G., and Raffensperger, J.P., 1997, Hydrogeology and geochemistry of Shanks, W.C., ed., Cameron volume on unconventional mineral deposits:
ore genesis in sedimentary basins, in Barnes, H.L., ed., Geochemistry of hy- Society of Economic Geology and Society of Mining Engineers, AIME,
drothermal ore deposits, 3rd ed.: New York, John Wiley and Sons, p.125–189. New York, p. 44–88.
Garven, G., Ge, S., Person, M.A., and Sverjensky, D.A., 1993, Genesis of Hall, C.M., York, D., Saunders, C.M., and Strong, D.F., 1989, Laser 40Ar/39Ar
strata-bound ore deposits in the midcontinent basins of North America. 1. dating of Mississippi Valley-type mineralization from western Newfound-
The role of regional groundwater flow: American Journal of Science, v. 293, land [abs.]: International Geological Congress, Abstracts-Congres Ge-
p. 497–568. ologique Internationale, Resumes, v. 28, p. 2.10–2.11.
Garven, G., Appold, M.S., Toptygina, V.I., and Hazlett, T.J., 1999, Hydroge- Hall, W.E., and Friedman, I., 1963, Composition of fluid inclusions Cave-in-
ologic modeling of the genesis of carbonate-hosted lead-zinc ores: Hydro- Rock fluorite district, Illinois, and Upper Mississippi Valley lead-zinc dis-
geology Journal, v. 7, p. 108–126. trict: ECONOMIC GEOLOGY, v. 58, p. 886–911.

0361-0128/98/000/000-00 $6.00 602


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 603

Hamilton, J.M., Bishop, D.T., Morris, H.C., and Owens, O.E., 1982, Geology Hudson, M.R., 2000, Coordinated strike-slip and normal faulting in the
of the Sullivan orebody, Kimberley, B.C., Canada: Geological Association of southern Ozark dome of northern Arkansas: Deformation in a late Paleo-
Canada, Special Paper 25, p. 597–665. zoic foreland: Geology, v. 28, p. 511–514.
Hanor, J.S., 1979, The sedimentary genesis of hydrothermal fluids, in Barnes, Hunt, J.M., 1996, Petroleum geochemistry and geology: New York, Free-
H.L., ed., Geochemistry of hydrothermal ore deposits: New York, Wiley- man, 743 p.
Interscience, p. 137–142. Huston, D.L., and Logan, G.A., 2004, Barite, BIFs and bugs: Evidence for
——1994, Origin of saline fluids in sedimentary basins: Geological Society the evolution of the Earth’s early hydrosphere: Earth and Planetary Science
Special Publications, v. 78, p. 151–174. Letters, v. 220, p. 41–55.
——1996, Controls on the solubilization of lead and zinc in basin brines: So- Ireland, T. Bull, S.W, and Large, R.R., 2004, Mass flow sedimentology within
ciety of Economic Geologists Special Publication 4, p. 483–500. the HYC Zn-Pb-Ag deposit, Northern Territory, Australia: Evidence for
Hay, R.L., Lee, M., Kolata, D.R., Matthews, J.C., and Morton, J.P., 1988, syn-sedimentary ore genesis: Mineralium Deposita v. 39, p. 143–158.
Episodic potassic diagenesis of Ordovician tuffs in the Mississippi Valley Jackson, S.A., 1966, A study of Mississippi Valley-type lead-zinc mineraliza-
area: Geology, v. 16, p. 743–747. tion with special reference to sediment diagenesis: Unpublished master’s
Hazlett, T.J., 1997, A hydrogeologic analysis of the Irish carbonate-hosted thesis, Toronto, Ontario, University of Toronto, 93 p.
lead-zinc ore deposits: Unpublished Ph.D. thesis, Baltimore, Johns Hop- Jankovic, S., 1986, The mineral association and genesis of the lead-zinc-barite
kins University, 322 p. deposit at Gunga, Khuzdar district, Baluchistan, Pakistan: Records of the
Hearn, P.P., Jr., Sutter, J.F., and Belkin, H.E., 1987, Evidence for late-Paleo- Geological Survey of Pakistan Report 71, 11 p.
zoic brine migration in Cambrian carbonate rocks of the central and south- Jefferson, C.W., Kilby, D.B., Pigage, L.C., and Roberts, W.J., 1983, The
ern Appalachians: implications for Mississippi Valley-type sulfide mineral- Cirque barite-zinc-lead deposits, northeastern British Columbia: Miner-
ization: Geochimica et Cosmochimica Acta, v. 51, p. 1323–1334. alogical Association of Canada Short Course Handbook, v. 8, p. 121–139.
Heijlen, W., Muchez, P., Banks, D.A., Schneider, J., Kucha, H., and Keppens, Johnston, J.D., 1999, Regional fluid flow and the genesis of Irish Carbonif-
E., 2003, Carbonate-hosted Zn-Pb deposits in Upper Silesia, Poland: Ori- erous base metal deposits: Mineralium Deposita, v. 34, p. 571–598.
gin and evolution of mineralizing fluids and constraints on genetic models: Kelley, K.D., Dumoulin, J.A., and Jennings, S., 2004a, The Anarraaq Zn-Pb-
ECONOMIC GEOLOGY, v. 98, p. 911–932. Ag and barite deposit, northern Alaska: Evidence for replacement of car-
Heyl, A.V., 1983, Geologic characteristics of three Mississippi Valley-type dis- bonate by barite and sulfides: ECONOMIC GEOLOGY, v. 99, p. 1577–1591.
tricts, in Kisvarsanyi G, et al., eds., Proceedings of International Confer- Kelley, K.D., Leach, D.L., Johnson, C.A., Clark, J.L., Fayek, M., Slack, J.F.,
ence on Mississippi Valley-type lead-zinc deposits: Rolla, Missouri, Univer- Anderson, V.M., Ayuso, R.A., and Ridley, W.I., 2004b, Textural, composi-
sity of Missouri, Rolla Press, p. 27–30. tional, and sulfur isotope variations of sulfide minerals in the Red Dog Zn-
Heyl, A.V., Agnew, A.F., Lyons, E.J., and Behre, C.H., 1959, The geology of Pb-Ag deposits, Brooks Range, Alaska: Implications for ore formation:
the Upper Mississippi Valley zinc-lead district: U.S. Geological Survey Pro- ECONOMIC GEOLOGY, v. 99, p. 1509–1532.
fessional Paper 424–D, 310 p. Kelly, J.G., Andrew, C.J., Ashton, J.H., Boland, M.B., Earls, G., Fusciardi, L.,
Heyl, A.V., Landis, G.P. and Zartman, R.E., 1974, Isotopic evidence for the and Stanley, G., eds., 2003, Europe’s major base metal deposits: Dublin,
origin of Mississippi Valley-type mineral deposits: A review: ECONOMIC GE- Ireland, Irish Association for Economic Geology, 552 p.
OLOGY, v. 69, p. 992–1006. Kesler, S.E., 1996, Appalachian Mississippi Valley-type deposits: Pale-
Hitzman, M.W., 1986, Geology of the Abbeytown mine, Co. Sligo, Ireland, in oaquifers and brine provinces: Society of Economic Geologists Special
Andrew, C.J., Crowe, R.W.A., Finlay, S., Pennell, W.M., and Pyne, J.F., Publication 4, p. 29–57.
eds., Geology and genesis of mineral deposits in Ireland: Dublin, Irish As- Kesler, S.E., and Carrigan C.W., 2002, Discussion on “Mississippi Valley-type
sociation for Economic Geology, p. 341–354. lead-zinc deposits through geological time: Implications from recent age-
——1995, Mineralization in the Irish Zn-Pb-(Ba-Ag) orefield: Society for dating research:” Mineralium Deposita, v. 37, p. 800–802.
Economic Geologists Guidebook Series, v. 21, p. 25–61. Kesler, S.E., Martini, A.M., Appold, M.S., Walter, L.M., Huston, T.J., and
——1996, Hydrothermal alteration associated with Irish type zinc-lead-(sil- Furman, F.C., 1996, Na-Cl-Br systematic of fluid inclusions from Mis-
ver) and carbonate-hosted copper deposits [abs.]: Geological Society of sissippi Valley-type deposits, Appalachian basin: Constraints on solute
America Abstracts with Programs, v. 28, no. 7, p. 23. origin and migration paths: Geochimica Cosmochimica Acta, v. 60, p.
——1999, Extensional faults that localized syndiagenetic Zn-Pb deposits and 225–233.
their reactivation during Variscan compression: Geological Society of Lon- Kesler, S.E., Chesley, J.T., Christensen, J.N., Hagni, R.D., Heijlen, W., Kyle,
don Special Publication 155, p. 233–245. J.R., Misra, K.C., Muchez, P., and van der Voo, R., 2004, Discussion of
Hitzman, M.W., and Beaty, D.W., 1996, The Irish Zn-Pb-(Ba) orefield: Soci- “Tectonic controls of Mississippi Valley-type lead-zinc mineralization in
ety of Economic Geologists Special Publication 4, p. 112–143. orogenic forelands” by D.C. Bradley and D.L. Leach (2003): Mineralium
Hitzman, M.W., and Large, D., 1986, A review and classification of the Irish Deposita, v. 34, p. 512–514.
carbonate-hosted base metal deposits, in Andrew, C.J., Crowe, W.A., Pen- Kharaka, Y.K., Maest, A.S., Carothers, W.W., Law, L.M., Lamothe, P.J., and
nell, W.M., and Pyne, J.F., eds., Geology and genesis of mineral deposits in Fries, T.L., 1987, Geochemistry of metal-rich brines from central Missis-
Ireland: Dublin, Irish Association for Economic Geology, p. 217–238. sippi Salt Dome basin, U.S.A.: Applied Geochemistry, v. 2, p. 543–561.
Hitzman, M.W., O’Connor, P., Shearly, E., Schaffalitzky, C., Beaty, D.W., Kibitlewski, S., 1991, Tectonic control of the origin of the Zn-Pb deposits in
Allan, J.R. and Thompson, T., 1992, Discovery and geology of the Lisheen the Chrzanow region, Poland: Geological Quarterly, v. 37, p. 229–240.
Zn-Pb-Ag prospect, Rathdowney Trend, Ireland, in Bowden, A.A., Earls, Kisvarsanyi, G., Grant, S.K., Pratt, W.P., and Koenig, J.W., eds., 1983, Pro-
G., O’Connor, P.G., and Pyne, J.F., eds., The Irish Minerals Industry ceedings of International Conference on Mississippi Valley-type lead-zinc
1980–1990: Dublin, Irish Association for Economic Geology, p. 227–246. deposits: Rolla, University of Missouri, 603 p.
Hitzman, M.W., Layer, P.W., and Newberry, R.J., 1994, Argon-argon- Krebs, W., and Macqueen, R., 1984, Sequence of diagenetic and mineraliza-
stepheating studies of muscovite in the Upper Devonian Old Red Sand- tion events, Pine Point lead-zinc property, Northwest Territories, Canada:
stone: The first absolute dates for the age of Irish zinc-lead mineralization Canadian Petroleum Society Bulletin, v. 32, p. 434–464.
[abs.]: Geological Society of America Abstracts with Programs, v. 26, p. 381. Kusky, T.M., and Hudleston, P.J., 1999, Growth and demise of an Archean
Hitzman, M.W., Redmond, P.B., and Beaty, D.W., 2003, The carbonate- carbonate platform, Steep Rock Lake, Ontario, Canada: Canadian Journal
hosted Lisheen Zn-Pb-Ag deposit, County Tipperary, Ireland: ECONOMIC of Earth Sciences, v. 36, p. 565–584.
GEOLOGY, v. 97, p. 1627–1655. Kyle, J.R., 1981, Geology of the Pine Point lead-zinc district, in Wold, K.H.,
Hoagland, A.D., 1976, Appalachian zinc-lead deposits, in Wolf, K.H., ed., ed., Handbook of strata-bound and stratiform ore deposits: New York, El-
Handbook of strata-bound and stratiform ore deposits: Amsterdam, Else- sevier, v. 9, p. 643–741.
vier, v. 6, p. 495–534. Kyle, J.R., and Saunders, J.A., 1996, Metallic deposits of the Gulf Coast
Hobbs, B.E., Walshe, J.L., Ord, A., Zhang, Y., and Carr, G.C., 1998, The Bro- basin: Diverse mineralization styles in a young sedimentary basin: Society
ken Hill orebody: A high temperature, high pressure scenario: Australian of Economic Geologists Special Publication 4, p. 218–229.
Geological Survey Organisation Record 1998/2, p. 98–103. Lambert, I.B., and Scott, K.M., 1973, Implications of geochemical investiga-
Hoy, T., Brown, D., Lydon, J., and others, 2000, Sullivan and other Zn-Pb de- tions of sedimentary rocks within and around the McArthur zinc-lead-sil-
posits, southeastern British Columbia and northeastern Washington: Geo- ver deposit, northern Territory: Journal of Geochemical Exploration, v. 2,
Canada 2000 Field Trip Guidebook, no. 17, 132 p. p. 307–330.

0361-0128/98/000/000-00 $6.00 603


604 LEACH ET AL.

Land, L.S., and Macpherson, G.L., 1992, Origin of saline formation waters, Leach, D.L., Bechstadt, T., Boni, M., and Zeeh, S., 2003, Triassic-hosted
Cenozoic section, Gulf of Mexico sedimentary basin: American Association MVT Zn-Pb ores of Poland, Austria, Slovenia and Italy, in Kelly, J.G., An-
of Petroleum Geologists Bulletin, v. 76, p. 1344–1362. drew, C.J., Ashton, J.H., Boland, M.B., Earls, G., Fusciardi, L., and Stan-
Large, D.E., 1983, Sediment-hosted massive sulfide lead-zinc deposits: An ley, G., eds., Europe’s major base metal deposits: Dublin, Irish Association
empirical model: Mineralogical Association of Canada Short Course Hand- for Economic Geology, p. 169–214.
book, v. 8, p. 1–30. Leach, D.L., Marsh, E., Emsbo, P., Rombach, C., Kelley, K.D., Reynolds, J.,
Large, D., and Walcher, E., 1999, The Rammelsberg massive sulphide Cu- and Anthony, M., 2004, Nature of hydrothermal fluids at the shale-hosted
Zn-Pb-Ba-deposit, Germany: An example of sediment-hosted, massive sul- Red Dog Zn-Pb-Ag deposits, Brooks Range, Alaska: ECONOMIC GEOLOGY,
phide mineralization: Mineralium Deposita, v. 34, p. 522–538. v. 99, p. 1449–1480.
Large, D.E., 1980, Sediment-hosted massive sulphide lead-zinc deposits: An Leach, D.L., Bradley, D., Gardoll, S., Huston, D., Marsh, E., 2005, The dis-
empirical model: Mineralogical Association of Canada Short Course Hand- tribution of SEDEX Pb-Zn deposits through Earth history [ext. abs.]: Bi-
book, v. 8, p. 1–29. enniel SGA Meeting, 8th, Beijing, China, Aug. 21–25, Extended Abstracts,
Large, R.R., and McGoldrick, P.J., 1998, Lithogeochemical halos and geo- in press.
chemical vectors to stratiform sediment hosted Zn-Pb-Ag deposits: 1. Lady Lebedev, L.M., 1973, Minerals of contemporary hydrotherms of Cheleken:
Loretta deposit, Queensland: Journal of Geochemical Exploration, v. 63, p. Geochemistry International, v. 9, p. 485–504.
37–56. Leitch, C.H.B., and Lydon, J.W., 2000, Fluid inclusion petrography and mi-
Large, R.R., Bodon, S., Davidson, G., and Cooke, D., 1996, The chemistry of crothermometry of the Sullivan deposit and surrounding area: Geological
Broken Hill-type ore formation—one of the keys to understanding the dif- Association of Canada, Mineral Deposits Division Special Volume 1, p.
ferences between SEDEX and Broken Hill-type deposits: University of 617–632.
Tasmania, Centre for Ore Deposit and Exploration Studies (CODES) Spe- Leitch, C.H.B., Turner, R.J.W., Ross, K.V., and Shaw, D.R., 2000, Wallrock
cial Publication 1, p. 105–112. Alteration at the Sullivan deposit, British Columbia, Canada: Geological
Large, R.R., Bull, S.W., Cooke, D.R., McGoldrick, P.J., 1998, A genetic Association of Canada, Mineral Deposits Division Special Publication 1, p.
model for the HYC deposit, Australia: Based on regional sedimentology, 633–651.
geochemistry, and sulfide-sediment relationships: ECONOMIC GEOLOGY, v. Leventhal, J.S., 1990, Organic matter and thermochemical sulfate reduction
93, p. 1345–1368. in the Viburnum Trend, Southeast Missouri: ECONOMIC GEOLOGY, v. 85, p.
Large, R.R., Bull, S.W., and McGoldrick, P.J., 2000, Lithogeochemical halos 622–632.
and geochemical vectors to stratiform sediment hosted Zn-Pb-Ag deposits. Lowther, J.M., Balding, A.B., Bowden, A.A., Dunphy, S., and McEvoy, F.M.,
Part 2. HYC Deposit, McArthur River, Northern Territory: Journal of Geo- 1999, The Galmoy mine: Some geological aspects of development from
chemical Exploration, v. 68, p. 105–126. 1995 to 1999, in Stanley, C.J., et al., eds., Mineral deposits: Processes to
Large, R.R., Bull, S.W., and Winefield, P.R., 2001, Carbon and oxygen iso- processing: Rotterdam, Balkema, v. 5, p. 881–884.
tope halo in carbonates related to the McArthur River (HYC) Zn-Pb-Ag de- Lydon, J.W., 1983, Chemical parameters controlling the origin and deposi-
posit: Implications for sedimentation, ore genesis, and mineral exploration: tion of sediment-hosted stratiform lead-zinc deposits: Mineralogical Asso-
ECONOMIC GEOLOGY, v. 96, p. 1567–1593. ciation of Canada Short Course Handbook, v. 9, p. 175–250.
Large, R.R., Bull, S.W., Yang, J., Cooke, D.R., Garven, G., McGoldrick, P.J., ——1996, Sedimenary exhalative sulphides (SEDEX): Geological Survey of
and Selley, D., 2002, Controls on the formation of giant stratiform sedi- Canada, Geology of Canada, no. 8, p. 130–152.
ment-hosted Zn-Pb-Ag deposits with particular reference to the north Aus- ——2004a, Genetic models for Sullivan and other SEDEX deposits, in Deb,
tralian: University of Tasmania, Centre for (CODES) Special Ore Deposit M., and Goodfellow, W.D., eds., Sediment-hosted lead-zinc sulfide de-
and Exploration Studies Publication 4, p. 107–149. posits: Attributes and models of some major deposits in India, Australia,
Large, R.R., Bull, S.W., McGoldrick, P.J., Walters, S., Derrick, G.M., and and Canada: New Delhi, India, Narosa Publishing House, p. 149–190.
Carr, G.R., 2005, Stratiform and strata-bound Zn-Pb-Ag deposits in Pro- —— 2004b, Geology of the Belt-Purcell basin and the Sullivan deposit, in
terozoic sedimentary basins, northern Australia: ECONOMIC GEOLOGY 100TH Deb, M., and Goodfellow, W.D., eds., Sediment-hosted lead-zinc sulfide
ANNIVERSARY VOLUME, p. 000. deposits: Attributes and models of some major deposits in India, Australia,
Larter, R.C.L., Boyce, A.J., and Russell, M.J., 1981, Hydrothermal pyrite and Canada: Narosa Publishing House, p. 100–148.
chimneys from the Ballynoe barite deposit, Silvermines, Co. Tipperary, Ire- Lyle, J.R., 1977, Petrography and carbonate diagenesis of the Bonneterre
land: Mineralium Deposita, v. 16, p. 309–318. Formation in the Viburnum Trend area, Southeast Missouri: ECONOMIC
Lavery, N.G., Leach, D.L., and Saunders, J.A., 1994, Lithogeochemical in- GEOLOGY, v. 72, p. 420–434.
vestigations applied to exploration for sediment-hosted lead-zinc deposits: Ma, G., Beaudoin, G., Qi, S., and Li, Y., 2004, Geology and geochemistry of
Society for Geology Applied to Mineral Deposits Special Publication 10, p. the Changba SEDEX Pb-Zn deposit, Qinling orogenic belt, China: Miner-
393–428. alium Deposita, v. 39, p. 380–395.
Leach, D.L., 1980, Nature of mineralizing fluids in the barite deposits of cen- Marshall, B., and Spry, P.G., 2000, Discriminating between regional meta-
tral and southeast Missouri: ECONOMIC GEOLOGY, v. 75, p. 1168–1180. morphic remobilization and syntectonic emplacement in the genesis of
——1994 Genesis of the Ozark Mississippi Valley-type metallogenic massive sulfide ores: Reviews in Economic Geology, v. 11, p. 39–79.
province, Missouri, Arkansas, Kansas, and Oklahoma, USA, 1994: Society Mavrogenes, J.A., Hagni, R.D. and Dingess, P.R., 1992, Mineralogy, parage-
for Geology Applied To Mineral Deposits Special Publication 10, p. nesis, and mineral zoning of the West Fork mine, Viburnum Trend, South-
104–138. east Missouri: ECONOMIC GEOLOGY, v. 87, p. 113–124.
Leach, D.L., and Rowan, E.L., 1986, Genetic link between Ouachita foldbelt McArdle, P., 1990, A review of carbonate-hosted base metal-barite deposits
tectonics and the Mississippi Valley-type lead-zinc deposits of the Ozarks: in the Lower Carboniferous of Ireland: Chronique de la Miniere
Geology, v. 14, p. 931–935. Recherche, v. 500, p. 3–29.
Leach, D.L., and Sangster, D.F., 1993, Mississippi Valley-type lead-zinc de- McGoldrick, P.J., Bull, S.W., Cooke, D.R., and Large, R.R., 1998, Northern
posits: Geological Association of Canada Special Paper 40, p. 289–314. Australian Proterozoic stratiform sediment-hosted (SEDEX) Zn-Pb-Ag de-
Leach, D.L., Viets, J.G., Kozlowski, A., and Kibitlewski, S., 1996, Geology, posits: University of Tasmania, Centre for Ore Deposit and Exploration
geochemistry, and genesis of the Silesia-Cracow zinc-lead district, south- Studies (CODES) Special Publication 2, p. 11–22.
ern Poland: Society of Economic Geologists Special Publication 4, p. McKnight, E.T., 1935, Zinc and lead deposits of northern Arkansas: U.S. Ge-
144–170. ological Survey Bulletin 853.
Leach, D.L., Bradley, D., Lewchuk, M.T., Symons, D.T.A., de Marsily, G., McLimans, R.K., Barnes, H.L., and Ohmoto, H., 1980, Sphalerite stratigra-
and Brannon, J., 2001, Mississippi Valley-type lead-zinc deposits through phy of the Upper Mississippi Valley zinc-lead district: ECONOMIC GEOLOGY,
geological time: Implications from recent age-dating research: Mineralium v. 75, p. 351–361.
Deposita, v. 36, p. 711–740. Meinert, L., Dipple, G., Chang, L., and Megaw, P., 2005, World skarn de-
Leach, D.L., Bradley, D., Lewchuk, M., Symons, D.T.A., Premo, W., Bran- posits: ECONOMIC GEOLOGY 100TH ANNIVERSARY VOLUME, p. 000.
non, J., and Marsily, G., 2002, Reply to Discussion on “Mississippi Valley- Meyer, C., 1981, Ore-forming processes in geologic history: ECONOMIC GE-
type lead-zinc deposits through geological time: Implications from recent OLOGY 75TH ANNIVERSARY VOLUME, p. 6–41.
age-dating research” by S.E. Kesler and C.W. Carrigan (2002): Mineralium ——1988, Ore deposits as guides to geologic history of the earth: Annual Re-
Deposita, v. 37, p. 803–805. view of Earth and Planetary Science, v. 16, p. 147–171.

0361-0128/98/000/000-00 $6.00 604


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 605

Mitchell A.H.G., 1985, Mineral deposits related to tectonic events accompa- Plumlee, G.S., Leach, D.L., Hofstra, A.H., Landis, G.P., Rowan, E.L., and
nying arc-continent collision: Transactions of the Institution of Mining and Viets, J.G., 1994, Chemical reaction path modeling of ore deposition in
Metallurgy, sec. B, v. 94, p. B115–B125. Mississippi Valley-type Pb-Zn deposits of the Ozark region, U.S. midconti-
Moldovanyi, E.P., and Walter, L.M., 1992, Regional trends in water chem- nent: ECONOMIC GEOLOGY, v. 89, p.1361–1383.
istry, Smackover Formation, southwest Arkansas: Geochemical and physi- Qing, H., and Mountjoy, E.W., 1992, Large-scale fluid flow in the Middle
cal controls: American Association of Petroleum Geologists Bulletin, v. 76, Devonian Presqu’ile barrier, Western Canada sedimentary basin: Geology,
p. 864–894. v. 20, p. 903–906.
Montañez, I., 1996, Application of cathodoluminescent cement stratigraphy Raffensperger, J.P., 1996, Numerical simulation of sedimentary basin-scale
for delineating regional diagenetic and fluid migration events associated hydrochemical processes, in Corapciolglu, M.Y., ed., Advances in porous
with Mississippi Valley-type mineralization in the southern Appalachians: media 3: New York, Elsevier, p. 185–305.
Society of Economic Geologists Special Publication 4, p. 432–447. Ravenhurst, C.E., Reynolds, P.H., Zentilli, M., and Akande, S.O., 1987, Iso-
Moore, D.W., Young, L.E., Modene, J.S., and Plahuta, J.T., 1986, Geologic topic constraints on the genesis of Zn-Pb mineralization at Gays River,
setting and genesis of the Red Dog zinc-lead-silver deposit, western Brooks Nova Scotia, Canada: ECONOMIC GEOLOGY, v. 82, p. 1294–1308.
Range, Alaska: ECONOMIC GEOLOGY, v. 81, p. 1696–1727. Reed, C.P., and Wallace, M.W., 2001, Diagenetic evidence for an epigenetic
Morelli, R.M., Creaser, R.A., Selby, D., Kelley, K.D., Leach, D.L., and King, origin of the Courtbrown Zn-Pb deposit, Ireland: Mineralium Deposita, v.
A.R., 2004, Re-Os sulfide geochronology of the Red Dog sediment-hosted 36, p. 428–441.
Zn-Pb-Ag deposit, Brooks Range, Alaska, Alaska: ECONOMIC GEOLOGY, v. ——2004, Zn-Pb mineralisation in the Silvermines district, Ireland: A prod-
99, p. 1569–1576. uct of burial diagenesis: Mineralium Deposita, v. 39, p. 87–102.
Muir, M.D., 1983, Depositional environments of host rocks to northern Aus- Rhodes, D., Lantos, E.A., Lantos, J.A., Webb, R.J., and Owen, D.C., 1984,
tralian lead-zinc deposits, with special reference to McArthur River: Min- Pine Point orebodies and their relationship to the stratigraphy, structure,
eralogical Association of Canada Short Course Handbook, v. 8, p. 141–169. dolomitization, and karstification of the Middle Devonian barrier complex:
Mullane, M.M., and Kinnard, J.A., 1998, Synsedimentary mineralization at ECONOMIC GEOLOGY, v. 79, p. 991–1055.
Ballynoe barite deposit, near Silvermines, Co. Tipperary, Ireland: Transac- Rimstidt, J.D., 1997, Gangue mineral transport and deposition, in H.L.
tions of the Institution of Mining and Metallurgy, sec. B, v. 107, p. 48–61. Barnes, ed., Geochemistry of hydrothermal ore deposits: New York, John
Newhouse, W. H., 1933, Temperature of formation of the Mississippi Valley Wiley and Sons, p. 487–516.
lead-zinc deposits: ECONOMIC GEOLOGY, v. 28, p. 744–750. Rittenhouse, G., 1967, Bromine in oil-field waters and its use in determining
Nisbet, E.G., and Fowler, C.M.R., 1983, Model for Archean plate tectonics: possibilities of origin of these waters: American Association of Petroleum
Geology, v. 11, p. 376–379. Geologists Bulletin, v. 51, p. 2430–2440.
Ohle, E.L., 1959, Some considerations in determining the origin of ore de- Roberts, S.J., Nunn, J.A., Cathles, L.M., and Cipriani, F-D., 1996, Expulsion
posits of the Mississippi Valley-type. Part 1: ECONOMIC GEOLOGY, v. 54, p. of abnormally pressured fluids along faults: Journal of Geophysical Re-
769–789. search, v. 101, p. 28231–28252.
——1970, Mississippi Valley-type ore deposits, a general review: Washington Robinson, M., and Godwin, C.I., 1995, Genesis of the Blende carbonate-
Division of Mines Bulletin, v. 61, p. 5–15. hosted Zn-Pb-Ag deposit, North-Central Yukon Territory: Geologic, fluid
——1980, Some considerations in determining the origin of ore deposits of inclusion and isotopic constraints: ECONOMIC GEOLOGY, v. 90, p. 369–384.
the Mississippi Valley-type. Part II: ECONOMIC GEOLOGY, v. 75, p. 161–172. Roedder, E., Ingram, B., and Hall, W.E., 1963, Studies of fluid inclusions.
——1985, Breccias in Mississippi Valley-type deposits: ECONOMIC GEOLOGY, III. Extraction and quantitative analysis of inclusions in the milligram
v. 80, p. 1736–1752. range: ECONOMIC GEOLOGY, v. 58, p. 353–374.
Ohmoto, H., and Rye, R.O., 1979, Isotopes of sulfur and carbon, in Barnes, Rouvier, H., Perthuisot, V., and Mansouri, A., 1985, Pb-Zn deposits and salt-
H.L. ed., Geochemistry of hydrothermal ore deposits: New York, Wiley-In- bearing diapirs in southern Europe and North Africa: ECONOMIC GEOLOGY,
terscience, p. 509–567. v. 80, p. 666–687.
Oliver, J., 1986, Fluids expelled tectonically from orogenic belts: Their role Rowan, E.L., 1986, Cathodoluminescent zonation in hydrothermal dolomite
in hydrocarbon migration and other geologic phenomena: Geology, v. 14, p. cements: Relationship to Mississippi Valley-type Pb-Zn mineralization in
99–102. southern Missouri and northern Arkansas, in Hagni, R.D., ed., Process
——1992, The spots and stains of plate tectonics: Earth Science Reviews, v. mineralogy VI: Warrendale, Pennsylvania, Metallurgical Society, p. 69–87.
32, p. 77–106. Rowan, E.L., and Goldhaber, M.B., 1995, Duration of mineralization and
Orgeval, J.J., 1994, Peridiapiric metal concentration: Example of the Bou fluid-flow history of the Upper Mississippi Valley zinc-lead district: Geol-
Grine deposit (Tunisian Atlas): Society for Geology Applied to Mineral De- ogy, v. 23, p. 609–612.
posits Special Publication 10, p. 354–389. Rowan, E.L., and Marsily, G., 2001, Infiltration of Late Paleozoic evaporative
Orr, W.L., 1974, Changes in sulfur content and isotopic ratios of sulfur dur- brines in the Reelfoot rift: A possible source for Illinois basin formation wa-
ing petroleum maturation—study of Big Horn basin Paleozoic oils, Part 1.: ters and MVT mineralizing fluids: Petroleum Geoscience, v. 7, p. 269–279.
American Association of Petroleum Geologists Bulletin, v. 58, p. Roy, A.B., Kumar, S., Laul, V., and Chuahan, N.K., 2004, Tectonostratigraphy
2295–2318. of the lead-zinc-bearing metasedimentary rocks of the Rampura-Agucha
—— 1982, Rate and mechanism of non-microbial sulfate reduction [abs.]: mine and its neighbourhood, District Bhilwara, Rajasthan: Implications for
Geological Society of America Abstracts with Programs, v. 14, p. 580. metallogeny., in Deb, M., and Goodfellow, W.D., eds., Sediment hosted
Parr, J.M., and Plimer, I.R., 1993, Models for Broken Hill-type lead-zinc-sil- lead-zinc sulphide deposits: Attributes and models of some major deposits
ver deposits: Geological Association of Canada Special Paper, v. 40, p. in India, Australia and Canada: New Delhi, India, Narosa Publishing
253–288. House, p. 273–289.
Peace, W.M, 1999, Carbonate-hosted Zn-Pb mineralization within the Upper Rozendaal, A., 1980, The Gamsberg zinc deposit, South Africa: A banded
Pale Beds at Navan, Ireland: Unpublished Ph.D. thesis, University of Mel- stratiform base-metal sulfide ore deposit, in Ridge, J.D., ed., Proceedings
bourne, 248 p. of the Quadrennial IAGOD Symposium, 5: Stuttgart, E. Schweizer-
Peace, W.M., and Wallace, M. W., 2000, Timing of mineralization at the Zn- bart’sche Verlagsbuchhandlung (Naegele u. Obermiller), p. 619–633.
Pb deposit; A post-Arundian age for Irish mineralization: Geology, v. 28, p. Rozendaal, A., and Stumpfl, E.F., 1984, Mineral chemistry and genesis of
711–714. Gamsberg zinc deposit, South Africa: Transactions of the Institution of
Peace, W.M., Wallace, M.W., Holdstock, M.P., and Ashton, J.H., 2003, Ore Mining and Metallurgy, sec. B, v. 93, p. B161–B175.
textures within the U lens of the Navan Zn-Pb deposit, Ireland: Mineral- Russell, M.J., 1978, Downward-excavating hydrothermal cells and Irish-type
ium Deposita, v. 38, p. 568–584. ore deposits: Importance of an underlying thick Caledonian prism: Trans-
Person, M.A., and Garven, G., 1994, A sensitivity study of the driving forces actions of the Institution of Mining and Metallurgy, sec. B, v. 87, B168–171.
on fluid flow during continental rift evolution: Geological Society of Amer- Russell, M.J., and Skauli, H., 1991, A history of theoretical developments in
ica Bulletin, v. 106, p. 461–475. carbonate-hosted base metal deposits and new tri-level enthalpy classifica-
Plumb, K.A., Derrick, G.M., and Wilson, I.H., 1980, Precambrian geology of tion: ECONOMIC GEOLOGY MONOGRAPH 8, p. 96–116.
the McArthur River-Mount Isa region, northern Australia, in Henderson, Sangster, D.F., 1988, Breccia-hosted lead-zinc deposits in carbonate rocks, in
R.A., and Stephenson, P.J., eds. The geology and geophysics of northeast- James, N.P., and Choquette, P.W., eds., Paleokarst: Heidelberg, Germany,
ern Australia: Queensland, Geological Society of Australia, p. 71–88. Springer Verlag, p. 106–116.

0361-0128/98/000/000-00 $6.00 605


606 LEACH ET AL.

——1990, Mississippi Valley-type and sedex lead-zinc deposits: A compara- Slack, J.F., Kelley, K.D., Anderson, V.M., Clark, J.L., and Ayuso, R.A., 2004b,
tive examination: Transactions of the Institution of Mining and Metallurgy, Multistage hydrothermal silicification and Fe-Tl-As-Sb-Ge-REE enrich-
sec. B, v. 99, p. B21–B42. ment in the Red Dog Zn-Pb-Ag district, northern Alaska: Geochemistry,
——1996, Cabonate-hosted lead-zinc deposits: Society of Economic Geolo- origin, and exploration applications: ECONOMIC GEOLOGY, v. 99, p.
gists Special Publication 4, 664 p. 1481–1508.
—— 2002, The role of dense brines in the formation of vent-distal sedimen- Snyder, F.G., 1967, Genesis of strata-bound lead-zinc-barite-fluorite deposits
tary-exhalative (SEDEX) lead-zinc deposits: field and laboratory evidence: in carbonate rocks: ECONOMIC GEOLOGY MONOGRAPH 3, p. 1–13.
Mineralium Deposita, v. 37, p. 149–157. Snyder, F.G. and Gerdemann, P.E., 1968, Geology of the southeast Missouri
Sangster, D.F., and Carriere, J.J., 1991, Preliminary studies of fluid inclusions lead district, in Ridge, J.D., ed., Ore deposits of the United States,
in sphalerite from the Robb Lake Mississippi valley-type deposit, British 1933–1967, Graton-Sales Volume: New York, American Institute of Min-
Columbia: Geological Survey of Canada Paper 91–1E, p. 25–32. ing, Metallurgical, and Petroleum Engineers, p. 327–358.
Sangster, D.F., and Hillary, E. M., 1998, SEDEX lead-zinc deposits—pro- Solomon, M., and Heinrich, C.A., 1992, Are high-heat producing granites es-
posed subtypes and their characteristics: Exploration and Mining Geology, sential to the origin of giant lead-zinc deposits at Mount Isa and McArthur
v. 7, p 341–357. River, Australia?: Exploration and Mining Geology, v. 1, p. 85–91.
Sass-Guskiewicz, M., and Dzulynski, S., 1998, On the origin of strata-bound Solomon, M., Groves, D.I., and Jaques, A.J., 2000, The geology and origin of
ores in the Upper Silesia, Poland: Annales Societatis Geologorum Poloniae, Australia’s mineral deposits: Centre for Ore Deposit Research, University
v. 68, p. 267–278. of Tasmania, and Centre for Global Metallogeny, University of Western
Schaefer, M.O., 2002, Paleoproterozoic Mississippi valley-type Pb-Zn de- Australia [Perth], 957 p.
posits of the Ghaap Group, Transvaal Supergroup in Griqualand West, Stalder, M., and Rozendaal, A., 2004, Apatite nodules as an indicator of de-
South Africa: Unpublished Ph.D. thesis, Johannesburg, South Africa, Rand positional environment and ore genesis for the Mesoproterozoic Broken
Afrikaans Unviersity, 373 p. Hill-type Gamsberg Zn- Pb deposit, Namaqua province, South Africa: Min-
Selby, D., and Creaser, R.A., 2003. Direct dating of bitumen through the ap- eralium Deposita, v. 39, p. 189–203.
plication of the 187Re-187Os isotope system: First results from the Polaris Stanley, C.J., Walters, S.G., Jeffrey, S., Lawie, D., 2001, Primary lithogeo-
Mississippi Valley-type Zn-Pb deposit, Nunavut, Canada [abs.]: Geological chemical halos associated with the Cannington Broken Hill-type Ag-Pb-Zn
Association of Canada-Mineralogical Association of Canada Annual Meet- sediment hosted massive sulphide deposit, Mt Isa inlier, Queensland: Im-
ing, Vancouver, 2003, Abstracts 76, CD archive.\ plications for genesis and mineral exploration [abs.]: International Geo-
Sevastopulo, G.D., and Redmond, P., 1999, Age of mineralization of carbon- chemical Exploration Symposium, 20th, Santiago, Chile, 2001, Abstracts
ate-hosted base metal deposits in the Rathdowney Trend, Ireland, in Mc- with Programs, p. 158–162.
Caffrey, K.W., Lonergan, L., and Wilkinson, J.J., eds.: Fractures, fluid flow Stormo, S., and Sverjensky, D.A., 1983, Silicate hydrothermal alteration in a
and mineralization, Geological Society, London, Special Publications, v. Mississippi Valley-type deposit-Viburnum Trend, southeast Missouri [abs.]:
151, p. 303–311. Geological Society of America Abstracts with Programs, v. 15, p. 699.
Shah, N., 2004, Rampura-Agucha, a remobilized SEDEX deposit southeast- Sverjensky, D.A., 1981, The origin of a Mississippi Valley-type deposit in the
ern Rajasthan, India, in Deb, M., and Goodfellow, W.D., ed., Sediment Viburnum Trend, southeast Missouri: ECONOMIC GEOLOGY, v. 76, p.
hosted lead-zinc sulphide deposits; Attributes and models of some major 1848–1872.
deposits in India, Australia and Canada: New Delhi, India, Narosa Pub- ——1984, Oil field brines as ore-forming solutions: ECONOMIC GEOLOGY, v.
lishing House, p. 290–303. 79, p. 23–37.
Shanks, W.C., III, Woodruff, L.G., Jilson, G.A., Jennings, D.S., Modene, J.S., ——1986, Genesis of Mississippi Valley-type lead-zinc deposits: Annual Re-
and Ryan, B.D., 1987, Sulfur and lead isotope studies of stratiform Zn-Pb- view of Earth and Planetary Sciences, v. 14, p. 177–199.
Ag deposits, Anvil Range, Yukon: basinal brine exhalation and anoxic bot- Swenson, J., Person, M., Raffensperger, J.P., Cannon, W., Woodruff, L., and
tom-water mixing: ECONOMIC GEOLOGY, v. 82, p. 600–634. Berndt, M.E., 2004, A hydrogeologic model of stratiform mopper mineal-
Sharp, J.M., Jr., and Domenico, P.A., 1976, Energy transport in thick se- ization in the Midcontinent rift system, northern Michigan, USA: Geoflu-
quences of compacting sediment: Geological Society of America Bulletin, ids, v. 4, p. 1–22.
v. 87, p. 390–400. Symons, D.T.A., 2002, Paleomagnetic dating of Navan and East Tennesssee
Shearley, E. Redmond, P., King, M., and Goodman, R., 1996, Geological con- ores: Don’t believe the geologists!: Dublin, Irish Association for Economic
trols on mineralization and dolomitization of the Lisheen Zn-Pb-Ag de- Geology Lecture, Irish Geological Survey.
posit, Co. Tipperary, Ireland, in Strogen, P., Somerville, I.D., and Jones, G. Symons, D.T.A., Smethurst, M.T., Ashton, J.H., 2001, Paleomagnetic dating
Ll., eds., Recent advances in Lower Carboniferous geology: Geological So- of the Navan Zn-Pb deposit, Ireland [abs]: Geological Society of America
ciety, London, Special Publications, v. 107, p. 23–33. Abstracts with Programs, v. 33, no. 6, 338 p.
Sherlock, R.L., Lee, J.K.W., and Cousens, B.L., 2004, Geologic and ——2002, Paleomagnetism of the Navan Zn-Pb deposit, Ireland: ECONOMIC
geochronologic constraints on the timing of mineralization at the Nanisivik GEOLOGY, v. 97, p. 997–1012.
zinc-lead Mississippi Valley-type deposit, Northern Baffin Island, Nunavut, Taylor, B.E., and Beaudoin, G., 2000, Sulphur isotope stratigraphy of the Sul-
Canada: ECONOMIC GEOLOGY, v. 99, p. 279–293. livan Pb-Zn-Ag deposit, B.C.: Evidence for hydrothermal sulphur, and bac-
Siebenthal, C.E., 1916, Origin of the zinc and lead deposits of the Joplin re- terial and thermochemical sulphate reduction: Geological Association of
gion, Missouri, Kansas, and Oklahoma: U.S. Geological Survey Bulletin Canada, Mineral Deposits Division Special Volume 1, p. 696–719.
606, 283 p. Taylor, B.E., Turner, R.J.W., Leitch, C.H.B., Watanabe, D.H., and Lydon,
Simms, M.A., and Garven, G., 2004, Thermal convection in faulted exten- J.W., 2000, Isotopic stratigraphy of carbonate in wall rocks and bedded
sional sedimentary basins: Theoretical results from finite-element model- ores, Sullivan Pb-Zn-Ag mine, British Columbia: Geological Association of
ing: Geofluids, v. 4, p. 109–130. Canada, Mineral Deposits Division Special Volume 1, p. 673–695.
Sinclair W.D., Chorlton L.B., Laramee R.M., Eckstrand O.R., Kirkham Thorkelson, D.J., Mortensen, J.K., Davidson, G.J., Creaser, R.A., Perez,
R.V., Dunne K.P.E., and Good D.J., 1999, World minerals geoscience W.A., and Abbott, J.G., 2001, Early Mesoproterozoic intrusive breccias in
databaseproject: Digital databases of generalized world geology and min- Yukon, Canada: the role of hydrothermal systems in reconstructions of
eral deposits for mineral exploration and research, in Stanley C.L., et al, North America and Australia: Precambrian Research, v. 111, p. 31–55.
eds., Mineral deposits: Processes to Processing: Rotterdam, Balkema, p. Titley, S.R., 1993, Relationship of strata-bound ores with tectonic cycles
1435–1437. of the Phanerozoic and Proterozoic: Precambrian Research, v. 61, p.
Slack, J.F., Shaw, D.R., Leitch, C.H.B., and Turner, R.J.W., 2000, Tourma- 295–322.
linites and coticules from the Sullivan Pb-Zn-Ag deposit and vicinity, Trude, K.J., and Wilkinson, J.J., 2001, A mineralogical and fluid inclusion
British Columbia: Geology, geochemistry, and genesis: Geological Associa- study of the Harberton Bridge Fe-Zn-Pb deposit, County Kildare, Ireland:
tion of Canada, Mineral Deposits Division Special Publication 1, p. Journal of the Geological Society, London, v. 158, p. 37–46.
736–767. Turner, R.J.W., 1990, Jason stratiform Zn-Pb barite deposit, Selwyn basin,
Slack, J.F., Dumoulin, J.A., Schmidt, J.M., Young, L.E., and Rombach, C.S., Canada: Geological setting, hydrothermal facies and genesis, in Abbott,
2004a, Paleozoic sedimentary rocks in the Red Dog Zn-Pb-Ag district and J.G., and Turner, R.J.W., eds., Mineral deposits of the northern Canadian
vicinity, western Brooks Range, Alaska: Provenance, deposition, and metal- Cordillera: International Association on the Genesis of Ore Deposits, Field
logenic significance: ECONOMIC GEOLOGY, v. 99, p. 1385–1414. Trip 1, Guidebook, p. 137–175.

0361-0128/98/000/000-00 $6.00 606


SEDIMENT-HOSTED Pb-Zn DEPOSITS: A GLOBAL PERSPECTIVE 607

——1992, Formation of Phanerozoic stratiform sediment-hosted zinc-lead Werdon, M.B., 1996, Drenchwater, Alaska: Zn-Pb-Ag mineralization in a
deposits: Evidence for the critical role of ocean anoxia: Chemical Geology, mixed black shale-volcanic environment: Geology and Ore Deposits of the
v. 99, p. 165–188. American Cordillera, Geological Society of Nevada Symposium,
Turner, R.J.W., Leitch, C.H.B., Höy, T., Ransom, P.W., Hagen, A., and De- Reno/Sparks, Nevada, April 1995, Proceedings, p. 1341–1354.
laney, G.D., 2000, Sullivan graben system: District-scale setting of the Sul- Wheatley, C.J.V., Whitfield, G.G., Kenny, K.J., and Birch, A., 1986, The Per-
livan deposit: Geological Association of Canada, Mineral Deposits Divi- ing carbonate-hosted zinc-lead deposit, Griqualand West, in Anhaeusser,
sionSpecial Publication 1, p. 408–439. C.R., Maske, S., eds., Mineral deposits of southern Africa: Johannesburg,
Velasco, F., Herrero, J.M., Yusta, I., Alonso, J.A., Seebold, I., and Leach, Geological Society of South Africa, p. 867–874.
D.L., 2003, Geology and geochemistry of the Reocin zinc-lead deposit, White, D.E., 1958, Liquid inclusions in sulfides from Tri-State (Missouri,
Basque-Cantabrian basin, northern Spain: ECONOMIC GEOLOGY, v. 98, p. Kansas, and Oklahoma) are probably connate in origin [abs.]: Geological
1371–1396. Society of America Bulletin, v. 69, p. 1660–1661.
Viets, J.G., 1983, Geochemical variations of major, minor, and trace elements ——1968, Environments of generation of some base metal ore deposits:
in samples of the Bonneterre Formation from drill holes transecting the ECONOMIC GEOLOGY, v. 63, p. 301–335.
Viburnum Trend Pb-Zn district of Southeast Missouri, in Kisvarsanyi, G., Wilkinson, J.J., 2003, On diagenesis, dolomitisation and mineralization in the
Grant, S.K., Pratt, W.P., and Koenig, J.W., eds., International Conference Irish Zn-Pb orefield: Mineralium Deposita, v. 38, p. 968–983.
on Mississippi Valley-type Lead-Zinc Deposits. Proceedings volume: Uni- Wilkinson, J.J., and Earls, G., 2000, A high-temperature hydrothermal origin
versity of Missouri-Rolla, Rolla Press, p. 174–186. for black dolomite matrix breccias in the Irish Zn-Pb orefield: Mineralogi-
Viets, J.G., Hofstra, A.H., Emsbo, P., and Kozlowski, A., 1996, The composi- cal Magazine, v. 64, p. 1017–1036.
tion of fluid inclusions in ore and gangue minerals from Mississippi Valley- Wilkinson, J.J., and Lee, M.J., 2003, Cementation, hydrothermal alteration,
type Zn-Pb deposits of the Cracow-Silesia region of southern Poland: Ge- and Zn-Pb mineralization of carbonate breccias in the Irish Midlands: Tex-
netic and environmental implications, in Gorecka, E., and Leach, D.L., tural evidence from the Cooleen zone, near Silvermines, County Tipper-
eds., Carbonate-hosted zinc-lead deposits in the Silesian-Cracow area, ary—a reply: ECONOMIC GEOLOGY, v. 98, p. 194–198.
Poland: Warsaw, Poland, Prace Panstwowego Instytuti Geologicznego, v. Wilkinson, J.J., Eyre, S.L., and Boyce, A.J., 2005, Ore-forming processes in
154, p. 85–104. Irish-type carbonate-hosted Zn-Pb deposits: Evidence from mineralogy,
Voss, R.L., and Hagni, R.D., 1985, The application of cathodoluminescence chemistry, and isotopic composition of sulfides at the Lisheen mine: ECO-
microscopy to the study of sparry dolomite from the Viburnum Trend, NOMIC GEOLOGY, v. 100, p. 63–86.
Southeast Missouri, in Hausen, D.M., and Kopp, O.C. eds., Mineralogy; Williams, N., 1978, Studies of the base metal sulfide deposits at McArthur
applications to the minerals industry: Proceedings of the Paul F. Kerr River, Northern Territory, Australia. I. The Cooley and Ridge deposits:
Memorial Symposium, p. 51–68. ECONOMIC GEOLOGY, v. 73, p. 1005–1035.
Wachowiak, N.M., Gummer, P.K., Pope, M., and Grotzinger, J.P., 1997, Willis, I.L., 1996, Exploration for Broken Hill-type Pb-Zn-Ag deposits: Uni-
Stratabound Zn-Pb mineralization to the Esker horizon, Paleoproterozoic versity of Tasmania, Centre for Ore Deposit and Exploration Studies
Rocknest Formation, Wopmay orogen, Canada [abs.]: NWT Geoscience (CODES) Special Publication 1, p. 145–152.
Forum 25th Anniversary, Department of Indian and Northern Affairs, Yel- de Wit, M.J., 1998, On Archean granites, greenstones, cratons and tectonics:
lowknife, Program and Abstracts, p. 96–98. Does the evidence demand a verdict?: Precambrian Research, v. 91, v.
Walcher, E.H., 1986, Geologic-mineral deposit investigations of the Ram- 181–227.
melsburg deposit: Unpubliahed Ph.D. dissertation, Clausthal-Zellerfeld, Yang, J., Bull, S., and Large, R., 2004, Numerical investigation of salinity in
Federal Republic of Germany (DEU) Technical University Clausthal. controlling ore-forming fluid transport in sedimentary basins: Example of
Walters, S.G., 1998, Broken Hill-type deposits: AGSO Journal of Australian the HYC deposit, northern Australia: Mineralium Deposita, v. 39, p.
Geology and Geophysics, v. 17, p. 229–237. 622–631.
——2001, New applications of resistate indicator minerals for base metal ex- Young, L.E., 2004, A geologic framework for mineralization in the western
ploration: An example from Broken Hill-type Pb-Zn-Ag deposits: Brooks Range, Alaska: ECONOMIC GEOLOGY, v. 99, p. 1281–1306.
Townsville, Queensland, Economic Geology Research Unit Contribution Zartman, R.E., 1984, The lower crust; missing reservoir in the lead isotope
59, p. 209–210. paradox: Transactions, American Geophysical Union, v. 65, 230 p.
Walters, S.G., and Bailey, A., 1998, Geology and mineralization of the Can-
nington Ag-Pb-Zn deposit: An example of Broken Hill-type mineralization
in the Eastern Succession, Mount Isa inlier, Australia: ECONOMIC GEOL-
OGY, v. 93, p. 1307–1329.
Walters, S.G., Skrzeczynski, B., Whiting, T., Bunting, F., and Arnold, G.,
2002, Discovery and geology of the Cannington Ag-Pb-Zn deposit, Mount
Isa Eastern Succession, Australia: Development and application of an ex-
ploration model for Broken Hill-type deposits: Society of Economic Geol-
ogists Special Publication 9, p. 95–118.

0361-0128/98/000/000-00 $6.00 607

View publication stats

You might also like