Вы находитесь на странице: 1из 23

2005 Society of Economic Geologists, Inc.

Economic Geology, v. 100, pp. 13251347

The Environment of Vein Formation and Ore Deposition in the


Purisima-Colon Vein System, Pachuca Real del Monte District, Hidalgo, Mexico*
JOHN E. DREIER
13790 Braun Road, Golden, Colorado 80401

Abstract
The Pachuca Real del Monte district is a leading example of a low total sulfide epithermal vein-type silver
district. In its 500-year history, Pachuca has produced about 40 106 kg of silver from more than 100 million
metric tons (Mt) of ore at an average total sulfide content of about 1 wt percent. Hosting the more than 100
veins of the district are calc-alkalic series volcanic rocks of Oligocene-mid Miocene age that total about 2,700
m in thickness. K-Ar dating of the volcanic rocks and the veins indicates that the veins formed within ~1 m.y.
of the termination of volcanism at about 21 Ma.
The present study is of the north-southtrending Purisima-Colon vein system of the Pachuca Real del Monte
district where field, petrographic, and fluid inclusion studies provide a three-dimensional and time transgressive
view of vein formation and important clues to the genesis of the general class of low total sulfide silver veins. The
Purisima-Colon vein deposits are located in a high-angle normal fault system where vertical offset varies sys-
tematically from ~30 m in the center to nil at the north and south ends. The fault system is blind and its top de-
scribes a concave downward arc, the highest point coinciding with the area of maximum offset and the lowest
points, at the distal north and south ends, with the least offset. The top of the fault transitions upward into a zone
of distributed fracturing that extends upward an additional 300 to 350 m. Illite-pyrite-calcite-chlorite alteration
pervades the zone of distributed fracturing above the fault, whereas propylitic alteration dominates wall rocks
below the top of the fault. Ore-grade silver mineralization is within a 650-m vertical interval that extends down-
ward from the top of the vein to a bottom determined by silver grade. The vein system is vertically zoned, from
the top down, with respect to thickness, sulfide content, and silver grade as follows: (1) an upper zone of 50-m
vertical extent in which the vein is <0.5 m wide and has low total sulfides (<<1 wt %) and low Ag grade (<100
g/t), (2) an intermediate zone of about 100-m vertical extent that averages 0.5 to 1 m wide and contains low to
moderate sulfides (1 wt %) and low to moderate Ag values (150300 g/t), (3) a main ore zone of 350-m verti-
cal extent in which the vein is 2 to 5 m wide and contains moderate to high sulfides (averages 12 wt %), and
high to moderate Ag values (5002,000 g/t), and (4) a lower zone of unknown vertical extent (extending >500 m
below the mine workings) that is similar to the main ore zone but contains only base metal sulfides.
The veins developed in four stages: (1) brecciation of the host rock, (2) a silicate stage, (3) an ore stage, and
(4) a postore stage. Stage 1, brecciation of the host rock, took place prior to the onset of hydrothermal activity
when broken wall rocks accumulated in openings produced by separation of the fault walls. Stage 2, the silicate
stage, is dominated by various forms of fine-grained to microcrystalline quartz and chalcedony plus coarse-
grained crystalline quartz, johansennite, K-feldspar, albite, clinozoisite, epidote, and hematite. Petrographic and
fluid inclusion data indicate that much of the fine-grained to microcrystalline quartz and chalcedony of stage 2
was precipitated rapidly from <180C fluids that were supersaturated to highly supersaturated with silica,
whereas the coarse-grained crystalline quartz and associated johansennite, K-feldspar, and clinozoisite precipi-
tated from fluids at near-quartz saturation and at temperatures of 230 to 310C. The coarse crystalline quartz
was deposited by fluids that ascended the veins along the boiling curve, as evidenced by all-vapor fluid inclusion
assemblages and by fluid homogenization temperatures plotted versus depth; PCO2 for the silicate-stage fluids
was <102 bar and PO2 was in the hematite field. Fluid inclusion ice-melting temperatures from quartz, adularia,
epidote, and johansennite average ~1 wt percent NaCl equiv (range = 04 wt % NaCl equiv). The silicate-stage
mineral assemblage, quartz + adularia + clinozoisite, in conjunction with fluid homogenization and freezing data,
indicates that the silicate stage was deposited from low-salinity, oxidizing, neutral pH hydrothermal fluids simi-
lar to the equilibrated, deeply circulated meteoric fluids of modern dilute, neutral pH geothermal systems.
The ore stage was deposited at 270 to 300C by fluids that were acid, of moderate salinity, and which did
not deposit silica or quartz except at the very beginning of the ore stage or as very local alteration products of
johansennite. Sulfides were precipitated mainly by fluid-rock reactions in which wall rocks and stage 2 alumino-
silicate minerals were altered or totally replaced by sulfides + illite + kaolinite + carbonate + Mg-chlorite. Ore-
stage fluids, with salinity in the range of 5 to 7 wt percent NaCl equiv, resemble the magmatic fluids of
Giggenbach (1997) diluted by one to two parts of equilibrated meteoric fluids. Whereas most sulfides precip-
itated from water-rock reactions, sulfides lining vugs probably precipitated due to cooling and/or dilution.
The separate quartz and sulfide deposition at Pachuca is believed to be a common feature of silver-domi-
nant epithermal veins based on descriptions of other mineral deposits of this type. This observation suggests
that the absence of quartz at Pachuca may be a defining feature of this class of deposits and might provide im-
portant clues to the ultimate source of the metals. The separate genesis of quartz and ore-stage minerals also
requires caution when applying the results of fluid inclusion studies in quartz to problems of ore transport and
deposition in such systems.
E-mail, jedreier@att.net

*The Appendix is available electronically at <www.segweb.org/EG/papers/Abs100-7_files/DreierAppendix.pdf>

0361-0128/05/3546/1325-23 $6.00 1325


1326 JOHN E. DREIER

Introduction normal faults and are generally coextensive with wall-rock


breccias, although several important veins were actually
THE PACHUCA REAL DEL MONTE district (Pachuca), located lodes. Principal vein minerals are quartz, chalcedony, jo-
in Hidalgo, Mexico, about 90 km north of Mexico City (Fig. hansennite, K-feldspar, base metal sulfides, and argentite.
1), is one of the worlds great mining camps. Renowned for its Overall, the total sulfide content of the veins is about 1 per-
great bonanzas and very large silver production, it remains ac- cent by weight and recoverable metals up to about 1950 av-
tive today nearly 500 years after discovery. Based on mineral- eraged 350g/t Ag, 0.5 percent Zn, 0.05 percent Pb, and 0.01
ogy, sulfide content, and environment of formation, Pachuca percent Cu (McGrath, 1950; Geyne et al., 1963).
is classified as a low total sulfide epithermal Ag vein district The great size and richness of the bonanzas at Pachuca
(also referred to as low sulfidation, adularia-sericite, or low compelled early and prolonged study, beginning with Burkart
sulfur epithermal by Bonham, 1986; Heald et al., 1987; and (1836) and including important contributions by Aguilera et
White and Hedenquist 1990; or intermediate sulfidation by al. (1897), Ordonez and Rangel (1899), Ordonez (1902),
Hedenquist et al., 2000). Pachucas metal production of 40.4 Wisser (1937, 1941a, b, 1942, 1946, 1948, 1951, 1964),
106 kg (1.3 109 oz) Ag, 187,000 kg (6 106 oz) Au and 1 Thornburg (1945, 1952), Bastin (1948), and Geyne et al.
106 t of Zn + Pb + Cu (Geyne et al., 1963, 1990; Real del (1963). The Geyne et al. (1963) study, conducted from 1949
Monte Company, unpub. data) is the largest among low total to 1963 as a joint investigation by the Compania de Real del
sulfide epithermal Ag vein deposits. Monte y Pachuca and the U.S. Geological Survey, was the cul-
The district contains hundreds of productive veins spread mination of an intensive geologic investigation of the district
out over an area of more than 70 km2, but most of the Ag pro- begun in about 1920. It is a detailed report of the stratigraphy,
duction and all of the bonanzas were in three compact sub- structure, and mineralization of an area exceeding 120 km2
districts: (1) western Vizcaina-Santa Ana-Maravillas, (2) Real and a more generalized geologic evaluation of the surround-
del Monte, and (3) Santa Gertrudis (Fig. 2). Considering the ing 4,000 km2. The present study is part of an ongoing effort
size of the district and the large number of veins, the mineral (Dreier, 1976, 1982, 1984; McKee et al., 1992) to extend the
deposits are remarkably consistent in form, mineralogy, work of Geyne et al. (1963) by investigating fluid inclusions,
grade, structure, paragenesis, and vertical extent. In keeping vein mineral relationships, wall-rock alteration, geochronol-
with this general uniformity, results of detailed studies of vein ogy, and structural aspects of vein control at Pachuca and
mineralogy, paragenesis, and structure by Bastin (1948), other low total sulfide epithermal Ag veins.
Wisser (1948), and Geyne et al. (1963) showed that district- This paper reports the results of a study of the mineralogy,
wide vein formation was simultaneous and that sulfide min- structure, fluid inclusions, and wall-rock alteration of the
eralization was a single event. The veins fill open zones in Purisima-Colon vein system of the Real del Monte subdistrict

FIG. 1. Regional geologic map of the Pachuca district modified from Geyne et al. (1963). Locations of veins outside the
district are from Wisser (1964) and unpublished sources. Trans-Mexico volcanic province is modified from Suter et al. (2001)
and Henry and Aranda-Gomez (1992).

0361-0128/98/000/000-00 $6.00 1326


PURISIMA-COLON VEIN SYSTEM, PACHUCA REAL DEL MONTE DISTRICT, MEXICO 1327

FIG. 2. Vein map of the Pachuca district modified from Geyne et al. (1963). Veins: C = Colon, D = Dios Te Guie, M =
Maravillas, P = Purisima, SG = Santa Gertrudis, SI = Santa Ines, T = Tapona, V = Vizcaina. Highlighted areas, 1 (western
Vizcaina-Santa Ana-Maravillas), 2 (Real del Monte), and 3 (Santa Gertrudis), show the most intensely mineralized parts of
the district and the loci of the great bonanzas. Heavy lines = veins, light lines = postmineral faults. The datum planes for the
western and eastern halves of the district are taken at the elevations of maximum data and the greatest extent of mine work-
ings for each half.

(Figs. 2, 3). With an estimated Ag production of 2.4 to 3.3 M Finlandia vein and fluid inclusion studies at Fresnillo, Za-
kg (77100 Moz) and a ~1 Mt bonanza grading >1.5 kg/t Ag, catecas, and elsewhere in Mexico (Simmons et al., 1988; Al-
the Purisima-Colon vein system was among the greatest dis- binson, 1995; Albinson et al., 2001), indicate a complex his-
coveries of the district. The system was also undisturbed by tory of vein formation involving multiple compositionally
supergene oxidation and leaching, it was reasonably accessi- distinct hydrothermal fluids of up to 14 wt percent NaCl
ble in mine workings (Fig. 3D), it was well tested by drilling, equiv. The principal mechanisms invoked for Ag deposition
and the rocks above it were widely exposed in outcrop. were boiling, dilution, and cooling (Heald et al., 1987; White
Over the past 30 years or so, many published studies have and Hedenquist, 1990), generally in that order. One of the
focused on the general class of epithermal veins and the principal features of epithermal precious metals vein deposits
smaller class of low total sulfide epithermal Ag veins. Heald is their restriction to a relatively narrow vertical interval. In
et al. (1987) summarized the results of fluid inclusion melting the case of low total sulfide epithermal Ag deposits, the verti-
point depression studies carried out on low total sulfide ep- cal interval of Ag mineralization is generally <700 m (Heald
ithermal Ag veins by ONeil and Silberman (1974), Dreier et al., 1987) and the top of the Ag horizon is variously ascribed
(1976), Buchannan (1979), Vikre (1980), and Fahley (1981) to (1) the presence of an overlying, mechanically incompetent
and deduced that low total sulfide Ag veins were formed from stratigraphic unit (Ferguson, 1921; Steven and Rattee, 1965),
neutral pH, low-salinity (<3 wt % NaCl equiv) hydrothermal (2) the development of a hydrothermally generated clay cap
solutions. This conclusion was generally supported by White above a boiling zone (Heald et al., 1987), or (3) boiling in an
and Hedenquist (1990) and by the study of Clarke and Titley open channel (Buchannan, 1979, 1980).
(1988). However, the results of fluid inclusion studies at The objectives of this study were to determine if (1) the
Creede, Colorado (Barton et al., 1977), a comparative analy- Pachuca veins were formed from a neutral pH, low-salinity
sis of epithermal Ag and Au systems in Heald et al. (1987), fluid or by the interplay of several fluids, (2) the Ag was de-
including Creede, Julcani, Lake City, Goldfield, and the posited by boiling, dilution, cooling or other processes, and

0361-0128/98/000/000-00 $6.00 1327


1328 JOHN E. DREIER

FIG. 3. Longitudinal projections of the Purisima-Colon vein system. A. Geology and mined areas and vein map of the
450-m level. Geology was projected from geologic maps and cross sections of Geyne et al. (1963). The stope outlines are from
the mine database and the vein map of the 450-m level is from mine level maps. B. Vein widths and vertical offsets on the
fault system. Vein widths were compiled from mine level maps and assay longitudinal projections. Vertical offsets, in meters,
on the fault system were compiled from maps and cross sections of Geyne et al. (1963) and show the vertical offset on the
fault at the approximate elevation of the 450-m level at the coordinate of each line. C. Ag grade distribution compiled from
mine assay longitudinal projections, diamond drill logs, and mine level maps. D. Locations of samples used in the study and
isotherms derived from maximum fluid inclusion homogenization temperatures of stage 2b quartz at the numbered sample
sites.

0361-0128/98/000/000-00 $6.00 1328


PURISIMA-COLON VEIN SYSTEM, PACHUCA REAL DEL MONTE DISTRICT, MEXICO 1329

(3) the upper surface of the Ag horizon was controlled by faults occur only in the Real del Monte area (Fig. 2) and were
stratigraphy, a boiling-induced clay cap, or something else. active for a relatively brief period of time just prior to and
Neutral pH geothermal systems have been proposed as close during mineralization.
modern analogues to low total sulfide epithermal veins
(Ewers and Keys, 1977; White, 1981; Henley and Ellis, 1983; Description of the Purisima-Colon Vein System
Krupp and Seward, 1987, 1990; Spycher and Reed, 1989; The Purisima-Colon vein system fills a complex north-
White and Hedenquist, 1990). One of the principal objectives southtrending, 75 to 85 east-dipping normal fault system
of this study was to compare and contrast the geochemistry of extending from about 8,500 N to ~13,500 N (Figs. 2, 3A).
low-salinity, neutral pH geothermal systems and the Vertical offset along the system ranges from a maximum of
Purisima-Colon vein forming fluids. about 30 m near 11,000 N to nil at both the north and south
The study is based on field and underground mapping, ends (Fig. 3B). The vein system contained three main ore-
sampling of the Purisima-Colon vein system (Fig. 3D) and its bodies: (1) the Colon orebody, which extended from about
surrounding and overlying wall rocks, and logging and sam- 8,800 N to 10,100 N and which continued to the northeast
pling of several thousand meters of drill core. Vein mineral- along the La Rica vein for another 200 m past the orebody of
ogy was studied in more than 300 thin sections by standard Figure 3A-C; (2) the main Purisima orebody, which extended
optical petrographic methods augmented by X-ray diffraction from 10,350 N to 11,350 N; (3) the north Purisima orebody,
and electron microprobe work, extending the work of Dreier which extended from 11,600 N to 12,450 N. With a strike
(1976, 1982). Vein mineral relationships and textures also length of 1,000 m and a vertical extent of about 650 m, the
were studied in many mine workings and in several hundred main Purisima orebody was the largest deposit of the three.
hand samples. Separating the main Purisima and Colon orebodies is a struc-
turally complex region of narrow, low-grade to barren
Geology stringers and unmineralized faults. Assay maps for different
The Pachuca district occurs within a ~700 km2 Oligocene mine levels in this region show Ag values mostly below 50 g/t,
to early Miocene calc-alkaline volcanic field located along the in many places below 20 g/t, and in still more, the prospects
northeastern edge of the Trans-Mexico volcanic province for commercial mineralization were judged so poor that sam-
(Delgado-Granados et al., 2000; Fig. 1). Rocks within the ples were not taken. Between the main and north Purisima
Pachuca field range in composition from basaltic andesite to orebodies, the Purisima fault is a narrow gouge zone that does
rhyolite with andesite and dacite flows and flow breccias as not contain vein material.
the principal components. A maximum thickness for the field
is about 2,400 m. The volcanic pile is divisible into a lower Mineralization
succession of mostly intermediate composition flows and Commercial Ag mineralization, as defined by a ~200 g/t Ag
breccias and subsidiary volcaniclastic sedimentary rocks of ~1-m grade-thickness cutoff, occurs within a 650-m verti-
generally uniform thickness comprising 75 percent of the cal interval between 2,100 and 2,750 m above mean sea level
field, and an upper succession of dacite breccias and rhyolite (Fig. 3A-C). Estimates from assay plans indicate an average
flows and tuffs of highly variable thickness making up the re- Ag grade of 350 to 400 g/t, or about the district average,
mainder. The two successions are separated from each other whereas scattered assays and observations at sample sites
by an angular unconformity that locally cuts up to 1,000 m (Fig. 3D) suggest that base metals grades were perhaps twice
from the lower succession (Geyne et al., 1963). Four sets of the district average. Silver grades are strongly zoned in the
east-westtrending dikes are cogenetic with the upper suc- Purisima-Colon system (Figs. 3C, 4). The highest Ag grades
cession; the youngest dikes are feeders to volcanic domes. occur in the mid to mid-lower levels of the central Purisima
Whereas the lower succession was deposited under condi- vein where they define a bonanza zone about 900 m long and
tions of relative tectonic quiescence, the upper succession 200 m high, averaging above 1.5 kg/t Ag. Away from the bo-
records a history of episodic volcanism, erosion, warping, in- nanza zone, Ag values progressively diminish in all directions;
trusion, and faulting. Geologic relationships indicate that the lowest Ag grade zones are located at the top, bottom, and
mineralization occurred at or near the end of the deposition distal ends of the system. Base metal values are also zoned,
of the upper succession (Wisser, 1951; Geyne et al., 1963). but the absence of a global base metals database prohibits the
The close timing of volcanism, faulting, and mineralization construction of base metal isocons. Scattered assay data and
deduced from field observations is supported by K-Ar ages of qualitative observations reveal local areas of multipercent
about 21.6 Ma for the youngest volcanic and intrusive rocks combined base metals throughout the workings and suggest
of the upper succession (Geyne et al., 1990) and 20.3 Ma for that values are generally <1 percent in the upper mine levels
vein adularia from the Maravillas vein (McKee et al., 1992). and ~2 to 4 percent below about 2,300 m (Aguilera et al.,
Rocks in the Pachuca volcanic field are cut by east- 1897; Ordonez and Rangel, 1899; Turban, 1947; Wisser, 1948,
westtrending normal faults with dips of 55 to 70 and verti- 1951; Geyne, 1949). Comparison of Figure 3A and B indi-
cal offsets of up to 500 m and north-south normal faults with cates a positive relationship between vein width and Ag
dips >75 and maximum vertical offsets of only 30 m (Fig. 2). grade.
The east-west faults are present throughout the district and Deep drilling and an isopach diagram demonstrate that
its surroundings, began their growth during the deposition of the Purisima-Colon vein system persists to depth where silver
the lower succession, and parallel the structural grain of the grades decrease (Fig. 3B, C) and base metal grades increase.
Trans Mexico volcanic province (Henry and Aranda-Gomez, As the Purisima-Colon vein system and its base metals com-
1992; Suter et al., 2001). By contrast, the north-south normal ponent extend far below the bottom of the mine workings, the

0361-0128/98/000/000-00 $6.00 1329


1330 JOHN E. DREIER

discontinuous quartz-calcite-pyrite veinlets (Fig. 4). Silver


grades in the fracture zone are 1 ppm and base metal con-
centrations are 200 ppm. The distributed fracturing and
quartz-calcite veining above the Purisima-Colon system are
typical of other studied structures in the district (Aguilera et
al., 1897; Ordonez and Rangel, 1899; Thornburg, 1945;
Wisser, 1951; Geyne et al., 1963; Geyne, 1968; Dreier, 1976,
1982). As a generalization, the width and intensity of fractur-
ing are greatest within 100 to 200 m above the vein and there-
after diminish with elevation up to 300 to 400 m above the top
of the vein system where they die out (Fig. 4). The absence of
fracturing and veining at high levels above the vein system is
confirmed by a surface traverse along the projected course of
the Colon vein at about 9,900 N where the top of the vein is
between 300 and 450 m below the surface (Fig. 3A). In that
area, the nearly continuous outcrops along the projected trace
of the vein consist of purple Vizcaina Formation andesite de-
void of faults, veins, and the oxidized remnants of sulfides.
The top of the Ag horizon forms a concave downward sur-
face that is unaffected by stratigraphic contacts, an unconfor-
mity, dikes, and faults. The highest point on the Ag horizon,
at about 11,000 N, coincides with the area of maximum offset
on the fault; at the lowest points, at the distal north and south
ends, the offset approaches nil (Fig. 3A-B).
Wall-rock alteration
Extending from the top of the Ag horizon down to the
deepest explored levels of the district, the wall rocks are
widely altered to the assemblage albite + K-feldspar + epi-
dote + chlorite + calcite + pyrite prehnite (Wisser, 1951;
Geyne et al., 1963; Dreier, 1982), with the most intense al-
teration occurring in fracture zones or permeable tuffs. Start-
ing about 100 to 150 m below the top of the Ag horizon along
FIG.. 4. Schematic cross section through the Purisima vein, showing av- the Purisima-Colon vein system, propylitic alteration adjacent
erage vein widths and the distribution of ore and alteration minerals. The di- to the individual veins is overprinted by a 1- to 2-m selvage of
agram is based on mine assay longitudinal projections, level maps, core log- weak illite + pyrite + chlorite alteration and quartz-calcite
ging, geologic mapping data in Geyne (1948) and Geyne et al. (1963), and a
description of vertical zoning of the adjacent north-south Santa Ines vein by veining that increases upward in width and intensity. Pro-
Wisser (1951). ceeding upward, the illite selvage expands to 10 m and then,
near the top of the veins, to more than 50 m. As the zone ex-
pands in width, it changes from a light dusting of illite at the
Ag horizon evidently represents the top of a large, zoned min- bottom to an illite-pyrite (pyrite ~510 vol %)-chlorite rock
eral system dominated by Ag in the upper part and base met- just above the top of the vein (Fig. 4). Above the top of the Ag
als in the lower part. The Purisima-Colon system is similar to horizon, the zone of illite alteration coincides with the zone of
the adjacent Santa Ines vein where Wisser (1951) described distributed fracturing and quartz-calcite veining.
four vertical zones as follows: (1) an upper zone of 100- to
150-m vertical extent in which the vein is 0.5 m wide and con- Depth of formation of the silver horizon
tains very sparse sulfides; (2) an intermediate zone of 50- to For the Pachuca district as a whole, the highest points of
100-m vertical extent in which the vein widens to 2 m and the silver horizon are located within the Vizcaina Formation,
contains moderately abundant sulfides; (3) a main ore zone of the lowest formation of the upper succession. Based on strati-
~300-m vertical extent in which the vein is 5 to 8 m wide and graphic reconstructions of the formations overlying the Viz-
contains abundant sulfides; and (4) a lower zone of unknown caina Formation at various locations, Geyne et al. (1963) esti-
vertical extent (extending below workings and drill intercepts) mated that the highest points of the top of the Ag horizon
that is visually identical to zone 3 but contains only base metal were originally located at a depth of about 300 m. They noted,
sulfides. however, that their estimate was compromised by the highly
The top of the Ag horizon (Fig. 3A-C) coincides approxi- variable thickness of the upper succession and the effects of
mately with the pinchout point of the vein (Fig. 4), although, faulting, tilting, and erosion. Based on maps and cross sec-
locally, the vein extends 25 m or so above the top of the Ag tions of Geyne et al. (1963), the highest point of the Ag hori-
horizon. Based on mine mapping and core logging, the top of zon on the Purisima-Colon vein system (11,000 N) is within
the vein system occurs where the enclosing fault transforms an area that is more deeply eroded than its surroundings; in
upward into a zone of distributed fracturing characterized by the area of 11,000 N, the Vizcaina Formation is about 300 m

0361-0128/98/000/000-00 $6.00 1330


PURISIMA-COLON VEIN SYSTEM, PACHUCA REAL DEL MONTE DISTRICT, MEXICO 1331

thick and the top of the Ag horizon is about 100 m above the matrix material, and multiple sets of stockworks and subpar-
bottom of the Vizcaina Formation. Measurements of strati- allel veins. Study of these elements in mine workings and thin
graphic sections on cross sections of Geyne et al. (1963) show sections reveals an evolving paragenesis (Figs. 5, 6; Aguilera
that in surrounding areas where the full original thickness of et al., 1897; Ordonez and Rangel, 1899; Thornburg, 1945;
the Vizcaina Formation is preserved, it is between 500 and Wisser, 1948; Geyne et al., 1963). Based on crosscutting rela-
600 m thick and is locally overlain by erosional remnants, 20 tionships, the paragenesis is divisible into four stages: (1) rock
to 200 m in thickness, of rhyolite and volcaniclastic sedimen- breccia, (2) silicate, (3) ore, and (4) postore (Figs. 5, 6A). In
tary rocks of the upper succession; at the time of mineraliza- order of abundance by volume, the veins are composed of
tion, the upper rhyolites were probably thicker than they are rock fragments (4060%), quartz-chalcedony (3040%), jo-
today. If, at the time of mineralization, the highest point of hansennite (515%), K-feldspar + albite (510%), and sul-
the Ag horizon along the Purisima-Colon vein system was fides (12%). The principal ore mineral is argentite but spha-
covered by the original thickness of the Vizcaina Formation, lerite, galena, and chalcopyrite are also economically
minus 100 m (the top of the vein is located ~100 m above the important.
bottom of the Vizcaina Formation) plus the full original thick-
ness of the overlying rhyolite, then the thickness of cover Stage 1: Wall-rock breccia
above the top of the Ag horizon at the time of mineralization A principal component of the veins is a breccia composed
would have been between 400 and 700 m, depending on local of angular fragments of the adjacent wall rocks (andesite frag-
conditions of erosion and deposition, with the most likely ments dominate where the vein walls are in andesite, whereas
depth range of 550 to 650 m. rhyolite fragments are present where the veins follow or
crosscut rhyolite dikes) set in a matrix of fine, angular rock
Description of the Purisima-Colon Veins material and its alteration products (Fig. 6A-C). Breccia
Veins of the Purisima-Colon system are highly complex in fragments range in diameter from 0.3 m down to a few mi-
structure, consisting of angular rock fragments, fine-grained crons, with ~ 80 percent of the material >5 mm and ~10 to 15

FIG. 5. Paragenesis of the Purisima-Colon vein system. The earliest vein component (stage 1) consists of rock fragments.
In stage 2a, the rock-flour matrix of the breccia was dissolved or altered to fine-grained quartz + K-feldspar. In Stage 2b, the
deposition of plumose and milky quartz and chalcedony (deposition temperature = 200C) alternated with clear quartz +
adularia + hematite + johansennite or epidote (deposition temperature = 250320C). Clinozoisite and adularia were de-
posited at, or near the end of, stage 2b. The alteration minerals K-feldspar, Fe chlorite, epidote, and albite accompanied stage
2. Stage 2c, gray quartz, was deposited before and during the early stages of sulfide mineralization. Stage 3 consisted of sul-
fides accompanied by the alteration minerals illite, calcite, CaMg(CO3)2, Mg chlorite, and kaolinite. Quartz is rare in stage 3
except as a local alteration product of johansennite. Stage 4 consists of amethyst, crystalline quartz, and calcite. Faulting took
place before, during, and after vein formation.

0361-0128/98/000/000-00 $6.00 1331


1332 JOHN E. DREIER

FIG. 6. Photomicrographs and photographs of rock breccia, minerals, and mineral relationships of stage 2. A. Wall-rock
breccia fragments surrounded by stage 2a and cut by stage 2b, ore- and postore-stage veins. B. Wall-rock fragments sur-
rounded by stage 2a quartz. C. Fine wall-rock fragment altered to K-feldspar and quartz; fragment margin is intergrown with
stage 2a. Black areas are sulfides replacing K-feldspar in the fragment margin. D. Quartz intergrown with johansennite and
K-feldspar; note chalcedony layer. Dark spots are sulfides replacing K-feldspar and johansennite. E. Stage 2b banded vein
composed of four cycles; light bands are clear quartz, gray bands are milky quartz, thin bands are chalcedony. The center of
the vein (at the top of the photograph) is johansennite. F. Chalcedony band (dark) in milky quartz. Abbreviations: Ch = chal-
cedony; J = johansennite; K = K-feldspar; Q = quartz; R = rock fragment; Z = zoned quartz; 2a = stage 2a; 2b = stage 2b; S
= sulfides; 4 = stage 4.

percent <0.1 mm (rock flour). No complex fragments or frag- substages a through c, based on crosscutting relationships
ments containing truncated veins were seen in this study or (Fig. 6A).
by Geyne et al. (1963). On this basis, and because the frag- Stage 2a consists of (1) very fine grained (<0.010.03 mm)
ments are crosscut by all other vein stages, it is interpreted quartz-K-feldspar intergrowths that surround rock fragments;
that the wall-rock breccias belong to the first phase of vein (2) slightly coarser grained patches of similar material albite,
formation. Fe chlorite, and epidote that appear to represent the altered
remnants of fine rock fragments; and (3) fine-grained quartz-
Stage 2: Silicate stage K-feldspar-albite rinds around or incomplete replacements of
The silicate stage consists, in decreasing abundance, of rock fragments (Fig. 6A, C). Stage 2a never occurs as planar
quartz, johansennite, K-feldspar, clinozoisite, epidote, albite, veins (Fig. 6A, C). Most of the stage 2a material is honey-
and hematite (Figs. 6, 7); minor to trace amounts of prehnite, combed by irregular openings that are lined or filled by later
diopside, wollastonite, and tremolite were reported in other vein phases and cut by planar veins of all later stages (Fig.
veins by Geyne et al. (1963). The silicate stage is divisible into 6C).

0361-0128/98/000/000-00 $6.00 1332


PURISIMA-COLON VEIN SYSTEM, PACHUCA REAL DEL MONTE DISTRICT, MEXICO 1333

FIG. 7. Photomicrographs (crossed polars) of stage 2b silicate minerals. A. Cloudy quartz. B. Plumose quartz. C. Zoned
quartz intergrown with johansennite (dark fibrous mineral). D. Hematite intergrown with epidote and Fe chlorite in stage
2b quartz. E. Clinozoisite intergrown with adularia in a stage 2b quartz vein. F. Stage 2c quartz (gray quartz) cutting spha-
lerite and pyrite. Abbreviations: Cl = chlorite; Cli = clinozoisite; E = epidote; H = hematite; K = K-feldspar; Py = pyrite; Q
= quartz; S = sphalerite.

Stage 2b minerals occur as vug fillings in the irregular johansennite, and the several quartz forms (Fig. 6E). The
openings of stage 2a material and as linear veins that cut rock vein shown in Figure 6E contains four cycles of vein fill, each
fragments and stage 2a (Fig. 6A, C). The principal minerals of cycle consisting of clear and cloudy mosaic bands, but some
stage 2b are quartz, chalcedony, K-feldspar, johansennite, and cycles also contain chalcedony and mosaic quartz layers.
epidote; hematite is widespread but much less abundant Quartz stratigraphy is also present within individual quartz
(Figs. 6C-F, 7A-E). Milky and plumose quartz (Sander and crystals, termed zoned quartz, in which clear quartz bands
Black, 1988) are the most abundant quartz forms of stage 2b alternate with cloudy, mosaic quartz, or chalcedony layers
(Figs. 6D-F, 7A-B) but many grains dominated by these (Fig. 7C). The youngest growth band of most zoned quartz
forms have zones of clear quartz and chalcedony (Fig. 6D-E). grains is a crystalline variety that is commonly intergrown
Chalcedony bands, visible as linear extinction zones under with, or overgrown by, euhedral adularia and clinozoisite (Fig.
crossed polars (Fig. 6F), are common within masses or larger 7E).
veins of plumose + milky quartz. Whereas many stage 2b Due to the abundance and importance of johansennite as
veins lack well-defined internal organization, some have a a site for later ore-stage mineralization, this mineral deserves
crude banding parallel to the walls defined by chalcedony, special discussion. Johansennite was identified in this study

0361-0128/98/000/000-00 $6.00 1333


1334 JOHN E. DREIER

by X-ray diffraction, microprobe, and optical methods but Within the Ag horizon, the ore sulfides comprise fine
was misnamed by past workers (Bastin, 1948; Geyne et al., monomineralic grains or irregular aggregations of different
1963) as rhodonite-bustamite. Johansennite occurs as irregu- minerals; less commonly, they are present as well-developed
lar masses of randomly oriented, fine (<<0.1 mm), acicular 5- to 10-mm crystals lining vugs. In contrast, below the Ag
grains that surround andesite-dacite rock fragments, and as horizon, base metal sulfides tend to be present as larger
multiple layers within stage 2b where it is intergrown with grains and are concentrated in massive sulfide veins. Sulfide
clear quartz (Figs. 6D-E, 7C). The presence of johansennite minerals occur in a variety of settings as follows: (1) replace-
as encrustations on rock fragments, intergrowths with fine- ments of rock fragments and silicate-stage minerals; (2) dif-
grained quartz and K-feldspar, and as layers within stage 2b fuse bands around johansennite masses and rock fragments;
quartz indicates episodic deposition throughout stage 2b. (3) vug and microfracture linings and pore fillings between
Epidote occurs in similar settings to johansennite but is much quartz grains or groups of grains; (4) sulfide-only veins; and
less abundant. It is frequently intergrown with hematite (Fig. (5) replacements of earlier sulfides (Fig. 8A-H). According to
7D), which is also scattered throughout stage 2b veins as iso- Geyne et al. (1963), most of the ore sulfides are disseminated
lated grains and fine layers. in quartz, but Ordonez (1902), Bastin (1948), Thornburg
Stage 2b contains linear breccias, up to 1 mm wide, of fine- (1945), Wisser (1948), and this study found that most of the
grained, angular fragments of vein material surrounded by sulfides replace silicate-stage minerals and rock fragments.
very fine grained, anhedral quartz made up of multiple ex- The replacement nature of the sulfides is evident at all scales.
tinction zones. Based on the linear form of the breccias and At the district scale, the Santa Gertrudis (Geyne et al., 1963)
the presence within them of fragments of vein material, the and Tapona orebodies (Thornburg, 1945) were lodes defined
zones are interpreted as microfault planes. The very fine by assay limits and consisted of sheeted masses of intimately
grain size and the presence of multiple extinction zones in the mixed clay/chlorite-altered wall rock and ore sulfides with lit-
matrix to the breccias suggest that this material was deposited tle quartz vein material. At the hand specimen and outcrop
as chalcedony. scale in the Purisima-Colon veins and other described veins
Stage 2c consists of anhedral, clear, inclusion-free quartz throughout the district, sulfides, together with clays and car-
that occurs in gray vitreous veins. Under the microscope, bonates, coat or replace wall-rock fragments, form hairline
stage 2c quartz is variably intergrown with and cuts pyrite and veinlets within fragments, and replace silicate-stage minerals,
some sphalerite (Fig. 7F) but is cut by much sphalerite and particularly johansennite and epidote (Ordonez, 1902; Geyne
all galena, chalcopyrite, and argentite. These relationships in- et al., 1963). In thin section, sulfides commonly form pseudo-
dicate a temporal progression from early stage 2c quartz, to morphs after silicates or contain remnant inclusions of re-
one or more events of pyrite-sphalerite deposition inter- placed silicates, or they replace silicate-stage minerals (gen-
spersed with minor stage 2c quartz, to sphalerite, to galena- erally K-feldspar and clinozoisite) that line vugs or
chalcopyrite-argentite. microfractures (Fig. 8A-E). The replacement of parageneti-
Stage 2 was accompanied by the alteration minerals K- cally earlier sulfides by later sulfides was also an important
feldspar, albite, Fe chlorite, epidote, and quartz but these depositional mechanism, and it is common to find sphalerite
minerals vary in absolute and relative amounts depending on replacing earlier pyritohedral pyrite and galena, chalcopyrite,
the composition of the host rock, the proximity to a stage 2 and argentite replacing sphalerite (Fig. 8F-H). Most spha-
vein, and the density of stage 2 veins. Fine rock fragments, lerite is affected by chalcopyrite disease (Barton and Bethke,
rock fragments cut by many stage 2 veins, and tuffaceous 1987) where chalcopyrite blebs and stringers are present
fragments are altered to quartz-K-feldspar mosaics (Fig. 6C). along fractures or cleavage planes or form highly irregular
Fragments ranging from 0.5 to 1 mm in diameter or those cut bands associated with adjacent or intruding pyrite, chalcopy-
by fewer stage 2 veins retain some original texture but are rite, galena, and argentite grains (Fig. 8F). According to the
mostly converted to quartz, K-feldspar, albite, chlorite, and discussion in Bortnikov et al. (1991), the chalcopyrite disease
epidote. Fragments >5 mm and rocks >5 to 10 mm from at Pachuca is the result of the replacement of sphalerite by
stage 2 veins are characterized by propylitic alteration in chalcopyrite as opposed to coprecipitation.
which the rock matrix is only partly altered to K-feldspar, Argentite, the principal ore sulfide, occurs (1) by itself and
fine-grained quartz, and albite, and the plagioclase phe- with galena and chalcopyrite in the centers of vugs lined by
nocrysts partly converted to epidote, albite, and K-feldspar. zoned quartz and other stage 2b silicate minerals (where the
Beyond 10 to 20 cm of a stage 2 vein, the intensity of propy- sulfides replace K-feldspar and clinozoisite), (2) replacing sil-
litic alteration diminishes until plagioclase and the matrix are icate-stage minerals (particularly johansennite) and wall-rock
fresh. fragments, and (3) as veins cutting sphalerite (Fig. 8A-B, G-
H). Argentite is rare in samples from the lower levels of the
Stage 3: Ore stage mine (2,300-m level and below). Based on crosscutting rela-
The ore stage consists of pyrite, base metal sulfides, ar- tionships, sulfide paragenesis is in the following order: (1)
gentite, native silver, with alteration consisting of calcite, Ca- pyrite, (2) sphalerite, (3) galena, (4) chalcopyrite, (5) argen-
Mn carbonate, illite, Mg-Fe chlorite, minor silica, and kaolin- tite, and (6) native silver (Fig. 5; Bastin, 1948; Geyne et al.,
ite (kaolinite may be a product of postmining oxidation). 1963). Mill head assays confirm that the earliest sulfide min-
Mg-Fe chlorite also occurs as a vein filling. Although the av- erals are the most abundant (sphalerite > galena > chalcopy-
erage sulfide content of the Purisima-Colon veins within the rite > argentite). Where sulfides are in contact with quartz,
Ag horizon is about 2 wt percent, locally it is up to 20 wt per- they fill open spaces between quartz crystals (Fig. 8E), cor-
cent with combined Zn + Pb + Cu of up to 10 wt percent. rode the quartz, or cut the quartz. The absence of ore-stage

0361-0128/98/000/000-00 $6.00 1334


PURISIMA-COLON VEIN SYSTEM, PACHUCA REAL DEL MONTE DISTRICT, MEXICO 1335

FIG. 8. Photomicrographs of ore-stage minerals. A. Rock fragment replaced by argentite. B. Johansennite largely altered
to carbonate and clay and replaced by sulfides. C. Adularia altered to illite and replaced by sulfides. D. Enlargement of
showing illite. E. Sulfides in center of stage 2b quartz vein, triangular etch pits are in galena. F. Vein of galena, chalcopyrite,
and argentite cutting sphalerite; note chalcopyrite disease in sphalerite. G. Galena, and argentite cutting sphalerite. H.
Galena, chalcopyrite, and argentite cutting quartz. Abbreviations: A = argentite; CC = clay + CaMn carbonate; CPY = chal-
copyrite; G = galena; I = illite; K = adularia; R = rock; S = sphalerite; SLF = sulfides; 2b = stage 2b.

quartz is notable, except at the very beginning of the sulfide Alteration minerals associated with the ore stage include
stage, where pyrite and sphalerite are associated with gray two dominant assemblages: (1) illite + chlorite (talc-penni-
quartz, and within or immediately adjacent to replaced jo- nite) + kaolinite? +/- calcite and (2) Ca-Mn carbonate (50%
hansennite, where fine-grained quartz with remnant jo- Ca -50% Mn) + calcite + chlorite + clay +/- fine-grained
hansennite inclusions occurs with Ca-Mn carbonate (Fig. 8B). quartz. The illite-chlorite-kaolinite assemblage occurs as

0361-0128/98/000/000-00 $6.00 1335


1336 JOHN E. DREIER

selvages on sulfide veins that cut wall-rock fragments and im- they were aligned on fractures or trails that cut across growth
parts a light green-white bleached appearance to the rock. zones in the host mineral or several minerals. The morphol-
Where the density of sulfide veins is high or where dissemi- ogy of the fluid inclusions was studied and inclusions were
nated sulfide mineralization is intense, large bodies of rock classified into fluid inclusion assemblages, using the format of
are altered to the illite-chlorite-kaolinite assemblage; in the Bodnar et al. (1985) and Goldstein and Reynolds (1994). De-
Santa Gertrudis and Tapona lode deposits, illite-chlorite- termination of homogenization temperatures was limited to
kaolinite alteration occupies >106 m3 of rock. Locally along assemblages with consistent filling ratios and results were ac-
the Purisima-Colon system, sulfidation is intense, and in cepted only if temperature determinations for an assemblage
these locations rock fragments are converted to knots of sul- were within ~15C; for assemblages where temperature vari-
fides and clay or to cavities lined by sulfides and clay. The Ca- ations >15C it was assumed that necking down of individual
Mn carbonate + calcite + chlorite + clay fine-grained quartz fluid inclusions had taken place. For assemblages with con-
assemblage occurs within johansennite masses where it is sistent liquid/vapor ratios, fluid constituted 75 to 85 percent
generally much more intensely developed and pervasive than of the filling material and vapor, the remainder, although a
is the illite-chlorite-kaolinite assemblage within rock frag- few inclusions contained tiny opaque solid inclusions that
ments. The johansennite masses shown in Figure 8B have were insoluble on heating. Some quartz grains hosted assem-
been converted entirely to a mixture of carbonate, clay, and blages that contained both liquid- and vapor-rich inclusions
sulfides. Alteration-mineralization zoning in the johansennite suggestive of the trapping of separate liquid and vapor
masses comprises (1) an inner zone of isolated johansennite phases. A determined search was made in melting point runs
patches that are cut by irregular veinlets of alteration miner- for clathrates, but none were found, indicating that XCO2 was
als minor sulfides; (2) an intermediate zone of carbonates, below about 4 percent (PCO2 <7080 bars: Barton and Chou,
clay, and chlorite with local, irregular sulfide veins or patches; 1993; Hedenquist and Henley, 1985). Measurements of melt-
and (3) an outer rind of fine-grained quartz + fine-grained ing point temperature were converted to NaCl equiv salinity
sulfides + sulfide veinlets + isolated carbonate/clay patches using data from Bodnar (1993).
clay + chlorite. This alteration imparts a brown or greenish According to Bodnar et al. (1985), Goldstein and Reynolds
coloration to the johansennite masses on fresh surfaces and a (1994), and T. J. Reynolds (pers. commun., 2000) fluid inclu-
soft, muddy tan and/or black appearance in mine workings. sion assemblages may be divided into four types based on
morphology and relative consistency of liquid to vapor ratios
Stage 4: Postore as follows: irregular inconsistent (ii), irregular consistent (ic),
Postore minerals include amethyst, clear, coarse-grained, regular consistent (rc), and negative-crystal types (Fig. 9A-E).
crystalline quartz, and white calcite. These minerals are pre- This classification system, which is based on empirical obser-
sent as veins and coatings on ore-stage and earlier minerals. vations from many types of hydrothermal systems, although
Volumetrically, the postore minerals are unimportant. not tested under controlled laboratory conditions, provides
useful information about conditions of inclusion formation
Fluid Inclusion Study and whether leakage or necking occurred subsequent to de-
The fluid inclusion study is based on samples collected at position. According to Bodnar et al. (1985) and T. J. Reynolds
sites shown in Figure 3D. The sample sites were chosen to (pers. commun., 2000) type ii inclusions typically form below
represent the complete vertical and horizontal extent of the about 200C, type ic fluid inclusions form between 220 and
vein system, although some preselected sample sites were 230C, type rc fluid inclusions form above 250C, and nega-
abandoned due to problems of accessibility. Each site was ei- tive crystal fluid inclusions form above 260C. In this study,
ther photographed or recorded as a hand drawing, and sam- type ii fluid inclusions were not suitable for homogenization
pling was conducted to acquire samples of the various para- or freezing study because they result from necking down and
genetic stages. leakage as evidenced by the cooccurrence of all liquid and all
Fluid inclusions were studied in 150 doubly polished, 80- vapor inclusions.
m-thick plates from 41 sample sites. Inclusions were ob- The complete data set for this study is available as a digital
served in stage 2a quartz and K-feldspar, in stage 2b milky supplement to this paper at <www.segweb.org/EG/papers/
quartz, zoned quartz, K-feldspar, johansennite, clinozoisite, in Abs100-7_files/DreierAppendix.pdf>
ore-stage sphalerite, and in postore calcite and quartz. Heat-
ing and melting point determinations were made on a Fluid Stage 2
Inc.-adapted U.S. Geological Survey (USGS) gas-flow heat- Stage 2a (fine-grained quartz-K-feldspar): The cloudy and
ing and/or freezing stage at the USGS fluid inclusion labora- very fine grained character of this material (Fig. 7A) pre-
tory in Lakewood, Colorado. Fluid inclusion samples used in cluded determinations of melting point and homogenization
the heating and ice-melting runs were disks less than 7mm in temperature determinations.
longest dimension and fluid inclusions were placed within 3 Stage 2b: Much of the stage 2b quartz is plumose quartz,
mm of the thermocouple. The stage was calibrated every 5 to chalcedony, or some other form of amorphous to cryptocrys-
6 mo. Heating runs were reproducible to within 1C, and ice- talline silica, and most fluid inclusions in this material are too
melting point runs to within 0.1C. Fluid inclusions were small for study during heating and freezing runs. Further-
classified as primary or secondary by the criteria of Roedder more, postdepositional changes, including leakage and the
(1979). Fluid inclusions were considered primary if they oc- coalescence of fluid from microinclusions or hydrated silica,
curred as random arrays within a crystal or in symmetric compromise the utility of these fluid inclusions (Sander and
growth zones. Fluid inclusions were classified as secondary if Black, 1988). A low temperature of formation for milky quartz

0361-0128/98/000/000-00 $6.00 1336


PURISIMA-COLON VEIN SYSTEM, PACHUCA REAL DEL MONTE DISTRICT, MEXICO 1337

FIG. 9. Fluid inclusion assemblages and types in quartz and sphalerite. A. Type ii fluid inclusions in milky quartz. B.
Zoned quartz grain with 3 growth zones. C. Type rc and negative crystal fluid inclusions in the inner zone. D. Type ii
fluid inclusions in the outer zone. E. Type ic fluid inclusions in the middle zone. F. Secondary, vapor-filled, fluid inclusions
in the inner zone of (B). G. Primary fluid inclusions in sphalerite. H and I. Fluid inclusions attached to chalcopyrite inclu-
sions in sphalerite. Abbreviations: CPY = chalcopyrite; FI = fluid inclusion.

is suggested by the abundance of type ii fluid inclusions >260C from the center of the zoned grain of Figure 9B. Fig-
within it (Figs. 7A-B, 9A-B, D). This suggestion is supported ure 9E is a close-up of the surrounding band defined by type
by the common intergrowth of milky quartz with plumose ic fluid inclusions indicative of formation at 200 to 230C.
quartz and chalcedony, two forms that are interpreted to pre- Figure 9D is a close-up of the outer band of the crystal that is
cipitate at temperatures <180C (Fournier, 1985; Sander and defined by type ii fluid inclusions indicative of temperatures
Black, 1988; Saunders, 1990; Rimstidt, 1997). By contrast, the of <180C. Thus, for the zoned quartz grain of Figure 9B,
dominance of type rc, and negative-crystal fluid inclusions in fluid inclusions indicate overall declining fluid temperatures.
the clear bands of zoned quartz indicates depositional tem- As shown in Figure 6E, cycles of the type illustrated in Fig-
peratures >250C, a conclusion that is supported by homoge- ure 9B were repeated numerous times (four cycles are visible
nization temperatures of 250 to 320C (App., Fig. 10). Some in Fig. 6E) in which the outermost band of type ii fluid inclu-
assemblages of type rc and negative-crystal fluid inclusions are sions might be overgrown by a new cycle beginning with a
all vapor or vapor dominant (Fig. 9F), and these are inter- clear band with type rc negative-crystal fluid inclusions. In
preted to represent the trapping of vapor from a boiling fluid. Figure 6E some of the bands of type ii fluid inclusions are
Figure 9B to E illustrates the zoning of fluid inclusion capped by layers of chalcedony. As a result of the zoning re-
types within zoned quartz and would roughly correspond to lationships noted for the various fluid inclusion types of Fig-
the individual bands of Fig 6E. Figure 9C shows negative- ure 9B, it appears that stage 2b was deposited by fluids that
crystal and type rc fluid inclusions indicative of temperatures fluctuated widely in temperature, from <180 to >310C and

0361-0128/98/000/000-00 $6.00 1337


1338 JOHN E. DREIER

FIG. 10. Histograms of fluid inclusion homogenization temperatures


arranged by host mineral.

that these fluctuations were ordered into cycles such that


each cycle began with the influx of a hot fluid and ended with
a cooler fluid.
If the maximum fluid inclusion homogenization tempera-
tures are arranged by elevation, they provide evidence for a
vertical temperature gradient within the Purisima vein. The
maximum temperatures in samples from the lower mine lev- FIG. 11. Histograms of stage 2b and ore-stage fluid inclusion salinity (wt
els were >300C, whereas in the upper levels they were % NaCl equiv). For stage 2b, each measurement represents a fluid inclusion
~280C (Fig. 3D). assemblage.
Ice-melting temperatures in fluid inclusions from stage 2b
quartz range from 0 to 2.3C (avg = 0.8C; App.). The data zones are virtually identical, but in one grain, inclusions in the
show a slight tendency toward a bimodal distribution with one inner (clear quartz) zone had ice-melting temperatures of
subset in the range of 0 to 0.5C (01 wt % NaCl equiv) and 1.5C, whereas in the outer, cloudy zone they were 0.1C.
the other in the range of 1.0 to 1.5C (23 wt % NaCl For the Purisima vein, sample statistics are similar (0.7C for
equiv; Fig. 11). The bimodal distribution of the data is accen- type ii inclusions and 0.9C for type rc and negative-crystal
tuated by a comparison of fluid inclusion ice-melting temper- inclusions) but suggest that clear quartz, which contains all
atures from the Colon (avg 0.2C) and Purisima veins (avg = type rc and negative-crystal fluid inclusions, was deposited by
1C) and suggests a difference in salinity between the stage slightly more saline fluids than was milky quartz which con-
2 Purisima and Colon fluids (0.3 vs. 1.4 wt % NaCl equiv) if tains most of the type ii fluid inclusions. Ice-melting temper-
the ice-melting temperatures are ascribed entirely to NaCl. A atures also were measured in fluid inclusions in K-feldspar
comparison of ice-melting temperatures from different zones and johansennite. The average ice-melting temperature for
of the same stage 2b quartz grain shows no clear pattern. In these fluid inclusions was 1.4C, consistent with ice-melting
some grains, ice-melting temperatures for inner and outer data in the associated quartz. Collectively, the ice-melting

0361-0128/98/000/000-00 $6.00 1338


PURISIMA-COLON VEIN SYSTEM, PACHUCA REAL DEL MONTE DISTRICT, MEXICO 1339

data from stage 2b minerals indicates that stage 2b was de- high-salinity, high-temperature end of the stage 2b popula-
posited from fluids with salinities in the range of 0 to 3.9 wt tion overlaps the low end of the ore-stage population.
percent NaCl equiv and averaging 1.2 wt percent NaCl equiv
(Fig. 11, App.). Interpretation of Vein Formation
Evidence from geologic mapping and the study of vein
Ore stage structure, petrography, and fluid inclusions suggests that the
Ore-stage fluid inclusions were observed and measured in Purisima-Colon vein system formed in an evolving geother-
sphalerite, in quartz that replaced johansennite, and in sec- mal system wherein the vertical location, geometry, structure,
ondary trails that radiate from sulfides. The fluid inclusions in paragenesis, and Ag and base metals grades of the system re-
sphalerite are primary inclusions controlled by growth bands sulted from the complex interplay of structural geology, fluid
or distributed irregularly within a grain (Fig. 9G) and sec- flow, the behavior of dissolved silica, and metals in hy-
ondary fluid inclusions located along fractures and/or at- drothermal fluids, and variable sources of the fluids.
tached to chalcopyrite, galena, or argentite grains (Fig. 9H-I)
that also occur along fractures. Fluid inclusions attached to Control on the top of the vein system
sulfide inclusions within sphalerite are interpreted to be co- As discussed by Berg (1932) and Wisser (1948), epithermal
genetic with the younger sulfides. veins occupy dilatant zones in faults that resulted from pro-
Homogenization temperatures of ore-stage fluid inclu- gressive separation of the fault walls. In epithermal Ag dis-
sions range from about 250 to 295C and average 280C tricts where a nearly complete stratigraphic section is pre-
(Fig. 10, App.). Ice-melting temperatures for ore-stage fluid served, blind veins are the norm and are variously ascribed to
inclusions range from 1.5 to 4.4C (App.) and average boiling, the development of a clay cap, changes in fault atti-
2.8C; the lowest ice-melting temperatures were measured tude, variations in the mechanical properties of the wall rocks,
in sphalerite from a specimen of massive sulfide collected or the disappearance of the host fault within a mechanically
from below the silver horizon in the Dios Te Guie vein, a incompetent rock unit (Ferguson, 1921; Nolan, 1930, 1935;
north-south vein located about 150 m east of the Purisima Wisser, 1951; Geyne et al., 1963; Heald et al., 1987). Because
vein (Fig. 2). The ice-melting data translate to salinity values the top of the Purisima-Colon vein system ignores stratigra-
of 2.6 to 7 wt percent NaCl equiv (Fig. 11). phy, a major unconformity, and crosscutting dikes and faults,
As shown in Figure 11, the ice-melting data indicate the it does not appear to have been controlled by variations in the
presence of two fluid populations: stage 2b with an average internal mechanical properties of its host rocks. In the ab-
salinity of 1.2 wt percent NaCl equiv and the ore stage sence of multistage fragments and fragments with truncated
with an average salinity of 4.5 wt percent NaCl equiv. The veins, it can be deduced that the initial openings on the host

FIG. 12. Salinity-temperature diagram for stage 2b hydrothermal fluids based on fluid inclusion homogenization and
freezing temperature data. The distribution of the salinity-temperature data suggests that the silicate stage was formed
by three fluids: (1) a low-salinity, <180C fluid possibly of unreacted, shallow meteoric origin; (2) a low- to moderate-
salinity, 250 to 310C reacted fluid; and (3) a low-salinity, 250 to 310C fluid possibly derived by vapor loss from fluid
2. Symbols in the diagram represent fluid inclusion homogenization and salinity determinations for individual fluid in-
clusion assemblages.

0361-0128/98/000/000-00 $6.00 1339


1340 JOHN E. DREIER

fault predated hydrothermal activity. Thus, boiling or the de- presence of epidote which is stable only above >225C (Bird
velopment of a clay cap could not have been involved in the and Helgeson, 1981). Boiling of some stage 2b fluids is sug-
process that caused the upward termination of the veins. In gested by the presence of type rc fluid inclusion assemblages
the case of the Purisima-Colon vein system, two lines of evi- with high homogenization temperatures and coexisting vapor-
dence suggest that the top of the vein is located at the top of and fluid-rich fluid inclusions (Fig. 9F).
the fault: the host fault terminates upward into a zone of dis- A temperature-salinity diagram (Fig. 12) suggests that
tributed fracturing which dies out upward; and the highest stage 2b resulted from three hydrothermal fluids with the fol-
point on the vein system is located in the area of maximum lowing temperatures and salinities: (1) 160 to 200C and av-
vertical offset on the host fault and the lowest points occur at eraging ~ 0.5 to 1 wt percent NaCl equiv; (2) 260 to 310C
the distal ends where offset is nil. The latter point suggests and ~ 2.5 to 3 wt percent NaCl equiv; (3) 260 to 310C and
that the host fault terminates upward where offset falls below averaging ~ 0.5 wt percent NaCl equiv. Interpretations based
some critical value. on fluid inclusion morphology suggest that fluid (1) was
largely responsible for the low-temperature quartz, fluid (2)
Genesis of the rock breccias was responsible for the clear quartz + johansennite + K-
The Purisima-Colon veins are generally coextensive with feldspar epidote hematite and adularia + clinozoisite as-
crudely tabular breccias of angular rock fragments that fill semblages, and fluid (3) was responsible for much or all of the
open zones along the fault; fragments are similar in composi- vein material in the Colon vein. Following Henley (1985),
tion to the adjacent vein walls. As the breccias are interpreted some of the low-salinity, higher temperature fluid inclusions
to predate hydrothermal activity, a passive process, such as in the Purisima vein (fluid 3) may have been derived by vapor
the collapse of rocks into open space, is envisioned for their loss from fluid (2). The formation of stage 2b by the periodic
formation. Based on the presence of remnant fine rock parti- replacement of one type of fluid by another is consistent with
cles and stage 2a quartz-K-feldspar intergrowths in the ma- scenarios proposed for Creede (Barton et al., 1977; Bethke
trix, it is interpreted that the breccias were originally sup- and Rye, 1979; Plumlee and Heald Whitehouse-Veaux, 1994),
ported by finely comminuted rock material (rock flour) that Zacatecas (Albinson, 1995), and elsewhere in Mexico (Al-
was replaced during stage 2a. binson et al., 2001) where fluid inclusion and/or oxygen-deu-
terium isotope studies point to multiple sources for the vein-
Genesis of stage 2 forming fluids.
Prior to the onset of mineralization, breccia fragments The contrasting salinity of the Purisima (up to 3.9 wt %
ranged in diameter from 0.3 m down to a few microns, with NaCl equiv) and Colon (0.020.05 wt % NaCl equiv) fluids
~80 percent of the material >5 mm and ~10 to 15 percent suggests that the two veins were hydrologically isolated from
<0.1 mm (rock flour). In the earliest phase of vein formation, each other during the silicate stage and that the central
hydrothermal fluids dissolved some of the rock flour and al- Purisima vein was above a principal upflow zone in the hy-
tered the remainder to a fine-grained mosaic of quartz + K- drothermal system. The overall zoning of Ag away from the
feldspar. At the same time, most fragments in the 0.1- to 0.5- central Purisima bonanza zone also suggests that the central
mm range were altered to fine- to medium-grained patches of Purisima orebody was an upflow zone.
anhedral quartz + K-feldspar + albite + chlorite epidote. In By analogy with dilute modern geothermal waters (Arnors-
active geothermal systems, mineral assemblages of this type son et al., 1978; Giggenbach, 1981, 1997; Arnorsson and
are interpreted to result from the action of neutral pH hy- Gunnlaugsson, 1983) and as a first approximation, the com-
drothermal fluids (Giggenbach, 1981, 1997). position of the stage 2b fluids may be further constrained by
Stage 2b minerals formed by fluid flow in the tortuous pas- assuming that virtually all of the salinity is due to NaCl and
sageways in stage 2a quartz-K-feldspar intergrowths and in KCl and that the >250C fluids were in equilibrium with al-
linear fractures that crosscut earlier vein material. Petro- bite, K-feldspar, clinozoisite, and hematite but not in equilib-
graphic and fluid inclusion studies indicate that the stage 2b rium with prehnite, wairakite, tremolite, or diopside. Under
fluids fluctuated considerably in temperature, and the study these conditions, log aK+/aH+ is 4.1 to 4.5, log aCa2+/aH+2 is be-
of crystal and vein banding (Figs. 6E, 9B) suggests that tem- tween about 8.5 and 9.1, and log aMg2+/aH+2 is below about 4.5
perature fluctuations were cyclical. While the total number of (Figs. 13, 14). The average salinity value of 1.2 wt percent
cycles is unknown, up to four cycles are recorded in single NaCl equiv for the stage 2 fluids corresponds to log mNaCl =
crystals and veins. The abundance of milky, mosaic, and 0.68. For dilute chloride solutions in equilibrium with albite
plumose quartz, and chalcedony relative to coarse-grained, and K-feldspar at 250C, log aNa+/aK+ 0, aK+ aNa+ (Henley
crystalline quartz suggests that most stage 2b material was de- et al., 1984) and, within the error of the salinity measure-
posited very rapidly from <200C fluids that were supersatu- ments, log mNa mK 0.98. Using the data and approach
rated to highly supersaturated with respect to quartz. This in- of Henley et al. (1984) to calculate activity coefficients gives
terpretation is based on field and laboratory work showing log aK+ 1.3 and log aH+ of 5.4 to 5.8 (pH = 5.45.8). Val-
that precipitation of quartz is kinetically inhibited at temper- ues for log aCa2+ and log aMg2+ of 3.1 to 1.7 and <6.3, re-
atures below ~180C (Fournier, 1985; Sander and Black, spectively, were calculated by substituting the value for log
1988; Saunders, 1990; Rimstidt, 1997). Fluid inclusion ho- aH+ of 6 into the relationships 8.5 < log aCa2+/aH+2 <9.1 and
mogenization data indicate that the assemblages (1) clear log aMg2+/aH+2 <4.5, which correspond to log mCa of 2.2 to
quartz + johansennite epidote hematite and (2) adularia 0.81 and for log mMg of <6.6. The absence of calcite in the
+ clinozoisite were deposited as a result of periodic influxes of stage 2b vein assemblage requires that log PCO2 was <102,
230 to >300C fluids. This conclusion is supported by the well below the level that would require adjustments to

0361-0128/98/000/000-00 $6.00 1340


PURISIMA-COLON VEIN SYSTEM, PACHUCA REAL DEL MONTE DISTRICT, MEXICO 1341

FIG. 14. Activity-activity diagram for the system CaO-Al2O3-MgO-SiO2


balanced on Al2O3 at 250C. Calculated from data in Henley et al. (1984).
Variations in the compositions of chlorite, represented here by chrysotile,
which was reported by Geyne et al. (1963), and clinozoisite are such that the
FIG. 13. Activity-activity diagram for the system CaO-Al2O3-K2O-SiO2 fields of these minerals are only approximate. The substitution of compo-
balanced on Al2O3 at 250C. Calculated from data in Henley et al. (1984). nents into these minerals would probably enlarge their fields in the diagram.
Based on the common vein assemblage K-feldspar-clinozoisite, the silicate- The >250C silicate-stage fluids were within the zoisite field (solid square).
stage fluids were mostly on the zoisite-K-feldspar boundary (solid square). Ore-stage fluids were initially in the chlorite field or along the kaolinite-chlo-
Because clinozoisite forms a continuous solid solution with epidote, the in- rite boundary (solid dot). Reaction of the ore-stage fluids with wall-rock frag-
clusion of variable amounts of Fe3+ within it would probably enlarge the ments and johansennite lowered PCO2 and tended to move fluid composition
zoisite field of the diagram. The absence of calcite in the silicate stage re- toward the zoisite field as shown by the arrow. Symbols as in Figure 13.
quires that PCO2 was lower than 102. The presence of abundant illite + kaoli-
nite indicates that the ore-stage fluids were initially along the muscovite-
kaolinite boundary (solid circle) if vein kaolinite is a primary hydrothermal
mineral as opposed to the product of postmine oxidation. Reaction of the
ore-stage fluids with wall rocks and johansennite tended to decrease PCO2 and
move fluid composition toward the zoisite-adularia field boundary as shown
by the arrows. Abbreviation: Ca-Mont = Ca montmorillonite.

ice-melting temperatures (Hedenquist and Henley, 1985).


The presence of hematite in stage 2b indicates that O2 >1036
(Fig. 15); the relatively oxidized nature of the fluids may have
resulted either from equilibration of the fluids with hematite
in the volcanic rocks or an influx of oxidized meteoric fluids.
These conditions are summarized in Figure 15.
As shown in Figure 16, the silicate-stage fluids were simi-
lar to those circulating in well-studied, dilute, neutral pH ge-
othermal systems as reported by Arnorsson et al. (1978),
Giggenbach (1981, 1997), and Arnorsson and Gunnlaugsson FIG. 15. pH vs. log aO2 diagram for the system S-Fe-O2 at 250C and
(1983). Fluids of this type are considered to represent the end mS = 102, showing the general locations of silicate- and ore-stage fluids.
Lines for stability field boundaries: Fe-O-S minerals = heavy solid; dissolved
product of fluid-rock interaction between deeply circulating sulfur species = light solid; argentite-native Ag = dotted; silicates = dot-dash.
meteoric waters and rocks of the geothermal reservoir (Hen- The stability boundaries for K-mica-K-feldspar using log aK+ = 1.5 and
ley, 1985). This interpretation is supported by oxygen-deu- wairakite-zoisite using log aCa2+ = 3 are based on fluid inclusion ice-melting
terium isotope data from quartz-hosted fluid inclusions from data from stage 2b silicate minerals and calculations described in the text.
The stability boundaries for kaolinite-K-feldspar using log aK+ = 0.5 and K-
Pachuca reported by ONeil and Silberman (1974) that indi- mica-K-feldspar using log aK+ = 0.5 are based on fluid inclusion ice-melting
cate a large amount of isotopic exchange between the stage 2 data from sphalerite and calculations given in the text. Diagram calculated
hydrothermal fluid and the rocks at depth. from data in Henley et al. (1984).

0361-0128/98/000/000-00 $6.00 1341


1342 JOHN E. DREIER

base metals. The presence of ore-stage illite and/or kaolinite


and the absence of montmorillonite and wairakite suggests a
fluid composition as represented by the solid circles in Fig-
ures 13 and 14. Combining equilibria calculations for the ap-
propriate assemblages in Figures 13 and 14 with fluid inclu-
sion salinity data, as described for the stage 2b fluids, shows
that mNa mK 0.5 to 1.5, mCa 105 and mMg ~104 (note
that the calculated mMg value for the ore-stage fluids is ~103
higher than for the stage 2b fluids). The pH of the ore-stage
fluid was in the range of 3.5 to 4.5 as constrained by the kaoli-
nite/K-mica and K-mica/K-feldspar boundaries (Fig. 15). The
presence of pyrite, as opposed to hematite, also indicates a
lower O2 for the ore-stage fluids. Whereas the silicate-stage
fluids closely resemble the deeply circulated meteoric fluids
of modern geothermal systems (Fig. 16), the ore-stage fluids
have intermediate temperatures and salinities, a relatively low
pH, and elevated Mg more similar to the magmatically de-
rived fluids of Giggenbach (1997). Reaction of the ore-stage
fluids with rock fragments and preexisting silicate-stage min-
erals would have driven the composition of the ore-stage fluid
in the direction of the arrows in Figures 13 and 14. The pres-
ence of unaltered johansennite within otherwise altered
masses of rock suggests local equilibration of the ore-stage
fluids with silicate-stage or rock-forming minerals.
The wide association of sulfides and clay alteration pro-
FIG. 16. pH vs. temperature diagram for hydrothermal systems (modi- vides strong evidence that the reaction of H+ with K-feldspar
fied from Henley, 1985). Lines are for the equilibrium between K-mica and and other silicate minerals caused the precipitation of argen-
K-feldspar. Note the similarity between the low-salinitylow-temperature
fluids of stage 2b (fluid 1) and fluids in the Hvergaderi and Te Kopia geot-
tite according to the reaction:
hermal systems and the low- to moderate-salinity, 250 to 310C fluids of
stage 2b (fluid 2) and fluids in the Broadlands, Kawerau, and Wairakei geot-
3KAlSi3O8 + 2AgCl21(aq) + 2H+ + S2(aq) + 12H2O =
hermal systems. The ore-stage fluids are distinctly more acid than those of K-feldspar
most neutral pH, low-salinity geothermal fluids. Data for modern geothermal
systems is from Arnorsson et al. (1978), Giggenbach (1981, 1997), and Hen-
KAl2Si3AlO10(OH)2 + 6H4SiO4(aq) + 2K+(aq) + Ag2S + 4Cl(aq).
ley et al. (1984). illite (1)

Genesis of the ore stage Ag precipitates when AgCl21(aq) is destabilized by the con-
sumption of H+. The abundance of johansennite in the veins
The ore stage is interpreted to represent a discrete event and its widespread alteration to CaMn(CO3)2 quartz sug-
of relatively short duration in which metal-rich fluids of mod- gests that CO2 was controlled by the reaction:
erate salinity penetrated the upper part of the meteoric water
aquifer. The presence of sulfides as microveins, pore and vug 2CO2 + 4H2O + CaMn(SiO3)2 =
fillings, stockwork-like and large, planar veins indicates that johansennite
fluid flow during the ore stage took place in the fine pores of CaMn(CO3)2 + 2H4SiO4(aq). (2)
the veins, the tortuous openings within partly healed frac- kutnohorite
tures, and large, planar, through-going channels that opened
in response to contemporaneous faulting. The abundance of Although thermodynamic data are not presently available for
sulfides replacing silicate minerals requires that a large pro- johansennite and kutnohorite, the structural similarity of
portion of sulfide deposition was somehow related to water- these minerals to diopside and dolomite, respectively, and the
rock reaction. By contrast with stage 2b, the absence of quartz formation of continuous solid solutions by the johansennite-
in the ore stage is notable and has been reported from other diopside and kutnohorite-dolomite pairs suggests that the al-
important low total sulfide epithermal Ag districts, including teration of diopside to dolomite can be used as a proxy for the
the Comstock Lode (Becker, 1882) and Tonopah (Bastin and carbonation of johansennite:
Laney, 1918; Nolan, 1930, 1935), and is implied by ice-melt-
ing and isotopic data from fluid inclusions at Creede, Col- CaMg(SiO3)2 + 2CO2 = CaMg(CO3)2 + 2SiO2. (3)
orado, and many districts in Mexico (Bethke and Rye, 1979; diopside dolomite
Albinson et al., 2001). The absence of quartz in the ore stage
is a widespread characteristic of low total sulfide epithermal Using thermodynamic data for diopside, dolomite, and
Ag vein deposits and may be fundamental to the genesis of CO2 from Helgeson (1969), PCO2 is found to be ~6 bars at
these systems. 275C and ~10 bars at 290C and would contribute about
The ore-stage fluids also differed from those of the silicate 0.5C to fluid inclusion melting point depressions. Similar
stage in pH, PCO2, O2, S, and concentrations of Mg, Ag, and values for PCO2 are attained by applying the rhodochrosite +

0361-0128/98/000/000-00 $6.00 1342


PURISIMA-COLON VEIN SYSTEM, PACHUCA REAL DEL MONTE DISTRICT, MEXICO 1343

rhodonite + quartz and calcite + chalcedony + calc-silicate point where boiling took place. On this basis and in the
buffers of Barton and Chou (1993). absence of vapor-dominant fluid inclusions or mineralogical
Clay caps to epithermal veins and geothermal systems are evidence of boiling such as ore-stage K-feldspar, fine-grained
widespread (Heald et al., 1987; Simmons and Browne, 2000) quartz and/or chalcedony, and lamellar calcite (Simmons and
and have been ascribed to a model in which boiling of as- Browne, 2000), it is concluded that boiling was not an impor-
cending hydrothermal fluids at or very close to the top of a tant factor in ore deposition. By contrast with the ore-stage
ore horizon releases H2S and CO2 from the geothermal reser- fluids, a plot of the highest fluid inclusion homogenization
voir that rises above the veins and then condenses on contact temperatures from stage 2b is located along a boiling curve
with cold descending meteoric water. Figure 17 shows boiling for water based on the interpreted depth to the top of the
point curves for ore-stage fluids of 5 and 10 wt percent NaCl vein system at the time of mineralization (Fig. 17).
equiv adjusted for PCO2 values noted above. The depth-tem- Deep drilling indicates that the Ag horizon of the
perature profile for the ore-stage fluids derived from fluid in- Purisima-Colon vein system represents the upper part of a
clusion homogenization temperatures intersects the boiling very large zoned base metal-Ag system. In this respect, the
point curve for a 5 percent NaCl fluid and PCO2 = 6 to 10 bars Purisima-Colon vein system is similar to other studied ep-
at a depth of about 570 m and for a 10 percent NaCl fluid at ithermal Ag vein districts, including Tonopah (Nolan, 1930),
650 m. This depth range is consistent with geologic estimates Guanajuato (Randall et al., 1994), Creede (Heald et al.,
of 550 to 650 m for the top of the vein system (Fig. 7). The 1987), Zacatecas (Albinson 1995), and Bolanos (Lyons, 1988).
dashed line in Figure 17 gives the elevation of the top of the The lowest ice-melting temperatures were measured in
Purisima vein at the time of mineralization, and the shaded yellow sphalerite free of chalcopyrite disease from below the
area represents the zone of illite alteration above the vein. If silver ore horizon. The highest ice-melting temperatures are
the region of illite alteration was at or just above the boiling from secondary fluid inclusions in sphalerite that are associ-
zone, then the bonanza Ag horizon was located well below the ated with argentite, galena, and chalcopyrite. Together, these
data suggest that the earlier and deeper fluids were more
saline than the later, higher fluids. This notion is supported by
Figure 18, a salinity-enthalpy diagram (Fournier, 1979; Hen-
ley et al., 1984), in which the ore-stage fluids lie along a hy-
pothetical mixing line between the low-temperature, dilute,
silicate-stage fluids (similar in geochemistry to deeply circu-
lated meteoric waters of modern neutral pH geothermal sys-
tems) and magmatically derived ore forming fluids of
mesothermal vein/manto-skarn base metal-silver systems (see
figs. 2, 3b of White, 1974; Sawkins, 1964; Simmons et al.,
1988). It is proposed that the Pachuca ore-stage fluid repre-
sents the upper, cooler, more dilute derivative of a hotter and
much more saline, magmatically derived mesothermal vein-
and/or skarn-forming fluid at depth. That fluids in even the
deepest mine levels at Pachuca are distinctly less saline than
those at Provedencia and Fresnillo (Sawkins, 1964; Simmons
et al., 1988) requires that fluid mixing began well below the
silver horizon.
Salinity data in conjunction with calculations of Cole and
Drummond (1986) indicated that the base metals and silver
were transported almost entirely as chloride complexes; data
from Henley et al. (1984) indicated that AgCl2- was the
strongest Ag-Cl complex. The dilution of a 300C brine satu-
rated in Ag2S from 1- to 0.5-m NaCl precipitates about 75
percent of the Ag, a pH increase from 4 to 5 precipitates >95
percent of the Ag, and cooling the brine from 300 to 250C
FIG. 17. Depth vs. temperature profile through the Purisima-Colon vein
system at ~11,000 N, showing the ore-stage and high-temperature silicate- precipitates about 95 percent of the Ag (Henley et al., 1984).
stage fluids in relationship to the top of the vein, the position of the clay-cal- These calculations are consistent with a model in which the
cite alteration zone, the Ag horizon, and the bonanza zone based on fluid in- bottom of the Ag horizon represents a surface at which the
clusion ice-melting and homogenization data. The boiling point curves combined effects of dilution, cooling, and pH increase due to
assume a hydrostatic pressure gradient (Hedenquist and Henley, 1985). The
silicate-stage boiling point curve (dotted line) was calculated for pure water.
fluid-rock reaction drove fluid composition to Ag2S satura-
The 5 and 10 percent NaCl equiv boiling point curves (solid lines) were ad- tion. The very late stage native Ag (and covellite) indicates a
justed to show the effect of PCO2 calculated from the reaction: diopside + CO2 late increase in O2 (Fig. 17) or the exhaustion of dissolved S.
= dolomite + quartz. The ore-stage depth-temperature profile (dashed line) That the high-grade Ag zones within the Purisima-Colon
is based on the average sphalerite fluid inclusion homogenization tempera- vein system occur generally within the widest vein segments
ture for each mine level and intersects the boiling curves for 5 and 10 per-
cent NaCl solutions in the depth range 650 to 570 m (between the 150- and provides strong evidence that Ag emplacement was directly
50-m mine levels). Milky quartz, plumose quartz, and chalcedony of stage 2b related to rates of fluid flow. This is also evident at the district
were deposited at temperatures well below the boiling curve. scale. Areas 1 to 3 of Figure 2, which host all of the great

0361-0128/98/000/000-00 $6.00 1343


1344 JOHN E. DREIER

FIG. 18. Salinity vs. enthalpy diagram showing the relationship between fluids of the silicate and ore stages, Purisima-
Colon vein system, and fluids from vein-manto-skarn systems elsewhere in Mexico (Sawkins, 1964; Simmons, 1988). Salinity
values for the Purisima-Colon ore-stage fluids, as represented by fluid inclusion data from sphalerite, lie on a mixing line be-
tween silicate-stage fluids of postulated wall-rock equilibrated meteoric origin and saline, mesothermal skarn-manto fluids of
Sawkins (1964) and Simmons et al. (1988). Following Fournier (1979) and Henley (1985), it is suggested that the ore-stage
fluids represent an original magmatic fluid (about one part) diluted by an equilibrated meteoric fluid (about one to two parts).

bonanzas, are zones of intense faulting and complex fault in- the postore stage provide evidence that hydrothermal activity
tersections that appear to have acted as large zones of hy- terminated before the last openings created by faulting could
drothermal upflow. be filled by mineral precipitates.
Whereas mineral zoning indicates horizontal flow in the
Creede district and the upper part of the Broadlands-Ohaaki Conclusions
geothermal system (Plumlee and Heald Whitehouse-Veaux, The Purisima-Colon veins were formed in open zones in
1994; Simmons and Browne, 2000), vertically zoned wall-rock normal faults at minimum depths of 550 to 650 m. The ele-
alteration, sulfide mineralogy, and isotherms point to vertical vation of the top of the vein system is positively related to the
flow in the Purisima-Colon vein system (Geyne, 1949; Wisser, amount of offset on the fault system and probably corre-
1951). Finally, geologic mapping and the study of level maps sponds to the location where the fault dies out. Prior to the
indicate that the major vein segments along the Purisima- onset of mineralization the open zones in the fault system
Colon vein system were separated from each other by barren were filled with rock fragments.
faults or areas of narrow, sporadic, barren to weakly mineral- Mineral deposition took place in three stages from hy-
ized veins. drothermal fluids that flowed directly up the veins to the top
The intense illite-pyrite alteration and quartz and/or cal- of the Ag horizon and then subhorizontally within the zone of
cite veining that pervades rocks for up to 350 m above the distributed fracturing and illite-pyrite alteration. The first
veins indicates that a significant proportion of the fluid flow stage of mineral deposition was characterized by conditions of
rose well above the top of the vein system, but the absence or fluctuating temperature or temperature cycles and is domi-
scarcity of faults, fractures, alteration, and veining above 350 nated by quartz. Much of the quartz was deposited as chal-
m provides strong evidence for vertical limits to upward cedony or cryptocrystalline material from hydrothermal fluids
flowi.e., that the system was capped at ~350 m above the that were highly supersaturated in silica at temperatures
top of the vein system. After the hydrothermal fluids exited <180C. Some of the quartz, however, was deposited as
the top of the vein system they must have flowed subhorizon- quartz between ~230 and 310C and the higher temperature
tally in the zone of distributed fracturing and clay alteration fluids were boiling. The silicate stage, which also includes K-
where they mixed with downward flowing ground water, as feldspar, johansennite, epidote, albite, clinozoisite, and
suggested by the low-salinity fluid inclusions from this region hematite, was deposited from dilute-very dilute, oxidized,
(Dreier, 1976, Fig. 19). neutral pH hydrothermal fluids similar to those circulating in
many dilute, neutral pH modern geothermal systems. Such
Genesis of the postore stage geothermal fluids are interpreted by Henley (1985) as deeply
The abundance of amethyst and calcite indicates that the circulated, wall-rock equilibrated meteoric fluids.
postore fluids were relatively oxidizing and that PCO2 was one The second stage of mineral deposition was deposited
to two orders of magnitude higher than in the stage 2 fluids. from acid-reducing, moderate-salinity fluids at temperatures
The high PCO2 of the fluids probably resulted from the of 270 to 290C that were similar to those proposed as mag-
ingress of meteoric fluids into the veins as the hydrothermal matic derivatives by Giggenbach (1997). The ore-stage fluids
system collapsed. The large open cavities characteristic of at Pachuca and possibly at many other epithermal Ag deposits

0361-0128/98/000/000-00 $6.00 1344


PURISIMA-COLON VEIN SYSTEM, PACHUCA REAL DEL MONTE DISTRICT, MEXICO 1345

FIG. 19. Schematic longitudinal projection of the Purisima-Colon vein system, illustrating fluid flow in relationship to
principal geologic and alteration features during vein formation. Data from geologic and fluid inclusion studies suggest that
the vein-forming fluids flowed directly up the fault and/or veins. They flowed out of the top of the open zone in the fault and
then horizontally away from the principal site of upwelling within the fracture zone above the vein system, driven by down-
ward-flowing meteoric waters. The absence of longitudinal mineral zoning along the vein system of the type reported for the
Broadlands-Ohaaki system by Simmons and Browne (2000) and the presence of barren fault segments between principal
orebodies indicate that significant horizontal, return flow did not take place within the ore horizon. The low-salinity fluid in-
clusions from outside the upflow zone (Dreier 1976) result from dilution by descending meteoric waters.

did not deposit quartz. An important implication of this ob- epithermal environment. Finally, the fluid inclusion study was
servation is that fluid inclusions in quartz should not be relied completed with much input and assistance from T. J.
upon to provide meaningful information on the genesis of Reynolds who taught me the basics of fluid inclusion study,
economic Ag mineralization and that genetic models based spent many hours looking at fluid inclusions, discussing con-
on such studies should be reevaluated. Ag deposition resulted cepts, and reviewing this manuscript and who took most of
largely from fluid-rock acid neutralization reactions but dilu- the photomicrographs.
tion and cooling of the ore fluids by deeply circulated, equili-
brated meteoric waters were contributing factors. July 12, 2002; July 12, 2005
The postore stage was deposited from dilute, oxidizing me-
REFERENCES
teoric fluids circulating as the hydrothermal system collapsed.
Aguilera, J.G., Ordonez, E., Sanchez, P.C., Rangel, M., Gonzalez, I.O., and
Acknowledgments Castro, C., 1897, El Mineral de Pachuca: Instituto Geologico de Mexico,
Boletins 79, 183 p.
I thank S.R. Titley for suggesting the study area, helping Albinson, T., 1995, Bosquejo de evolucion estructural e hidrothermal del dis-
to arrange funding, guiding the early research, and twice re- trito de Zacatecas: Convencion nacional, 21st, Acapulco, Guerrero, October
viewing this manuscript. The Consejo de Recursos de Mex- 1821, Memorias Tecnicas, p. 114.
Albinson, T., Norman, D.I., Cole, D., and Chomiak, B., 2001, Controls on
ico funded initial field work and I thank the late Guillermo formation of low-sulfidation epithermal deposits in Mexico: Constraints
Salas, former Director of the Consejo, for his support, inter- from fluid inclusion and stable isotope data: Society of Economic Geolo-
est, encouragement, and input. Petrographic and fluid inclu- gists Special Publication 8, p.132.
sion work was carried out at the Colorado School of Mines Arnorsson, S., Gronvold, K., and Sigurdsson, S., 1978, Aquifer chemistry of
four high-temperature geothermal systems in Iceland: Geochimica et Cos-
and the U.S. Geological Survey in Lakewood and I thank mochimica Acta, v. 42, p. 523536.
Murray Hitzman and David Leach for the use of their equip- Arnorsson, S., and Gunnlaugsson, E., 1983, The chemistry of geothermal wa-
ment. Reviews by Sam Romberger, Jeff Hedenquest, Mark ters in Iceland. II. Mineral equilibria and independent variables controlling
Hannington, and Tawn Albinson provided many helpful water compositions: Geochimica et Cosmochimica Acta, v. 47, p. 547566.
comments and greatly improved the manuscript. Robert Barton, P.B., Jr., and Bethke, P.M., 1987, Chalcopyrite disease in sphalerite:
Pathology and epidemiology: American Mineralogist, v. 72, p.451467.
Fournier and Paul Barton provided stimulating discussions Barton, P.B., Jr., and Chou, I.M., 1993, Refinement of the evaluation of the
on the interpretation of the various forms of silica and on the role of CO2 in modifying estimates of the pressure of epithermal mineral-
possible scenarios conducive to silica undersaturation in the ization: ECONOMIC GEOLOGY, v. 88, p. 873884.

0361-0128/98/000/000-00 $6.00 1345


1346 JOHN E. DREIER

Barton, P.B., Jr., Bethke, P.M., and Roedder, E., 1977, Environment of ore 1968, Guides for the interpretation of barren and weak surface expres-
deposition in the Creede mining district, San Juan Mountains, Colorado: sion of buried orebodies in the Pachuca-Real del Monte district, Mexico:
Part III. Progress toward interpretation of the chemistry of the ore-form- ECONOMIC GEOLOGY, v. 63, p. 702.
ing fluid for the OH vein: ECONOMIC GEOLOGY, v. 72, p. 124. Geyne, A.R., Fries, C., Jr., Segerstrom, K., Black, R.F., and Wilson, I.F.,
Bastin, E.S., 1948, Mineral relationships in the ores of Pachuca and Real del 1963, Geology and mineral deposits of the Pachuca-Real del Monte dis-
Monte, Mexico: ECONOMIC GEOLOGY, v. 43, p. 5365. trict, State of Hidalgo, Mexico: Consejo de Recursos Naturales no Renov-
Bastin, E.S., and Laney, F.B., 1918, The genesis of the ores at Tonopah, ables Publication 5E, 203 p.
Nevada: U.S. Geological Survey Professional Paper 104, 50 p. 1990, Geology and mineral deposits of the Pachuca Real del Monte dis-
Becker, G.F., 1882, Geology of the Comstock Lode and the Washoe district: trict, Hidalgo, Mexico: Society of Economic Geologists Guidebook, v. 6, p.
U.S. Geological Survey Monograph 3, 422 p. 241257.
Berg, G., 1932, Vein filling during the opening of fissures: ECONOMIC GEOL- Giggenbach, W.F., 1981, Geothermal mineral equilibria: Geochimica et Cos-
OGY, v. 27, p. 8794. mochimica Acta, v. 45, p. 393410.
Bethke, P.M., and Rye, R.O., 1979, Environment of ore deposition in the 1997, The origin and evolution of fluids in magmatic-hydrothermal sys-
Creede mining district, San Juan Mountains, Colorado: Part IV. Source of tems, in H.L.Barnes ed., Geochemistry of hydrothermal ore deposits, 3rd
fluids from oxygen, hydrogen, and carbon isotope studies: ECONOMIC GE- ed.: New York, John Wiley and Sons, p. 737796.
OLOGY, v. 74, p. 18321851. Goldstein, R.H., and Reynolds, T.J., 1994, Systematics of fluid inclusions in
Bird, D.K., and Helgeson, H.C., 1981, Chemical interaction of aqueous so- diagenetic minerals: Society for Sedimentary Geology Short Course v. 31,
lutions with epidote-feldspar mineral assemblages in geologic systems. II. 199 p.
Equilibrium constraints in metamorphic/geothermal processes: American Heald, P., Foley, N.K., and Hayba, D.O., 1987, Comparative anatomy of vol-
Journal of Science, v. 281, p. 576614. canic-hosted epithermal deposits: Acid-sulfate and adularia-sericite types:
Bodnar, R.J., 1993, Revised equation and table for determining the freezing ECONOMIC GEOLOGY, v. 82 p. 126.
point depression of H2O-NaCl solutions: Geochimica et Cosmochimica Hedenquist, J.W., Arribas, R.A., and Gonzalez-Urien, E., 2000, Exploration
Acta, v. 57, p. 683684. for epithermal gold deposits: Reviews in Economic Geology, v. 13, p.
Bodnar, R.J., Reynolds, T.J., and Kuehn, C.A., 1985, Fluid inclusion system- 245277.
atics in epithermal systems: Reviews in Economic Geology, v. 2, p. 4559. Hedenquist, J.W., and Henley, R.W., 1985, The importance of CO2 on freez-
Bonham, H., 1986, Models for volcanic-hosted epithermal precious-metal ing point measurements of fluid inclusions: Evidence from active geother-
deposits: A review: International Geological Congress, Symposium V, mal systems and implications for epithermal ore deposition: ECONOMIC
Hamilton, New Zealand, February 1986, Proceedings, p. 1317. GEOLOGY, v. 80, p. 13791406.
Bortnikov, N.S., Genkin, A.D., Dobrovolskaya, M.G., Muravitskaya, G.N., Helgeson, H.C., 1969, Thermodynamics of hydrothermal systems at elevted
and Filimonova, A.A., 1991, The nature of chalcopyrite inclusions in spha- temperatures and pressures: American Journal of Science, v. 267, p. 729804.
lerite: Exsolution, coprecipitation, or disease?: ECONOMIC GEOLOGY, v. Henley, R.W., 1985, The geothermal framework for epithermal deposits: Re-
86, p. 10701082. views in Economic Geology, v. 2, p. 124.
Buchannan, L.J., 1979, The Las Torres mine, Guanajuato, Mexico: Ore con- Henley, R.W., and Ellis, A.J., 1983, Geothermal systems ancient and modern:
trols of a fossil geothermal system: Unpublished Ph.D. dissertation, A geochemical review: Earth Science Reviews, v. 19, p. 150.
Golden, CO, Colorado School of Mines, 156 p. Henley, R.W., Truesdell, A.H., Barton, P.B., Jr, and Whitney, J.A., 1984,
1980, Ore controls of vertically stacked deposits, Guanajuato, Mexico: Fluid-mineral equilibria in hydrothermal systems: Reviews in Economic
Society of Mining Engineers of the American Institute of Mining, Metal- Geology v. 1, 267 p.
lurgical and Petroleum Engineers Preprint 8082, 26 p. Henry, C., and Aranda-Gomez, J., 1992, The real southern Basin and Range:
Burkart, H.J., 1836, Aufenhalt und Reisen in Mexico in den Jahren 1825 bis Mid- to late Cenozoic extension in Mexico: Geology, v. 20, p. 701704.
1834: Stuttgart, Germany, 678 p. Krupp, R.E., and Seward, T.M., 1987, The Rotokawa geothermal system,
Clarke, M., and Titley, S.R., 1988, Hydrothermal evolution in the formation New Zealand: An active epithermal gold-depositing environment: ECO-
of silver-gold veins in the Tayoltita mine, San Dimas district, Mexico: ECO- NOMIC GEOLOGY, v. 82, p. 11091129.
NOMIC GEOLOGY, v. 83, p. 18301840. Lyons, J.I., 1988, Geology and ore deposits of the Bolanos silver district,
Cole, D.R., and Drummond, S.E., 1986, The effect of transport and boiling Jalisco, Mexico: ECONOMIC GEOLOGY, v. 83, p. 15601582.
on Ag/Au ratios in hydrothermal solutions: A preliminary assessment and McGrath, J.M., 1950, Santa Julia plant of Cia. de Real del Monte y Pachuca:
possible implications for the formation of epithermal precious-metal ore Denver, Colorado, Denver Equipment Company Bulletin M4B103, 6 p.
deposits: Journal of Geochemical Exploration, v. 25, p. 4579. McKee, E.H., Dreier, J.E., and Noble, D.C., 1992, Early Miocene hy-
Delgado-Granados, H., Aguierre-Diaz, G.J., and Stock, J.M., 2000, Preface, drothermal activity at Pachuca-Real del Monte, Mexico: An example of
in Cenozoic tectonics and volcanism of Mexico: Geological Society of space-time association of volcanism and epithermal Ag-Au vein mineraliza-
America Special Paper 334, p. v-vii. tion: ECONOMIC GEOLOGY, v. 87, p. 16351637.
Dreier, J.E., 1976, Geochemical environment of ore deposition in the Nolan, T.B., 1930, The underground geology of the western part of the
Pachuca Real del Monte district, Hidalgo, Mexico: Unpublished Ph. D. dis- Tonopah mining district, Nevada: University of Nevada Bulletin v. 24, no.
sertation, Tucson, AZ, University of Arizona, 126 p. 4, 35 p.
1982, Distribution of wall rock alteration and trace elements in the 1935, The underground geology of the Tonopah mining district,
Pachuca-Real del Monte district, Hidalgo, Mexico: Mining Engineering, v. Nevada: University of Nevada Bulletin v. 29, no. 5, 49 p.
34, p. 699704. ONeil, J.R., and Silberman, M.L., 1974, Stable isotope relationships in ep-
1984, Regional tectonic control of epithermal veins in the western ithermal Au-Ag deposits: ECONOMIC GEOLOGY, v. 69, p. 902909.
United States and Mexico: Arizona Geological Society Digest, v. 15, p. Ordonez, E., 1902, The mining district of Pachuca: American Institute of
2850. Mining Engineers Transactions, v. 32, p. 224241.
Ewers, G.R., and Keys, R.R., 1977, Volatile and precious metal zoning in the Ordonez, E., and Rangel, M., 1899, El Real Del Monte: Instituto Geologico
Broadlands geothermal field: ECONOMIC GEOLOGY, v. 72, p. 13371354. de Mexico Boletin 12, 105 p.
Fahley, P.M., 1981, Fluid inclusion study of the Tonopah district, Nevada: Plumlee, G.S., and Heald Whitehouse-Veaux, P., 1994, Mineralogy, paragen-
Unpublished M.Sc. thesis, Golden, CO, Colorado School of Mines, 106 p. esis, and mineral zoning of the Bulldog Mountain vein system, Creede dis-
Ferguson, H.G., 1921, The Mogollon district New Mexico: U.S. Geological trict, Colorado, ECONOMIC GEOLOGY, v. 89, p. 18831905.
Survey Bulletin 715L, p. 174204. Randall R, J.A., Saldana A, E., and Clark, K.F., 1994, Exploration in a vol-
Fournier, R.O., 1979, Geochemical and hydrologic considerations and the cano-plutonic center at Guanajuato, Mexico: ECONOMIC GEOLOGY, v. 89, p.
use of enthalpy-chloride diagrams in the prediction of underground condi- 17721751.
tions in hot-spring systems: Journal of Volcanology and Geothermal Re- Rimstidt, J.D., 1997, Gangue mineral transport and deposition, in Barnes,
search, v. 5, p. 116. H.L., Geochemistry of hydrothermal ore deposits, 3rd ed.: New York, John
1985, The behavior of silica in hydrothermal solutions: Reviews in Eco- Wiley and Sons, p. 487515.
nomic Geology v. 2, p. 4559. Roedder, E., 1979, Fluid inclusions as samples of ore fluids, in Barnes, H.L.,
Geyne, A.R., 1949, Mineral relationships in the ores of Pachuca and Real del ed., Geochemistry of hydrothermal ore deposits, 2nd ed.: New York, Wiley,
Monte, Hidalgo, Mexicoa reply: ECONOMIC GEOLOGY, v. 44, p. 233234. p. 684737.

0361-0128/98/000/000-00 $6.00 1346


PURISIMA-COLON VEIN SYSTEM, PACHUCA REAL DEL MONTE DISTRICT, MEXICO 1347

Sander, M.V., and Black, J.E., 1988, Crystallization and recrystallization of Turban, F., 1947, Estudio de la mineralizacion a la profundidad en los min-
growth-zoned vein quartz crystals from epithermal systemsimplications erales de Pachuca y Real del Monte, Estado de Hidalgo: Minas y Petrolio
for fluid inclusion studies: ECONOMIC GEOLOGY, v. 83, p. 10521060. Boletin 16, p. 36.
Saunders, J.A., 1990, Colloidal transport of gold and silica in epithermal pre- Vikre, P.G., 1980, Fluid inclusions in silver-antimony-arsenic minerals from
cious-metal systems: Evidence from the Sleeper deposit, Nevada: Geology, precious metal deposits: ECONOMIC GEOLOGY, v. 75, p. 338339.
v. 18, p. 757760. White, D.E., 1974, Diverse origins of hydrothermal ore fluids: ECONOMIC
Sawkins, F.J., 1964, Lead-zinc ore deposition in light of fluid inclusion stud- GEOLOGY, v. 69, p. 954973.
ies, Providencia mine, Zacatecas, Mexico: ECONOMIC GEOLOGY, v. 59, p. 1981, Active geothermal systems and hydrothermal ore deposits: ECO-
883919. NOMIC GEOLOGY 75TH ANNIVERSARY VOLUME, p. 392423.
Simmons, S.F, and Browne, P.R.L., 2000, Hydrothermal minerals and pre- White, N.C., and Hedenquist, J.W., 1990, Epithermal environments and
cious metals in the Broadlands-Ohaaki geothermal system: Implications for styles of mineralization: Variations and their causes, and guidelines for ex-
understanding low-sulfidation epithermal environments: ECONOMIC GEOL- ploration: Journal of Geochemical Exploration, v. 36, p. 445474.
OGY, v. 95, p.971999. Wisser, E., 1937, Formation of the north-south fractures of the Real del
Simmons, S.F., Gemmell, J.B., and Sawkins, F.J., 1988, The Santo Nino sil- Monte area, Pachuca silver district, Mexico: American Institute of Mining
ver-lead-zinc vein, Fresnillo district, Zacatecas: Part II. Physical and chem- Engineers Transactions, v. 126, p. 442486.
ical nature of ore-forming solutions: ECONOMIC GEOLOGY, v. 83, p. 1941a, The environment of ore bodies: American Institute of Mining
16191641. Engineers Transactions, v. 144, p. 96110.
Spycher, N.F., and Reed, M.H., 1989, Evolution of a Broadlands-type ep- 1941b, Discussion, in McKinstry, H.E., ed., Symposium on some obser-
ithermal ore fluid along alternate P-T paths: Implications for the transport vations in ore search: American Institute of Mining Engineers Transac-
and deposition of base, precious, and volatile metals: ECONOMIC GEOLOGY, tions, v. 144, p. 140145.
v. 84, p.328359. 1942, The Pachuca Silver district, Mexico, in Newhouse, W.H., ed. Ore
Steven, T.A., and Ratte, J., 1965, Geology and structural control of ore depo- deposits as related to structural features: Princeton, N.J., Princeton Uni-
sition in the Creede district, San Juan Mountains, Colorado: U.S. Geologi- versity Press, p. 229235.
cal Survey Professional Paper 487, 87 p. 1946, Some applications of structural geology to mining in the Pachuca-
Suter, M., Lopez-Martinez, M., Quintero-Legorreta, O., and Carillo-Mar- Real del Monte area, Pachuca silver district, Mexicoa reply: ECONOMIC
tinez, M., 2001, Quaternary intra-arc extension in the central Trans-Mexi- GEOLOGY, v. 41, p. 7786.
can volcanic belt: Geological Society of America Bulletin 113, p. 693703. 1948, Mineral relationships in the ores of Pachuca and Real del Monte,
Thornburg, C.L., 1945, Some applications of structural geology to mining in Hidalgo, Mexicoa reply: ECONOMIC GEOLOGY, v. 43, p.280292.
the Pachuca-Real del Monte area, Pachuca silver district, Mexico: ECO- 1951, Tectonic analysis of a mining district, Pachuca, Hidalgo, Mexico:
NOMIC GEOLOGY, v. 40, p. 283297. ECONOMIC GEOLOGY, v. 46, p. 459477.
1952, The surface expression of veins in the Pachuca silver district, Mex- 1964, Geology and mineral deposits of the Pachuca-Real del Monte dis-
ico: Mining Engineering v. 4, p. 594600. trict, Hidalgo, Mexicoa review: ECONOMIC GEOLOGY, v. 59, p. 725732.

0361-0128/98/000/000-00 $6.00 1347

Вам также может понравиться