Geology of the Jemez Area, Chapter 5: The Early Tertiary

The table of contents may be found here.

Impact event
Artist's impression of Chicxulub imact.. NASA

The age of the dinosaurs ended in global catastrophe. Traces of this event are found in New Mexico, though not in the Jemez, where the relevant beds were eroded away during the Laramide Orogeny.

In this chapter, we will look at the Jemez area during the early Tertiary Period.

  1. Life on Earth Takes Another Hammering
  2. The Ojo Alamo Formation
  3. The Paleogene in the Jemez
    1. Faults
    2. Fission-track ages
    3. The Paleocene Epoch
      1. The Nacimiento Formation
    4. The Eocene Epoch
      1. The San Jose Formation
      2. El Rito Formation
      3. Diamond Tail and Galisteo Formations
    5. The Oligocene Epoch
      1. Ortiz Monzonite
      2. Espinaso Formation
        1. Debris flows
  4. The Birth of the Rio Grande RIft
    1. The  Cieneguilla Basanite
    2. Feldspathoids and silica saturation
    3. Ritito Conglomerate and Abiquiu Formation
      1. The Gilman Conglomerate
      2. Dikes
  5. The beginning of the Neogene and the Santa Fe Group
    1. "Every valley shall be exalted, and every mountain and hill made low"
    2. The Lower Santa Fe Group
      1. Tanos Formation
      2. The Zia Formation
      3. The Tesuque Formation
        1. The Chama-El Rito Member
        2. The Ojo Caliente Sandstone
        3. Marker Beds
      4. The Cerro Conejo Formation
      5. The Chamita Formation

Life on Earth Takes Another Hammering

The Mesozoic Era ends, and the Cenozoic Era begins, with a mass extinction, albeit one less catastrophic than the Great Dying at the end of the Permian. However, what it lacked in destructiveness was made up in drama and knowledge: We know a lot more about this much more recent extinction. Unfortunately, no record of this is preserved in the immediate Jemez area.

You have almost certainly heard that the age of the dinosaurs ended with a catastrophic asteroid impact near what is now the northern shore of the Yucatan Peninsula. This Chicxulub event has been popularized to the point where there is even a Pixar children's movie depicting it. The global consequences of this impact included the destruction of most of the dinosaurs and many other families of life. Only those dinosaur families whose descendants became modern birds survived the impact. Enough reptiles and mammals survived that they, too, became part of the fauna of the Tertiary. In fact, conditions became suitable for mammals to become the dominant form of land animal, leading ultimately to the ascent of Man.

The Cretaceous-Tertiary boundary is marked in many places in the world by a thin layer of clay that is enriched with traces of the noble metal, iridium. Iridium is extremely scarce in the universe, but even more so in the Earth's crust, since most of the small amount with which the Earth formed settled into the Earth's core along with iron and other heavy elements. Iridium is much more abundant (if still very scarce) in meteorites, which have over 500 times the iridium abundance of average crust rock. This was the first clue that an asteroid impact was the main trigger for the mass extinction. Eventually the deeply buried traces of the impact crater were discovered, mostly by petroleum geologists, and there is now a broad consensus that the asteroid impact was the major cause of the extinction. (There is still a vigorous debate about what other factors may have contributed.)

Although the Cretaceous-Tertiary boundary is not found in the Jemez, it is well exposed near Raton, at the K-T Boundary Trail.

Meloche Mesa
K-T Boundary Trail. 36 54.198N 104 27.077W

The hill is underlain by the Raton Formation, which is mostly shale with interbedded coal.

Here's a close up of the boundary layer.

K-T boundary
K-T Boundary Trail. 36 54.198N 104 27.077W

The boundary is the thin light layer separating lighter shale beneath from darker shale above.

The Ojo Alamo Formation

Digital relief map of Ojo Alamo Formation exposures in
        the Jemez Mountains
Relief map of the Jemez with Ojo Alamo Formation outcroppings highlighted in red.

By the end of the Cretaceous, the Laramide Orogeny was in full swing. Buckling and faulting of the crust had first become severe in Wyoming and Montana but began to spread south. By the beginning of the Tertiary, a large area was beginning to be thrown up in southern Colorado and northern New Mexico very close to the location of the ancient Uncompaghre Uplift. This is known as the Brazos-Sangre de Cristo Geanticline. A geanticline is a large-scale upwarp in the Earth's crust, and this uplift included the entire area from the present-day Tusas Mountains to the present-day Sangre de Cristo Mountains. This uplift began shedding sediments into the surrounding basins.

West of the Sierra Nacimiento, there are exposures of the Ojo Alamo Formation, whose exact age has been much debated. The current consensus seems to be that it straddles the Cretaceous-Tertiary boundary, with dinosaur fossils in its lowest beds and early Eocene fossils further up. There are several small discontinuities in the formation, and, unfortunately, the exact Cretaceous-Tertiary boundary is missing in one of these discontinuities.

Ojo Alamo Formation
          south of Cuba
Ojo Alamo Formation south of Cuba. 35 57.865N 106 59.446W

The Ojo Alamo shows a coarsening upwards sequence. The lower beds tend to be shale and mudstone, while the upper beds are increasingly coarse sandstone. This likely reflects the progression in the erosion of the source uplifts. The first beds exposed by uplift would include thick sequences of Lewis and Mancos Shale, providing ample clay and silt for shale and mudstone. As erosion progressed, the exposure of sandstone and basement rock would provide more coarse sediment. The accessible beds south of Cuba consist of a fairly “dirty” sandstone, of coarse, angular grains that include a fair amount of clay particles that are likely weathered feldspar grains eroded from the exposed basement rock.

Ojo Alamo
          Formation south of Cuba
Ojo Alamo Formation south of Cuba. 35 57.865N 106 59.446W

The upper beds of the Ojo Alamo are themselves resistant to erosion and form the cap of Portales Mesa, as we saw in the previous chapter. The formation is also exposed as a hogback in the Regina area in the far northwest of the Jemez region.

Ojo Alamo Formation in hogback
Ojo Alamo Formation in hogback. Looking southeast from 36 14.239N 106 55.506W

Although the period of time from 65 million to about 2.6 million years ago is still widely called the Tertiary in geologic literature, the name is no longer formally recognized. This period of time is divided into the Paleogene (65 to 23 million years ago) and the Neogene (23 to 2.58 million years ago.)

The Paleogene in the Jemez


Digital relief map of Paleogene exposures in the Jemez
      Mountains
Relief map of the Jemez with Paleogene outcroppings highlighted in red.

During the Paleogene, India collided with Asia and Antarctica moved towards the South Pole. This caused the hot, moist climate of the Cretaceous to give way to a cooler, drier climate. The rise of the Himalayas exposed great masses of silicate rock that absorbed carbon dioxide from the atmosphere, reducing greenhouse warming of the Earth. The rifting of Australia away from Antarctica allowed the Southern Ocean Current to became established around Antarctica, which also contributed to cooling. The ocean basins increased in volume and the seas receded off the continents. The world was beginning to take the form familiar today, although South America was not yet joined to North America and the Atlantic was narrower.

The Laramide Orogeny continued to affect northern New Mexico into the mid-Paleogene, raising uplifts along fault lines that often coincided with Paleozoic faulting and which remain active today. The Paleogene in New Mexico began with the crust beginning to crumple into mountains and basins as the crust was compressed; it ended with the crust beginning to tear apart down the center of the state as it was pulled back apart.

Faults

Digital relief map of major faults in the Jemez
        Mountains
Relief map of the Jemez with major faults highlighted in red.

The earth's upper crust is brittle. If the underlying ductile layers are deformed by tectonic forces, strain will build in the brittle upper crust. When this strain becomes great enough, the rocks beds of the upper crust will abruptly fracture, releasing the potential energy built up in the strained rock by shattering the rock and generating an earthquake. The rock beds will shift along the fracture to relieve the stress, leaving a zone of broken and deformed rock with intact rock beds on either side displaced from each other. This is the fault.

The map of major faults in the Jemez region shows that faults are not rare features. In addition to the major faults shown on the map, there are numerous minor faults crossing the entire area, most of which are relatively short and not particularly active. There are some patterns evident in this map. Faults in the Jemez region are more likely to run north to south than east to west. The faults also are not evenly distributed. There are particularly prominent chains of faults running along the west face of the Sierra Nacimiento (the Nacimiento Fault); along the eastern edge of the map (the Tusas-Picuris Fault); and from Black Mesa past the foothills of the the Sierra de los Valles and on south (the Pajarito and La Bajada Faults).  These are all associated with the Laramide Orogeny. There is also a swarm of north-south faults across the southern part of the map, associated with the opening of the Rio Grande Rift.

Faults are characterized by the angle of dip, which is the angle the fault plane makes with the vertical. Thus a high-angle fault, with an angle of dip close to 90 degrees, plunges nearly straight into the ground. The side beneath the fault is called the footwall and the side above the fault is called the hanging wall, from terminology used by miners (who often mine minerals deposited along a fault.)  The footwall would be under the miner's feet, while he would hang his lamp on the hanging wall, which hung over his head. The fault trace is the path where the fault intersects the surface of the earth.


Fault diagram

Almost all the faults in the Jemez area are normal faults, in which the hanging wall is displaced downwards relative to the footwall. If a normal fault is active enough, the footwall will form an escarpment above the fault trace. This will be eroded away if the fault is inactive for a long period of time.

Normal faults are characteristic of settings where tectonic forces are stretching the crust, such as in the Rio Grande Rift. Where the crust is being compressed, as it was in the Jemez area during the Laramide, one sees reverse faults where the hanging wall is displaced upwards relative to the footwall.

Normal and reverse faults

A reverse fault with a shallow dip (less than 45 degrees) is called a thrust fault. Active reverse faults of any kind are uncommon in the Jemez area today, with the Nacimiento Fault being the only major thrust fault in the area. However, reverse faults were more common during the Laramide.

Both normal and reverse faults (including thrust faults) are known as dip-slip faults. Whether normal or reverse, the side of the fault that is displaced downwards relative to the other side is described as the downthrown side of the fault. A fault can also be displaced laterally; such strike-slip faults are less common, but include such famous faults as the San Andreas Fault in California. The Embudo Fault along the south flank of Black Mesa is displaced both laterally and vertically and is described as an oblique-slip faults.

Strike-slip faults are further described as right lateral or left lateral faults. If you stand on one side of a strike-slip fault and look across at the other side, the direction of displacement of the far side determines which kind of strike-slip fault it is. Thus, if the far side is displaced to the right, it is a right lateral fault. You should be able to satisfy yourself that this is true no matter which side of the fault you are standing on, by picturing the geometry in your head.

Major faults tend not to occur as single isolated fractures. Instead, there are several parallel strands making up a fault zone. A strand can fork off the main fault and rejoin it further along the fault zone. Major faults may also end in a set of fault splays that branch off to spread the displacement over a broad area, almost like forks of a river delta.

Once a fault forms, it continues to be a zone of weakness in the crust for a very long time. If the deformation of the lower crust continues, as it typically does, there will be repeated earthquakes and repeated displacements along the fault. If new rock beds are being deposited on top of the fault, either by volcanic activity or by continued sedimentary activity, the older beds will be displaced by a greater amount than the younger beds. This permits geologists to learn something about the history of activity along the fault. The fault itself will be marked by a zone of finely crushed rock called fault gouge. If the fault ceases to be active, the fault gouge is sometimes cemented into hard rock by minerals carried by fluids moving along the old fault.

Some faults, particularly in sedimentary rock, display a phenomenon called fault drag. In the following photograph, the sedimentary beds forming the rocky outcrop, which are from a formation that actually dips to the west across the arroyo, curve over and dip steeply into a fault trace that is partially concealed by soil and vegetation just to the right of center.

Fault near Galisteo outcrops
Small fault strand east of San Miguel Mountains. 35.747753N 106.3419584W

The relatively ductile sedimentary beds have been deformed by slow motion, being (in effect) dragged along the direction of displacement. The less ductile igneous beds at right show a much smaller degree of fault drag.

One fault zone established during the Paleogene in the Jemez region is the Canones Fault Zone, located west of Abiquiu. The fault zone crosses the floor of Arroyo del Cobre and the highway to the south.

Arroyo del
          Cobre

Arroyo del Cobre. 36 14.740N 106 22.057W

The main strand of the Canones Fault Zone runs across the floor of the canyon from left to right. This fault zone is generally taken to mark the boundary between the Rio Grande Rift (on the near side) and the Colorado Plateau (on the far side), two major tectonic provinces. The rock beds on the near side of the fault zone are thrown down a considerable distance, so that the Jurassic rocks forming most of the rim we’re standing on are considerably younger than the Triassic rocks forming the far rim.  Just how great the displacement is is demonstrated by the presence of the white beds in the center of the valley floor. The main strand of the fault zone runs just beyond them, and they are the same formation (Poleo Sandstone) as the far canyon rim on the other side of the fault.

The cliffs in the left foreground are Jurassic Entrada Sandstone, with a thin covering of younger Tertiary El Rito Formation and Ritito Conglomerate. There was once a bed of Jurassic Toldilto Formation above the Entrada Sandstone, as shown by the yellowish color; this is characteristic of the uppermost Entrada Sandstone, just beneath the Todilto Formation. A few weathered remnants of the Todilto Formation lie between the Entrada Sandstone and the El Rito Formation, and on the far side of the cliffs the Todilto Formation is much better preserved.

The far canyon wall is underlain by dark red Arroyo del Agua Formation, the upper formation of the Permian Cutler Group. The lower formation, the Canon del Cobre Formation, is exposed in the valley floor and in the Canon del Cobre further north. These are capped by the tan beds of the Poleo Sandstone of the Triassic Chinle Group, with a significant discontinuity between them.

The fault trace is spectacular where it crosses a local saddle.

Canones Fault

Fault trace of main strand of Canones Fault. Looking soutwest from near 36 14.787N 106 22.471W

The fault trace is directly to the right of the yellowish outcrop of Entrada Sandstone capped with Todilto Formation. These Jurassic formations have been thrown down until they are in contact with Permian Arroyo del Agua red beds.

Canones Fault

Trace of main strand of Canones Fault Zone. 36 14.744N 106 22.573W

The zone of fine red material is the fault trace itself, which is filled with fault gouge apparently dominated by the soft red bed sediments of the Permian side of the fault. You are looking here at the closest thing to the exact boundary between two major tectonic provinces, the Rio Grande Rift province (to the left) and the Colorado Plateau (to the right).

There is a spring arising almost exactly on the fault trace (behind us as we’re facing here) that produces very brackish water. It must flow through the Todilto Formation, or some other formation rich in partially soluble minerals, at some point. Because the water is brackish, it does not support a local flourishing of vegetation; the spring is dominated by sparse salt grass instead.

In previous chapters, we visited the area just to the south where the main strand of the fault crosses the highway. You may recall that the fault has thrown down the contact between the Morrison Formation and the El Rito Formation. The Morrison Formation is of Jurassic age, about 150 million years old. Here it takes the form of fluvial sandstone, deposited in a river valley. All the rocks above the Morrison Formation were eroded away at this location, then buried again under the El Rito Formation some 50 to 30 million years ago. 

Discontinuity
        between Morrison Formation and El Rito Formation
A geological discontinuity between Jurassic and Tertiary beds. 36 13.920N 106 23.018W

The contact is clearly visible towards the top of the cliff, where the beds in the two formations lie at an angle to each other. This is a good example of what geologists call an angular unconformity. The contact is important for another reason: Whereas here the Tertiary El Rito Formation lies directly on Jurassic Morrison Formation rocks, we find that further west, across the Canones fault, the El Rito lies on top of younger sedimentary beds of Cretaceous age. This tells us that the Canones Fault existed at least thirty million years ago, and that at that time it was the west side of the fault that was thrown down. Geologists believe the Canones Fault was part of the Laramide orogeny, and has been reactivated by the opening of the Rio Grande Rift, this time with the east side being thrown down. Such fault reversal has been common across the western United States, with some of the best examples in the Precambrian basement exposed in the Grand Canyon.

The surprising conclusion is that the San Luis Valley and Espanola Basin were part of a highland area, the San Luis Uplift, during the Paleogene. The portion between Santa Fe and Los Alamos is called the Pajarito Uplift. These areas are now deep basins of the Rio Grande Rift.

This is illustrated in the following diagram.

Canones fault history

Here is a close up of the point of contact, where it comes close to the road level.

Contact between
        Morrison Formation and El Rito Formation
Contact between El Rito Formation and Morrison Formation. 36 13.920N 106 23.018W

This contact represents a gap in the geological record of at least 100 million years.

The Canones Fault Zone has been mostly inactive for the last three million years, since it does not displace volcanic rocks younger than that age.

Fission-track ages

Fission-track dating is a form of radiometric dating useful for certain minerals containing radioactive isotopes. It has mostly been applied to apatite, which contains radioactive thorium and uranium, and which is common in intrusive igneous rocks. It has special characteristics that make it useful for timing some of the events of the Laramide Orogeny.

A long-lived uranium or thorium nucleus usually decays by emitting an alpha particle. However, a few out of a trillion decays of U235 will take place through spontaneous fission, where the nucleus breaks into two halves plus a few neutrons. The two halves of the nucleus, which are still highly charged and relatively massive, penetrate a short distance into the surrounding mineral grain and leave a track of damaged crystal that can be brought out by polishing the mineral grain, then etching it with acid. Because these events are so rare, an apatite grain a few tens of millions of years old will have few enough of these fission tracks that it is possible to make out individual tracks under a microscope, count their density, and determine how long the grain has been accumulating radioactive damage.

Fission tracks are erased by relatively mild heating. Apatite has a particularly low closure temperature of about 70 to 110 C (160 to 230 F) and it is particularly rich in heavy radioisotopes. This means that the fission track age for a rock outcropping will be the age since the rock last experienced any kind of heating event.

Geologist Shari Kelley and her collaborators have collected fission track ages for many of the mountains around the Rio Grande Rift. The highest peaks of the Sangre de Cristo and San Pedro Mountains have fission track ages of around 65 million years, showing that these rocks were brought close to the surface by the Laramide Orogeny no earlier than the start of the Cenozoic. Locations further down the flanks of the mountains show progressively younger fission track ages, of about 35 million years at the feet of the mountains. This indicates continual uplift from 65 million to 35 million years ago, and it is possible to estimate the uplift rate at 0.072 mm per year (a little over a quarter of an inch per century.) Sandia Crest is relatively young, with fission track ages of around 20 million years or less. No fission track ages have been published yet for the Tusas Mountains, but the dating work is continuing.

We'll see stratigraphic evidence later in this chapter that is broadly consistent with these timings.

The Paleocene Epoch

Digital relief map of Paleocene exposures in the Jemez
        Mountains
Relief map of the Jemez with Paleocene outcroppings highlighted in red.

65 milllion years ago, northern New Mexico was beginning to look something like it does today. The Western Interior Seaway was no more. The San Juan Basin to the west was accumulating sediments eroded off the newly uplifted San Luis and Pajarito Uplifts to the east. The climate was hot and moist, as it was across almost the entire globe. Far to the west, Europe was just beginning to separate from Greenland and Greenland from North America, and there was an intermittent land bridge between Asia and North America, which would continue to open and close until almost the present day. South America had pulled away from Africa but was still separated from North America by a tropical sea, while Africa was separated from Europe by the Tethys Ocean. India was still well south of Asia but moving north. Australia was just beginning to pull away from Antarctica, and its bulk diverted southern ocean currents to ensure that warmth from the tropics was well mixed across the entire ocean area of the Earth.

This earliest epoch of the Paleogene has been named the Paleocene ("old recent"), and lasted from about 65 million years ago to 56 million years ago. This epoch left little trace in the Jemez proper, but Paleocene beds of the Nacimiento Formation are found west of the Sierra Nacimiento in the San Juan Basin.

The Nacimiento Formation

Digital relief map of Nacimiento Formation exposures in
        the Jemez Mountains
Relief map of the Jemez with Nacimiento Formation outcroppings highlighted in red.

During the early Paleogene, the San Juan Basin west of the Sierra Nacimiento was a lush, well-watered intermontane basin, and fossils of recognizable families of modern plants typical of semitropical climates are found in the Nacimiento Formation, along with numerous mammals. These included examples of the first true rodents and primates, which branched off from a common ancestor in the Cretaceous, 80 million years ago. That's right: Of all the common household pets, rats, mice, hamsters, and gerbils are our closest kin. The dog, man's best friend, is man's cousin 100 million years removed.

There are excellent exposures of the Nacimiento Formation throughout this area, generally forming slopes under cap rock of the San Jose Formation. Mesa de Cuba is an example.

Mesa de Cuba
Mesa de Cuba. Looking northwest from 35 58.057N 106 59.514W

The mesa is underlain by San Jose Formation sandstones and conglomerates forming a cap over softer shales of the Nacimiento Formation.

This sequence may be examined more closely along the main highway just northwest of Cuba.

San Jose and
        Nacimiento Formations
Nacimiento Formation and San Jose Formation northwest of Cuba. Looking north from 36 02.076N 106 58.076W

The San Jose Formation is the sandstone cliffs, while the slope-forming sedimentary beds beneath the cliffs are Nacimiento Formation.

The Paleocene marks the end of the long period of tectonic quiescence in the Jemez, and with it the paleotopographic map again becomes interesting.

Digital relief map of Paleocene paleotopography in the
        Jemez Mountains
Paleocene paleotopography map

Only the San Juan Basin to the northwest was accumulating sediments during the Paleocene. The remainder of the Jemez region was high ground. However, this map should be taken with some skepticism. The beds identified as Eocene beds resting directly on Cretaceous beds have unusually poor age control, and may be older than they appear.

The Eocene Epoch

The Hayden Expedition of 1869 was not exclusively concerned with locating coal supplies for a transcontinental railroad. Hayden was a well-trained and careful geologist and his observations established the first basic framework for the stratigraphy of the Santa Fe area. He took note of the sandstone beds of the Galisteo area, which he found dipped at angles from 5 to 50 degrees and displayed a wonderful variety of colors from almost white to a deep brick red. He was disappointed to find no fossils in the beds other than enormous petrified logs, including one remarkably well preserved log with a cavity 10" (25cm) in diameter running through it. He named these beds the Galisteo sand group, and the name has remained with some of the beds to the present day.

Digital relief map of Eocene exposures in the Jemez
        Mountains
Relief map of the Jemez with Eocene outcroppings highlighted in red.

The Eocene, which lasted from about 56 million years ago to 38 million years ago, was the warmest epoch of the Cenozoic Era. Ocean temperatures were at least 5 degrees Centigrade (9 degrees Fahrenheit) warmer than today. However, global temperatures began to decline in the middle of the epoch and plunged dramatically towards its end.

The start of the Eocene in the Jemez was marked by uplift of the Sierra Nacimiento to the west and further uplift of the Brazos-Sangre de Cristo geanticline to the north and east. A broad river valley lay between the Sierra Nacimiento and the Pajarito Uplift, where the Jemez Mountains are located today.

The tectonic compression of the Laramide Orogeny came to a close towards the end of the Eocene, and the uplifts began to erode down to fill the surrounding basins with sediments. The remnants of the Farallon Plate began to sink into the mantle, and hot aesthenosphere rose to take their place. This caused volcanism to break out in the Yellowstone area, and crustal extension began to take the place of compression, beginning in what is now western Montana.

The San Jose Formation

During the period between 1877 and 1892, the two leading paleontologists in the United States, Edward Drinker Cope of the Academy of Natural Sciences in Philadelphia and Othniel Charles Marsh of the Peabody Museum of Natural History at Yale, engaged in a bitter rivalry that become known as the Bone Wars. Cope had first come out to the fossil beds of the West as an unpaid member of the Hayden Survey, and he soon developed a reputation for ignoring instructions from Hayden and collecting fossils where he liked.

In late 1874, Cope ignored orders from Wheeler to proceed north from the Chama River to Colorado, and headed south to explore the area north of the San Pedro Mountains, where fossils had been reported. North of Regina, he found early Eocene fossils in the badlands and spent about seven weeks collecting fishes, reptiles, and what was then the oldest known mammal fossils in North America. He also collected the fossil of a giant flightless bird. This proved to be one of the most important fossil finds for life in the early Eocene in North America, which Cope wrote to his father was "the most important find in geology I have ever made". However, Cope was unable to find anything but a few bits of petrified wood in the underlying beds of the Nacimiento Formation, which he named the Puerco Marls.

Digital relief map of San Jose Formation exposures in
        the Jemez Mountains
Relief map of the Jemez with San Jose Formation outcroppings highlighted in red.

The San Jose Formation is separated by a definite discontinuity from the underlying Paleocene Nacimiento Formation. It dominates the area northwest of Cuba, where it tops mesas and other hills. It can also be a beautiful formation:

San Jose formation
San Jose Formation. 36 09.473N 106 58.259W

Here is a gorgeous exposure in a road cut. This exposure is heavily cross-bedded and strongly indurated.


Cross beds of San Jose
        Formation
San Jose Formation. 36 04.289N 106 59.026W

Sample of San
          Jose Formation

A fairly coarse, angular, and feldspar-rich sandstone, with some organic debris. Other exposures are quite dirty. In general, this formation is a relatively immature sandstone near Cuba, reflecting its origin as sediments off of the nearby Sierra Nacimiento and Brazos-Sangre de Cristo geanticline.

Locations near Cuba have fossils of the giant carnivorous bird, Diatryma. And if "giant carnivorous bird" seems indistinguishable from "dinosaur", I doubt many paleontologists would argue with you.

The formation shows internal discontinuities that suggest the lowest beds formed in the subsiding San Juan Basin, but later beds formed during discrete intervals of folding along the Nacimiento Fault that produced accommodation space for sediments. In other words, episodes of folding dropped the rock beds to the west so that there was room for sediments to accumulate on them.

The San Juan Basin west of the Sierra Nacimiento dips very gently to the north. As a result, successive resistant beds form nearly level mesa tops, with the youngest beds exposed on mesa tops to the north and the oldest to the south. This panorama shows the sequence of beds.

Eastern San Juan Basin panorama
Panorama of the eastern San Juan Basin. Looking west to southwest from 36 04.289N 106 59.026W


At left in the far distance is the Mount Taylor volcanic field. The northeastern part of the field includes several volcanic plugs, of which Cabezon Peak is the largest.

Just visible in front of the plugs is San Luis Mesa, which is topped by Lookout Point sandstone. The mesas at far left, which are labeled La Ventana Mesa North and La Ventana Mesa South on the geologic map, are topped by the next resistant formation, the Cliff House Formation (La Ventana Member) underlain by less resistant Menefee Formation.

The next four formations, the Lewis Shale, Cliff House Sandstone, and Fruitland and Kirtland Formations, are less resistant, while the Ojo Alamo Formation forms the resistant cap of Mesa Portales near the center of the panorama.

At right is Cuba Mesa, capped with resistant San Jose Sandstone over less resistant Nacimiento Formation.

In the middle distance in the center and right of the panorama is a hogback ridge formed by steeply dipping beds of the Mesaverde Group (Lookout Point, Menefee, and Cliff House Formations) that were exposed by the uplift of the Sierra Nacimiento. The flat ground to their east is underlain by Mancos Formation (here poorly exposed) while everything in the foreground is Dakota Formation or older.

El Rito Formation

Digital relief map of early Eocene exposures in the
        Jemez Mountains
Relief map of the Jemez with El Rito, Gallisteo, and Diamond Tail Formation outcroppings highlighted in green, red, and yellow, respectively.

The El Rito Formation extends from north of the Valles caldera to the Chama area in northernmost New Mexico. It is somewhat difficult to precisely date, having very few fossils, but is probably in the ballpark of 40 million years old. The formation is mostly interbedded red mudstone and sandstone, tending to form slopes.

The lowest bed of the El Rito Formation is very coarse conglomerate with ample sandy matrix, as shown in the earlier photograph of the contact between the El Rito Formation and the Morrison Formation near the Canones Fault. This is even more evident in the Red Wash Canyon area.

Basal
          conglomerate of El Rito Formation
Basal conglomerate of El Rito Formation. 36 14.428N 106 22.484W5

The clasts are mostly quartzite, of the type found in the Tusas Mountains to the north. This suggests that there was already a Tusas Range 40 million years ago, part of the San Luis Uplift, and geologists interpret the El Rito Formation as alluvial fans eroded off the early Tusas Mountains to the south and west. The lack of fossils and other characteristics of the formation suggest it was deposited under semiarid conditions.

Here is a sample of the El Rito Formation from its lowest bed.

El Rito Formation
El Rito Formation. 36 13.920N 106 23.018W

Notice that there is a quartzite pebble embedded in this sample. This was taken from near the bottom of the bed. The sandstone consists of somewhat rounded but poorly sorted clasts of quartz with considerable feldspar and lithic clasts, and much cement in the pore spaces — all consistent with an immature sandstone formed close to its source rock. Because of the proximity of the Canones Fault, which forms a path for ground water, this exposure is unusually well cemented.

El Rito Formation forms the south rim of Arroyo del Cobre.

El
          Rito Formation on south rim of Arroyo del Cobre

El Rito Formation on south rim of Arroyo del Cobre. 36 14.754N 106 22.123W

In this area, the El Rito Formation rests unconformably on Jurassic beds.

El Rito resting on Entrada

El Rito Formation resting on Entrada Formation. 36 14.754N 106 22.123W

This photograph nicely contrasts a conformable contact with an unconformable contact. At left, Todilto Formation limestone beds lie atop massive yellow beds of Entrada Formation standstone. The two formations are bedded at the same angle. At top center and right, red beds of the El Rito Formation lie atop the same massive beds of Entrada Formation, but the beds lie at a different angle with respect to each other. The contact at right represents an erosional surface across which around 130 million years of geologic history are missing.

There are also some good exposures north of the village of Canones.

El
          Rito Formation
El Rito Formation. Looking north from 36 11.271N 106 26.738W

A particularly accessible exposure is found in a road cut along the old road into Canones.

El Rito
          Formation
El Rito Formation. 36 11.463N 106 26.989W

Towards the top, the El Rito Formation gives way to the light-colored Ritito Conglomerate, which we'll describe shortly. Here's a sample of the El Rito Formation from this location.

El Rito
          Formation
El Rito Formation. 36 11.463N 106 26.989W

This sample is a medium sandstone of well-rounded but poorly-sorted quartz grains with some feldspar and a few lithics and a fair amount of reddish cement in the pore spaces. Note the distinctive muscovite flakes, generally aligned with the bedding planes of the sandstone. Mica is not usually abundant in sandstones and so serves as a distinguishing feature of the El Rito Formation. Its presence hints at an origin in the mica-rich pegmatites of the Tusas.

The El Rito Formation is not typically well exposed in the Jemez proper, but along Forest Road 100 in the north Jemez, the El Rito Formation is hinted at by reddish discoloration in road cuts just below outcrops of the younger Ritito Conglomerate. One can sometimes find clasts of El Rito sandstone resembling those from the Canones Area.

El Rito Formation
El Rito Formation. 36 08.616N 106 31.070W

Similar poor exposures continue west and south as far as upper Coyote Canyon. These exposures are significant because they are located northwest of the Canones fault zone, and suggest that this had not yet been reactivated by El Rito time.

Diamond Tail and Galisteo Formations

Another pair of formations in the Jemez area from about this same time are the Diamond Tail and Galisteo Formations. These crop out in only a few locations between the northern Sandia Mountains and the southern Sangre de Cristo Mountains, because they are mostly buried under younger sediments. However, there is an outcropping of Galisteo Formation along I-25 where the highway descends from the Cerros del Rio, at La Bajada. The outcropping shows up clearly on Google Maps as a red patch around the highway and to the northeast.

Gallisteo Formation on
        I-25
Galisteo Formation. 35 31.031N 106 13.552W

The resemblance of the red beds of the Galisteo Formation to the Abo Formation is not accidental. Like the Abo, the Galisteo formed in an arid fluvial environment, where rivers descending from the Pajarito Uplift and Sierra Nacimiento meandered south across a level valley. Similar environments produce similar rocks.

Galisteo Formation is prominent at Garden of the Gods.

Galisteo Formation at Garden of hte Gods

Garden of the Gods. 35 26.505N 106 04.927W

Here the beds have been tilted nearly vertical by the nearby intrusion of the Cerrillos Hills. We'll have more to say about the Cerrillos Hills shortly.

The most complete section of the Galisteo Formation is found in the Hagen Basin, located south and east of I-25 and north of the Sandia Mountains. The Hagen Basin is a large block of crust bounded by the La Bajada Fault to the east, the Budaghers Fault to the north, the San Francisco Fault to the west, and to the south by the South Montezuma Thrust Fault and zone of severe deformation to its east along the northern edge of the Sandia Mountains. This block of crust was uplifted to the west, tilting its beds, and erosion subsequently exposed a splendid sequence of formations dating from the Jurassic in the west to the Santa Fe Group to the east, with the particularly resistant beds of Espinaso Ridge forming a hogback. The Galisteo Formation is exposed just west of Espinaso Ridge.

Galisteo Formation at Espinaso Ridge

West of Espinaso Ridge. 35.3814564N 106.3191287W

Galisteo Formation underlies everything out to the middle ground. We see that it is a sequence of white, gray, and brick red sandstone and siltstone beds.

The formation also crops out on the east slopes of the San Miguel Mountains in the southeast Jemez, where it has been exposed by deep faulting along the Pajarito Fault Zone. 

Gallisteo
          Formation east of St. Peter's Dome
Galisteo Formation east of St. Peter's Dome. Looking west from 35 45.441N 106 22.135W

The block west of the fault zone has been tilted to the west. The tilting is discernible even in the relatively young volcanic beds atop St. Peter's Dome, but is greatest for the oldest beds, which show a dip of around 27 degrees. The increase in dip with age shows that the block has been slowly tilting over the last 40 million years. This tilted block forms a striking island of old rock among the much younger rocks of the Rio Grande Rift and Jemez volcanic field.

Gallisteo Formation east of
          St. Peter's Dome
Galisteo Formation east of St. Peter's Dome. 35.748755N 106.3475267W

The beds here vary from deep red to yellowish-white in color, with the lower beds being the deeper red.

Gallisteo Formation
          east of St. Peter's Dome
Galisteo Formation east of St. Peter's Dome. 35.749173N 106.3479167W

The red beds are a fairly fine-grained sandstone, though not well sorted and with abundant matrix and feldspar grains. This is typical of an immature sandstone.

Gallisteo Formation east of St.
          Peter's Dome
Galisteo Formation red sandstone. 35.7485094N 106.3420798W

The higher beds include some coarser conglomerate beds.

Gallisteo Formation east
          of St. Peter's Dome
Galisteo Formation conglomerate east of St. Peter's Dome. 35.749173N 106.3479167W

The clasts are well-rounded pebbles, mostly Precambrian quartzite but with some volcanic rock, in a white, very well-cemented, matrix. There is sufficient matrix that the pebbles barely touch, putting this on the boundary between a matrix-supported and a clast-supported conglomerate.  Coarse sediments like these are more common in these northernmost Galisteo Formation beds than than in the exposures further south. This supports the conclusion that the sediments came from highlands to the north, with the coarser sediments found closer to the source rock and the finer sediments found further away.

This bed is so thoroughly cemented that it resembles a lithic tuff, of which we'll learn more in the next few chapters. However, the rounded form and character of the clasts show that it is a conglomerate, though the abundant cement may have come from overlying volcanic ash beds.

The northernmost exposures of Galisteo Formation include outcrops in Capulin Canyon.

Gallisteo Formation in Capulin
          Canyon.
Galisteo Formation in Capulin Canyon. Looking northeast from 35.754263N 106.3360017W

The outcrops are the reddish beds at the base of the mesa. Here the main strand of the Pajarito Fault Zone throws down the area to the east, with the Galisteo beds marking the limit of pre-rift rock on the west side of the rift.

The Galisteo Formation was originally defined to include older beds that are extensively exposed south of the villages of Cerrillos and Galisteo and northeast of the ghost town of Hagen. Geologists later recognized that there was a significant discontinuity separating the two sets of beds, and the lower beds were split off into their own formation, the Diamond Tail Formation. This formation is exposed mostly at the southern limits of the Jemez region.

Diamond Tail Formation
Diamond Tail Formation south of Los Cerrillos. 35 24.865N 106 08.435W

As with the Permian Cutler Group and Abo Formations, the question arises whether the El Rito Formation and the Galisteo Formation are really distinct formations. They appear to be the same age and are similar in their characteristics. It is likely that they were deposited on opposite ends of a river valley running south between the Brazos-Sangre de Cristo geanticline and the Sierra Nacimiento whose middle section is now buried under the volcanic rocks of the Jemez. They also appear to correlate with the San Jose Formation west of the Sierra Nacimiento.

The Galisteo and Diamond Tail Formations contains numerous petrified logs in the Cerrillos area, mostly conifers, oak, and beech. Fossil pollen includes palm pollen, suggesting a subtropical forest giving way to a mixed pine-oak forest at higher elevations. The elevation of the Cerrillos area was probably around 550 m (1800') with a climate that was frost-free throughout the year. Cerrillos presently has an elevation near 1800 m (6000'), which is an indication of how much regional uplift has taken place in the last 50 million years.

The timing and cause of this uplift remain somewhat controversial. There are reasonably good time constraints for the integration of the modern Colorado River, suggesting it took its present form between 6.0 and 4.4 million years ago. This suggests relatively recent uplift.  The cause is also imprecisely known, though it seems likely it is related in some way to the approach of North America to the East Pacific Rise. One hypothesis is that the long history of subsidence of the Colorado Plateau was the result of slow conversion of garnet granulite in the underlying mantle to denser eclogite, and the recent uplift is the result of rapid conversion of eclogite back to lighter garnet granulite. There is some evidence supporting this hypothesis from xenoliths erupted across the Colorado Plateau, which change from eclogite to granulite around 25 million years ago.

The sediments of the Diamond Tail Formation appear to have come mostly from the northeast, where the Brazos-Sangre de Cristo geanticline provided a rich source of sediments. The Sierra Nacimiento became an additional source of sediments in Galisteo time, supporting the conclusion that the Sierra Nacimiento was not uplifted until the Eocene. There appears to be no contribution from the south, in the area of the present-day Sandia Mountains, suggesting that these were not uplifted until after the Eocene.

Fossils are not abundant in the Galisteo and Diamond Tail Formation. Curiously, the lowest beds contain Paleozoic brachiopods and Cretaceous shark teeth, eroded from underlying formations and redeposited as reworked fossils. You can imagine how these might have thrown off initial dating of the beds! However, two locations in these formations have some beds containing bone fragments, mostly mammals. A site in the Diamond Tail Formation in the Cerrillos Hills produces fossils belonging to the Wasatchean Stage, which extended from about 55.4 to 50.3 million years ago. These include freshwater fish scales, fragments of turtle shells and vertebra, tapir teeth, Coryphodon teeth, a Hyopsodus jaw, and Hyracotherium teeth. Coryphodon was one of the earliest really large grazing mammals, roughly the size of a horse; Hyposodus was a primitive mammal resembling a shrew but more closely related to modern horses; and Hyracotherium was a dog-sized ancestor of both horses and rhinoceroses.

The other fossil site is in the Galisteo Formation at Arroyo del Tuerto, and contains numerous fossil remains of titanotheres, very large mammals related to horses and rhinos. These date to the Duchesnean Stage, about 42 to 38 million years ago. The Stearns quarry where the mammal fossils were found is a protected area, but nearby are impressive exposures of petrified wood.

Petrified wood
Petrified wood in uppermost Galisteo Formation beds. 35.3812714N 106.3181237W

Towards the end of the Eocene, volcanic activity picked up in what is now the San Juan Mountains of Colorado. Ash and basalt flows from the San Juan field reached into northern New Mexico, though not as far as the Jemez region. This activity would peak in the Oligocene with the formation of at least 15 calderas, and the activity would not remain confined to southern Colorado.

The Oligocene Epoch

34 million years ago, northern New Mexico was rocked by a series of volcanic eruptions along a belt extending south from the Pajarito Uplift. Magma from the upper mantle had forced its way to near the surface, doming up the overlying sedimentary beds almost like blisters in the Earth's skin. In some places the magma broke through, forming volcanoes that deposited debris in a broad apron around the volcanic belt and choked the river valleys that had formerly run south through the area. This marked the beginning of the Oligocene Epoch in New Mexico.

The Oligocene Epoch lasted from 34 to 23 million years ago. It was a period of severe global cooling, caused by the rifting of Australia and South America away from Antarctica. This created the Southern Ocean and established the Antarctic Circumpolar Current, which isolated Antarctica from warmer ocean water to the north and turned the continent into an icebox. The earth has not returned to pre-Oligocene temperatures, leading some geologists to regard the Oligocene as the start of a Cenezoic Ice Age of which the Quaternary Ice Age (which we'll discuss later on) is only the latest episode. The drop in sea level accompanying the global cooling sharply increased continental weathering rates, with the 87Sr/86Sr of marine carbonates beginning a steep rise that continued almost to the present day.

The Oligocene was a period of volcanic upheaval in much of New Mexico. There were some truly spectacular eruptions in the area west of Socorro, along the southeast margin of the Colorado Plateau. These are distant enough from the Jemez that they will not come into our story further. Oligocene volcanism also included the Ortiz Mountains southeast of the Jemez. The Ortiz Mountains are part of the Ortiz Porphyritic Belt, a chain of volcanoes extending from South Mountain north to La Cienega. The bulk of the eruption was of monzonite and latite, which are intrusive and extrusive rocks formed from alkaline magma that is moderately rich in silica. This particular composition yields a rock composed mostly of plagioclase and alkaline feldspar, with traces of iron minerals and small amounts of quartz.

It is not clear how much of the eruption actually reached the surface. The igneous rocks in this area are primarily intrusive monzonite, coarsely crystallized and interbedded with older sedimentary rocks. Most of the extrusive latite has long since eroded away. However, there are beds of volcanic debris and ash atop the Galisteo Formation along Espinaso Ridge and west of Galisteo, named (logically enough) the Espinaso Formation. These are though to have come from the pulse of volcanism that produced the Ortiz Mountains.

It seems likely that the Ortiz Porphyritic Belt was a substantial volcanic field at its peak, comparable in size with the Jemez volcanic field we see today. It formed just as the Rio Grande Rift began to open, but its connection with the rift is unclear. However, the earliest stages of Ortiz volcanism, when laccoliths formed, was probably still a time of compression. Laccoliths are bodies of magma that pushed up the overlying rocks but did not quite break through to the surface, so that the magma cooled underground into visibly crystalline rock. Regional compression of the crust favors formation of laccoliths. The later stages, when dikes became more common, marks the transition to extension. The volcanic belt has since been deeply eroded, exposing the plumbing below the volcanoes.

As the eruptions died down, the crystallizing monzonite sweated fluids rich in incompatible elements, which are chemical elements that do not easily fit into the crystal structure of the common silicate minerals that form from magma. Gold and tungsten are incompatible elements, and both were mined in the Ortiz Mountains, with the last gold mine closing in 1986.

Ortiz Monzonite

Hayden visited the Ortiz Mine during his 1867 expedition and noted that both the Ortiz Mountains (which he knew as the Placiere Mountains) and the Cerrillos Hills to their north were composed of masses of intrusive rock and associated dikes. He naturally (but incorrectly) assumed that the dikes were the source of the nearby Cerros del Rio basalt and noted that one dike extended for a great distance east of the Cerillos Hills. This is El Crestón, which forms a striking landmark in the otherwise flat country surrounding the town of Galisteo.

Digital relief map of Ortiz Belt and Cieneguilla
        Basanite exposures in the Jemez Mountains
Relief map of the Jemez with Ortiz Belt and Cienegujilla basanite outcroppings highlighted in red and green respectively.

Southwest of Santa Fe, near where Cienega Creek joins the Santa Fe River, lies the village of La Cienega. In the language of the Spanish Southwest, a cienega is a wetlands in an otherwise arid region, characterized by permanent saturation of the soil by alkaline water. During the early Spanish period, the cienega southwest of Santa Fe supplied water to nearby El Rancho de las Golondrinas, "Ranch of the Swallows", which is preserved today as a living museum. El Rancho del las Golondrinas was the final paraje, rest stop, on the Camino Real, the Royal Road, joining Mexico City to Santa Fe. Here travelers could replenish supplies and rest for the night. The relatively abundant water and fertile soil made La Cienega the breadbasket of old Santa Fe.

La Cienega has escaped most of the gentrification that has affected many historic villages in the Santa Fe area. Its population remains small at about 3000 persons, and 70% of these identify as Hispanic. The village is set in a shallow valley surrounded by red exposures of Galisteo Formation, overlain in places by young sediments of the Ancha Formation that we'll examine later in the book.

Of immediate interest is the terrain north of La Cienega, which is low hills underlain with Ortiz monzonite and Espinaso brecciated latite. Ortiz monzonite is exposed south of the gravel road to Cañon.

Ortiz monzonite?


Ortiz
          monzonite

Ortiz Monzonite. 35.577627N 106.1325749W

Though fine grained, this rock is visibly crystalline under the loupe. In a few spots, the rock is quite coarsely crystalline and its makeup is clear.

Ortiz monzonite

Ortiz Monzonite. 35.5783697N 106.132627W

The rock is a mixture of translucent dark microcline, with obvious cleavage faces that are resistant to scratching by a metal probe (which rules out the remote chance that it is calcite), and opaque white plagioclase. Dark grains of augite with a few flakes of biotite are also present. The mixture of microcline and plagioclase without noticeable quartz is by definition a monzonite.

There is some variability in the Ortiz monzonite even in the La Cienega area, suggesting it was emplaced as multiple intrusions. Here's a sample from a few hundred yards south.

  Monzonite
Ortiz monzonite. 35 34.591N 106 07.893W

This sample is slightly lavender on fresh surfaces under sunlight, with a beautiful pattern of interlocking crystals. Again, these appear to be two different varieties of feldspar. There are also some large rusty grains that are probably weathered augite.

Still further south are rounded outcrops of yet another variety of monzonite.

Monzonite outcrop

Monzonite outcrop. 35 34.337N 106 07.969W

I've placed my walking stick for scale on the photograph below.

Foliated rocks
Monzonite outcrop. 35 34.325N 106 08.062W

The stuff weathers a bit like granite, unsurprising considering that it's mostly feldspar and it's mostly feldspar that makes granite weather the way granite weathers. Here’s a sample.

Monzonite
Monzonite. 35 34.325N 106 08.062W

This rock is much lighter in color and gives the impression of being porous, almost like a sandstone. However, under the loupe, one sees that it consists almost entirely of small feldspar grains with numerous small flecks of biotite and hornblende. No quartz is visible.

The unusual character of this rock may indicate that it has been weathered or hydrothermally altered. The rock comes from a ridge that looks suspiciously like a dike or fault trace, and an actual dike crops out along the road to Canon that is roughly aligned with this ridge.

This outcropping is located in the middle of La Cienega, is not mapped, but could be either Ortiz monzonite or Espinaso Formation.

Ortiz monzonite?

Ortiz Monzonite? 35.562N 106.126W

Sample of Ortiz monzonite?

Under the geologist’s loupe, one sees smaller gray crystals, probably alkali feldspar, with larger white crystals, probably plagioclase, and lots of oxidized iron that was probably once hornblende or another iron-rich mineral. The presence of both alkali and plagioclase feldspar is characteristic of monzonite, and the absence of visible quartz keeps this from being classified as a quartz monzonite.

The Ortiz porphyry belt continue south through the Cerrillos Hills and the Ortiz Mountains.

Cerillos
          Hills

Cerrillos Hills. Looking north from 35 25.104N 106 07.828W

Like most of the Ortiz Porphyry Belt, the Cerrillos Hills are a cluster of laccoliths. The highest point is Cerro Bonanza, also known as Santa Rosa Mountain, 7088′ tall. It is underlain by augite biotite monzonite, so called because it contains some biotite and lesser amounts of augite.

It's a popular belief that you can rid your body of toxins by making yourself sweat. This turns out not to be the case, unless you mean the modest quantities of urea and salt found in human sweat. But bodies of magma cooling underground really do “sweat” certain trace elements, dissolved in hydrothermal fluids, which penetrate the surrounding rock to deposit valuable metal ores. The second phase of Ortiz volcanism, when the crust had transitioned from compression to extension and numerous dikes formed, was also the phase of greatest mineralization. The dikes formed natural pathways for circulation of hydrothermal fluids.

The Ortiz porphyry belt has numerous old mines and mining claims, and gold was being mined here as recently as the 1980s.  But the mineral that has been mined the longest in the Cerrillos Hills is turquoise.

Turquoise
Turquoise from Blue Bell Mine. Near 35 28.219N 106 6.751W

Turquoise is a form of  hydrated copper aluminum phosphate, CuAl6(PO4)4(OH)8·4H2O, and it is the gemstone most closely associated with native American jewelry of the Southwest. It is not found in the original hydrothermal ore deposits of the Cerrillos Hills, but is a secondary or supergene mineral formed by weathering of the underlying ore veins. Meteoric water (rainwater or melted snow) carrying dissolved oxygen infiltrates the ore veins, where it oxidizes the copper sulfides to more soluble copper sulfates. The acidic, metal-bearing water then reacts with aluminum- and phosphorus-bearing minerals (such as feldspars and apatite) to deposit turquoise in veins and joints in the overlying rock.

Turquoise has been mined in the Cerrillos Hills since at least 700 A.D., the date of an archaeological site near Mexico City where turquoise has been found that has the chemical fingerprint of Cerrillos Hills turquoise. Mount Chalchihuitl in the Cerrillos Hills was the largest open pit mine of any kind in North American when the Spanish arrived in 1535. The native Americans may have mined the deposit by building large bonfires over the ore beds, then dousing the fires with cold water to fracture the rock by heat stress. The rest was done with stone tools. The Spanish subsequently mined small quantities of silver, but the big boom was between 1879 and and 1885 or so.

The boom was fueled mostly by speculators, with just enough mineral resources in the claims to keep the speculation going. Many investors bought rights to claims with less thought of profiting from the actual mineral extraction than from reselling the rights to someone else at a profit. This was not at the level of the famous tulip mania in Holland, but the same speculative mindset was involved.

The boom, and its collapse, were influenced by the laws for staking and developing mining claims. The laws governing such claims were derived from the economic theories of the Scottish Enlightenment, which strongly influenced the American founders. Under these theories, a person rightfully established ownership of previously unclaimed land by mixing his labor with the land. This could be done by clearing and tilling it for agriculture or excavating it for valuable mineral resources. This philosophy was reflected in the Homestead Act, which allowed settlers to take possession of public land they successfully cultivated for long enough, and in mining claim law. In most American jurisdictions, you could file a claim at the nearest county seat asserting that there were valuable mineral ores on a section of public land. You then marked your claim with "monuments", such as actual wooden or metal stakes, inscribed with the number of your claim. You then had exclusive rights to develop the claim and mine the ores. For the claim to remain valid, you had to do a minimum amount of development within a prescribed time frame. Eventually, if you developed the claim sufficiently, you could purchase title to the land from the government at a bargain price per acre. This was known as patenting the claim.

Mine claims were governed by state and local law previous to the Mining Law of 1872, which formally legalized mine claims on federal land. The law was enacted by Congress under the encouragement of the western states and against the objections of some Easterners who regarded miners as squatters stealing national resources. The Mining Law of 1872 was amended in 1920 to remove oil, gas, and coal from the claim process, and again in 1947 to allow sand, stone, gravel, and pumice to be mined without a claim. Subsequent amendments have added remediation requirements, but the basic provisions of the law remained in effect until 1 October 1994, when Congress ended the filing of new claims through the expedient of refusing appropriations to process the paperwork.

The claim boom in the Cerrillos Hills was fueled in part by relatively lax development requirements. You had a full year to dig just ten feet of tunnel. When this changed to a more stringent requirement — three months to dig ten feet of tunnel — the boom collapsed. The speculators weren’t interested in actually digging; they wanted to try to make money by trading the mineral rights.

This does not mean that there were not actually any valuable minerals in the Cerrillos Hills. In addition to turquoise, the Spanish mined considerable silver and lead, and iron and manganese were also produced in significant quantities. Manganese is still used in steel manufacture today. It is a powerful deoxidizer and desulfuring agent and is almost indispensable for making high-quality steel from low-grade ore

Picture Rock, also known as Lone Bluff, is a particularly picturesque isolated laccolith of the Cerrillos Hills.

Picture Rock

Picture Rock, Looking east from 35 29.597N 106 05.523W

The monzonite is readily accessible next to the road as it passes Picture Rock.

Picture Rock monzonite

Ortiz Monzonite near Picture Rock. 35 29.604N 106 04.855W

Notice the resemblance to the outcrop of monzonite at La Cienega. Another local landmark is The Devil's Throne, just northeast of Los Cerrillos.

The Devil's
          Throne

The Devil's Throne. Looking northwest from 35 26.479N 106 08.011W

Back
          side of the Devil's Throne

Back side of he Devil's Throne. 35 26.723N 106 08.084W

The geologic map for this area duly notes that one team of investigators classified the rock making up the Devil's Throne as hornblende quartz latite porphry and another as hornblende monzonite porphyry. The difference is subtle: Quartz latite is fine-grained rock with slightly more quartz than coarse-grained monzonite. A good petrological laboratory could settle the issue, but then it’s entirely possible the composition varies slightly from location to location. The geologic map chooses to map most of the rock underlying the Cerrillos Hills simply as porphyritic andesite. Here's a sample from further west, along the Jane Calvin Sanchez Trail at Cerrillos Hills State Park.

Porphyritic
          andesite

Porphyritic andesite along the Jane Calvin Sanchez Trail.  35 26.910N 106 07.151W.

The porphyritic andesite of the Cerrillos Hills is itself intruded by dikes that must have formed shortly after the main body of andesite cooled. These radiate away from the hills for great distances in all directions, Many are composed of a distinctive trachytic latite porphyry -- a rock moderate in silica and rich in potassium that contains large feldspar phenocrysts. One such dike is located along the Jane Calvin Sanchez Trail.

Dike

Trachytic latite porphyry dike along the Jane Calvin Sanchez Trail.  35 26.995N 106 07.199W

Close view of trachytic latite porphyry

Trachytic latite porphyry dike along the Jane Calvin Sanchez Trail.  35 26.995N 106 07.199W

A dike composed of similar rock extends well west of the Cerrillos Hills.

Rosario dike

Trachytic latite porphyry dike near Rosario.  35 27.678N 106 13.781W

Rosario
          dike

Trachytic latite porphyry dike near Rosario.  35 27.678N 106 13.781W

The rock here is reddish rather than gray in color, but such color differences are often among the least useful for distinguishing rocks.

Not all the dikes radiating from the Ortiz porphyry belts are porphyritic. Some are composed of what appears to be a uniform fine-grained basalt. One example is the El Creston dike, which forms the core of an impressive ridge extending for miles east of the Cerrillos Hills.

El Creston

El Creston dike. 35 24.539N 105 56.910W

This suggests two systems of dikes from two eruptive pulses in the Cerrillos Hills. A third system of basalt dikes located west of the Cerrillos Hills running in a north-south direction are likely younger dikes associated with the opening of the Rio Grande RIft.

A similar dike west of the Cerrillos Hills shows hornfelsization.

Dike
          with hornfels near Waldo

Dike with hornfelsization evident. 35 24.539N 105 56.910W

The dike is the dark gray rock with redish patina on the right. The whitish zone immediately to the left of the dike is hornfels, produced by the heat of the magma baking the Niobrara Member shale into which it intruded. Close to the dike, the shale layers have more or less fused together, while further out, the hornfels gradually transitions into unaltered shale.

The Ortiz Mountains are another cluster of laccoliths of the Ortiz monzonite.

Ortiz
          mountains

Ortiz Mountains. Looking south from 35 29.787N 106 05.653W

The Ortiz Mountains are gold mining country. Early mining exploited placer deposits, where gold that erodes out of bedrock accumulates deep in stream beds because of its high density. More recently, the bedrock ores themselves have been exploited.

Cunningham
          Hill

Ortiz Mountains. Deep telephoto shot looking south from 35 29.787N 106 05.653W

This photograph shows part of the tailings pile for the Cunningham Hill mine, which was first opened in 1828. The host rock containing the ore is the breccia filling a volcanic pipe through which much of the Espinaso Formation is thought to have been erupted. The content of gold in the ore is quite low, a few grams per ton, but this can still be profitably extracted using cyanide. Though extremely toxic, cyanide is one of the few chemicals that can pull gold into solution, so it is used to leach the gold out of the crushed ore. The gold-bearing solution is then treated with powdered zinc, which replaces the gold in solution. The result is a mixture of remaining zinc powder and metallic gold, from which the zinc can be removed by treatment with sulfuric acid.

The Cunningham Hill mine was most recently operated by Consolidated Gold Fields, which began open pit mining in 1973. Some 7 metric tons of gold (250,000 ounces), with a value of about $300 million at March 2017 prices, was extracted before the mine was closed again in 1986. Subsequent attempts to reopen the mine have foundered over environmental regulations and the fluctuating price of gold.

The northwest Ortiz Mountains include the Cerro Chato laccolith, where magma intruded into the relatively ductile beds of the Menefee Formation. The resulting rock has the same mixture of plagioclase and alkali feldspar as other Ortiz Monzonite outcrops, but with abundant hornblende needles.

Cerro
          Chato monzonite


Cerro Chato monzonite

Ortiz Monzoite of Cerro Chato laccolith. 35.3762124N 106.1762807W

It is striking that rock of such similar silica and alkali content and emplaced at about the same time in a similar manner can show such variety in its gross appearance.

Espinaso Formation

Digital relief map of Espinaso Formation exposures in
        the Jemez Mountains
Relief map of the Jemez with Espinaso Formation outcroppings highlighted in red.

The Espinaso Formation is slightly younger than the Ortiz Monzonite and is thought to be volcaniclastics from the portion of the Ortiz eruptions that made it to the surface. Volcaniclastics are beds of broken rock fragments produced by volcanic activity and subsequent erosion. The type section is at Espinaso Ridge itself.

Espinaso
          Ridge
Espinaso Ridge. Looking west from near 35.3897724N 106.2962957W

Here some 120 m (400') of volcaniclastics are exposed along Arroyo del Tuerto. 

Espinaso
          Ridge
Lower Espinaso Formation in Arroyo del Tuerto. 35.3821764N 106.3141617W

Debris flows

Much of the Espinaso Formation consists of beds made up mostly of large clasts rather than clay and silt. These are classified as debris flows. Debris flows form when loose shattered rock, of which there was an abundance in the Ortiz volcanic field, is lubricated by rainfall and begins moving downhill under the force of gravity. The beds in the previous photograph are dominated by clasts up to a few inches across.  As we will see shortly, It is not unusual for a debris flow to contain very large clasts.

It may be difficult to imagine how such a unit could once have flowed like a fluid, but debris flows have been captured on videotape:


Video of debris flow in Clear Creek Country, Colorado.

The destructive power of a debris flow is not hard to imagine.

There are some tuff beds interbedded with the debris flows of the Espinaso Formation.

Espinaso
          Ridge
Tuff beds in Espinaso Formation in Arroyo del Tuerto. 35.3816194N 106.3103107W

Tuff beds are formed of consolidated ash. There are appear to be two beds, with the lower bed rich in fragments of solid rock.

Espinaso
          Ridge lithic tuff
Lithic tuff in Espinaso Formation in Arroyo del Tuerto. 35.3816194N 106.3103107W

The mixture of dark fragments of volcanic rock and light volcanic ash is completely unsorted. This is very unlikely to happen in any sedimentary process. This was likely a pyroclastic flow, of which we'll learn much more later in the book.

Here is a thin tuff bed over a very coarse debris flow.

Espinaso
          Ridge debris flow
Espinaso Formation in Arroyo del Tuerto. 35.38184N 106.3089167W

Unfortunately, while Espinaso Ridge is located mainly on public lands, the road access is poor. More accessible exposures are found at La Cienega and along Galisteo Creek. This outcropping of Espinaso Formation breccia conglomerate is located north of La Cienega. It is formed mostly of broken fragments of andesite.

Espinaso Formation

Espinaso breccia. 35.570N 106.127W

A dike runs through the center of La Cienega that has been dated at about 20 million years in age. This is mapped as part of the Espinaso Formation.

Espinaso Formation dike

Espinaso dike. 35.564N 106.127W

The dike is composed of a very pretty hornblende trachyandesite. Trachyandesite is a rock formed from lava with an intermediate silica content, similar to andesite, but with an unusually high content of sodium and potassium. This trachyandesite contains numerous phenocrysts of hornblende.

Hornblende trachyandesite

Hornblende trachyandesite.. 35.564N 106.127W

The long needlelike hornblende crystals are obvious. The clump of crystals at the left end of the sample is either a clot (the crystals forming around some nucleus as the lava cooled underground before being erupted) or a xenolith.

Closer examination of the outcrop itself shows numerous xenoliths.

Xenolith

Xenoliths get geologists’ attention because they sample the rocks deep below the surface, including the lower crust and upper mantle. This dike has apparently gotten a lot of geologists’ attention, because there are drill holes all over it, presumably where geologists have taken samples of the xenoliths.

The Espinaso Formation is strikingly exposed northeast of Los Cerrillos.

Espinaso Formation

Espinaso Formation. 35 28.519N 106 04,199W

Espinaso Formation

The outcrop consists of numerous angular fragments of coarsely crystalline rock in a fine-grained matrix. This is almost certainly an example of a debris flow that has been preserved as conglomerate. The original flow was a mass of rock and mud that flowed off the slopes of the nearby volcanoes to produce the beds here. These were later cemented into hard rock.

Not all the Espinazo Formation is well cemented. In the Cañon area, there are beds of coarse conglomerate resembling river gravel that look deceptively young.

Espinaso
          Formation gravel beds


Espinaso
          Formation gravel beds

Dike intruding Espinaso Formation gravel beds. 35.578465N 106.132677W

The matrix between the clasts easily crumbles to the touch. However, two clues indicate these beds are much older than they appear. The first is that the clasts are entirely Ortiz Monzonite; there are no clasts from the younger Cieneguilla Basanite even though this formation is abundantly exposed just up the slope and as fresh fall on the road bed below. The second is the presence of a dike intruding the beds that cannot be geologically young.

Espinaso
          Formation dike

Espinaso Formation intruded by dike. 35.578465N 106.132677W

The dike consists of a dark rock thoroughly infiltrated with light veins. The rock could be an altered basanite, making this a source dike for the Cieneguilla Basanite, or it could be an unusual rock type such as a carbonatite. Laboratory tests will be be required to be sure. The veins are not calcite; they do not bubble in acid. They may be some form of zeolite. Regardless, this dike cannot be much younger than the Cienegilla Basanite (25 million years) and so proves the beds here are in the age range of the Espinaso Formation.

Further up the hillside, the Espinazo Formation is capped with what looks like a thin, crystal-rich tuff.

Espinaso
          Formation tuff bed


Espinaso
          Formation tuff bed sample

Espinaso Formation tuff bed. 35.57864N 106.1325372W

The bed is only a few feet thick. Above it are flows of the Cieneguilla Basanite, which I'll discuss shortly.  It is possible that this is actually a soil surface that was fused by the heat of the Cieneguilla Basanite lava, which was likely an unusually hot lava.

There are few exposures of sedimentary rocks of Oligocene age in New Mexico, and fewer fossils, but the end of the epoch marked an important milestone in the geologic history of the Jemez.

The Birth of the Rio Grande RIft

... the Rio Grande appears to follow a great fault zone along the foot of ranges which from Socorro northward face the east, while southward from that point they face the west. The relations of the larger structural valleys are too complex ... to permit an adequate description. The average width of this great structural trough is nearly 20 miles.

— W. Lindgren, L.C. Graton, and C.H. Gordon, 1910


The depression is a great rift belt... [and the] late Tertiary tectonic pattern is specialized and distinct enough to warrant the application of the term Rio Grande Rift Belt ...

— V.C. Kelley, 1952

In the last section, we saw that the rock beds at Espinaso Ridge and throughout the Hagan Basin dip to the east. However, when originally laid down, the beds were nearly level, showing that Hagen Basin is a block of crust that has been uplifted to the west. This deformation records the birth of the dominant tectonic feature of New Mexico, the Rio Grande Rift.


Rio grande
        rift
Wikimedia. Click to enlarge

Around 30 million years ago there was a major shift in the tectonics of the western United States. The compression produced by the subducting Farallon Plate along the west coast was replaced by tension and uplift as North American drifted onto the northern East Pacific Rise, a mid-ocean ridge. Precisely what began to take place under the western United States is still debated by geologists, but it seems likely that the remnants of the Farallon Plate began to sink deep into the mantle and  hot mantle rock rose to the level of the asthenosphere to take its place. As a result, for the last 30 million years, the western United States has been simultaneously uplifted and pulled apart, producing the Basin and Range Province so distinctive in topographic maps today. And the eastern boundary of that province runs right down the middle of New Mexico.

The Basin and Range Province is characterized by a series of north-trending ranges separated by valleys. The heart of this province is in Nevada and western Utah, but portions extend through Arizona into southwest New Mexico. The Colorado Plateau seems to be a block of particularly cold, rigid crust that has held together even as it has been pushed up by the rise of hot mantle beneath, but it has begun to split off the rest of North America along the valley of the Rio Grande. This is the Rio Grande Rift, which extends from near Leadville in central Colorado down at least as far as the area west of El Paso, where it blends into the surrounding basin and range geology. Seismological evidence shows that the crust has thinned significantly in the Rift, from a thickness of 50 km (30 miles) east and west of the Rift to 30 km (20 miles) under the Rift. There is also greater heat flow from deep in the earth through the Rift than in the surrounding crust.

The direction in which the Rift is pulling apart is not directly east-west. Instead, the separation is more west-southwest to east-northeast. This means that there is some left-lateral movement along the major faults bounding the rift. In the northern part of the Rift, this is consistent with the Colorado Plateau rotating slightly clockwise.

This map shows all exposures of rocks of Oligocene age or older in the Jemez region.

Digital relief map of pre-Neogene exposures in the
        Jemez Mountains
Relief map of the Jemez with pre-Neogene outcroppings highlighted in red.

These older rocks are useful for tracing the outline of the Rio Grande Rift, since they all formed before rifting began. Although the young volcanic rocks of the Jemez Mountains obscure part of the western margin of the Rift, one can see that this margin runs almost directly under the Jemez. The isolated exposure of Gallisteo Formation in the San Miguel Mountains, to the right of the center of the map, is an anomaly that suggests the western margin of the Rift is not sharply defined at the latitude of the southern Jemez. This is an accommodation zone in which the Rift shifts abruptly to the east as one moves north from the Albuquerque Basin to the Espanola Basin.

Espanola Basin
Espanola Basin. Looking north from 35 51.251N 106 08.896W

The Espanola Basin began to form late in the Oligocene as a broad, shallow basin, typical of ductile extension. The crust in north central New Mexico had been heated by the burst of volcanism in the early Oligocene, and it deformed relatively smoothly as it came under tension. At this time, the western boundary of the basin was located along the Sierrita Fault, just east of the southern Sierra Nacimiento. Similar basins had formed along the entire length of the Rio Grande Rift by 26 million years ago.

South of the accomodation zone, the displacement shifts west into the Santo Domingo Basin, sometimes regarded as a northern extension of the Albuquerque Basin.

Santo Domingo
          Basin
Santo Domingo Basin. Looking north from 35 18.883N 106 27.122W

North of the Jemez area, the displacement shifts east into the San Luis Basin.

The Cieneguilla Basanite

About 25 million years ago, silica-poor magma from the upper mantle reached the surface near La Cienega as a series of low-volume eruptions. This ultramafic magma was also rich in alkali metal oxides. Such alkaline magmas are thought to form from relatively undepleted mantle that has produced little magma in the past and so has retained most of its original alkali metal content. When undepleted mantle experiences a low degree of partial melting, so that only the most alkali-rich fraction of the mantle rock contributes to the magma, the magma is greatly enriched in alkali oxides.

Low-silica alkaline magma forms an uncommon kind of igneous rock called basanite. The excess sodium converts much of the feldspar that would otherwise have crystallized from the magma to nepheline, Na3KAl4Si4O16. Other common minerals in basanite are olivine, magnetite, and a calcium-rich pyroxene called clinopyroxene.

Cerro Seguro is the most prominent eruptive center for the basanite flows of the Cieneguilla Basanite.

Cerro Seguro
Cerro Seguro. Looking southwest from 35.5733543N 106.1329967W

Here a plug of basanite has intruded the underlying Ortiz Monzonite. The top of the hill is dark basanite while the lower slopes are light monzonite.  The lower slopes are covered with numerous small clasts of basanite.

 Basanite clasts
Basanite gravel. 35 34.310N 106 08.067W

The view from the top shows many features associated with the Rio Grande Rift.

Panorama
Panorama from atop Cerro Seguro. 35 34.228N 106 08.263W

The panorama begins with a view to the southwest, towards Sandia Crest. At extreme left are the nearer Ortiz Mountains. Mesa de Juanita is in the middle distance in front of Sandia Crest, just north of a bend in Interstate 25. It is underlain by relatively young Cerros del Rio Basalt with a cluster of small cinder cones atop the mesa.

The canyon that dominates much of the panorama, coming down from the north (right) and turning west to disappear into the horizon, is the canyon of the Santa Fe River. Northwest of the Santa Fe River canyon is Tsinat Mesa, also underlain by Cerros del Rio Basalt.

The hills just beyond the canyon are underlain by Ortiz monzonite, which rises like islands out of the much younger Cerros del Rio Basalt. Such islands of older rock completely surrounded by younger lava flows are described as kipukas, a Hawaiian term adopted by geologists.

The peak behind and to the right of these hills is Tetilla Peak, New Mexico's etymological counterpart to Grand Teton in Wyoming, and on the far size of the Santa Fe River canyon to the right you see exposures of highly weathered monzonite. North of these are darker hills of Cieneguilla Basanite flanking the canyon. The basanite hill on the west side of the canyon is marked with a white cross just visible in this panorama. This area is accessible via the road to the tiny settlement of Canon.

Agglomerate basanite
Cieneguilla Basanite knoll opposite Canon. 35.5784063N 106.1351187W

The basanite at the top of Cerro Seguro is described as agglomerated in a recent geological paper on this area, which confirms we are near the vent location. Rock with agglomerated texture is formed when blobs of semimolten lava are thrown out of the vent and pile up around it.

Agglomerate basanite
Agglomerated basanite from atop Cerro Seguro. 35 34.228N 106 08.263W

The rock does indeed look like individual blogs of magma tightly welded together. The rock is very black, very dense, and very fine-grained, but even with a loupe I doubt I would have recognized it as different from very dark basalt. It takes laboratory analysis to confirm the presence of nepheline and the low (44%) silica fraction.

Further down the hill, a pile of basanite boulders shown a distinctive red weathering.

Weathered basanite
Weathered basanite atop Cerro Seguro. 35 34.274N 106 08.167W

The magnetite in the basanite slowly oxidizes to red hematite when exposed to atmospheric oxygen. Basalt sometimes contains magnetite and gets a reddish patina, so this is not necessarily diagnostic for basanite, but there is a difference in degree. A sample from basanite here is more massive than the rock from the top of the cone.

Massive basanite
Weathered basanite from Cerro Seguro. 35 34.274N 106 08.167W

The gravel road to the settlement of Canon also passes readily accessible outcrops of the Cieneguilla Basanite. Here the basanite directly overlies conglomerate beds of the Espinaso Formation. A sample of the basanite from this area shows it to be a very uniform, dark, fine-grained rock.

Massive
          basanite
Basanite from Canon. 35.57864N 106.1325372W

Feldspathoids and silica saturation

Nepheline, Na3KAl4Si4O16, is an example of a feldspathoid. Like feldspars, feldspathoids are tektosilicates. Each silica or aluminum tetrahedron is joined to four neighbors at its corners to form a three-dimensional network. However, feldspathoids contain less silica than feldspars, and the framework of silica-alumina tetrahedra is more open. This makes room for the additional alkali metal atoms needed to supply the electrons to compensate for the added aluminum.

Feldspathoids generally crystallize only from magma that is enriched in alkali oxides relative to its silica content. In particular, feldspathoid is almost never found in the same rock as quartz, since any excess silica that might have formed quartz would have gone into converting feldspathoid into feldspar instead. Minerals that do not normally coexist with quartz are described as silica-undersaturated. A rock that contains such minerals is itself described as silica-undersaturated. A rock containing neither quartz nor silica-undersaturated minerals is described as silica-saturated. If the rock contains quartz, it is described as silica-oversaturated.

Those silica-undersaturated rocks which contain feldspathoids are described as alkaline, by contrast with the subalkaline tholeiitic and calc-alkaline suites discussed in the second chapter of this book. Alkaline magma is a product of a low degree of partial melting, because minerals bearing potassium and sodium are some of the first to melt under mantle conditions. Only if melting proceeds further will the magma become subalkaline.

Alkaline magmas are typical of continental rifts like the Rio Grande Rift. Here the crust is pulling apart relatively slowly, less than 2mm per year, versus centimeters per year for the mid-ocean ridges. In addition, the mantle under the continents is relatively undisturbed, making it more likely to be undepleted. The slow melting of undepleted crust can produce highly alkaline magmas. These are also more likely to penetrate the crust than subalkaline magmas, since alkaline magmas are more than 0.1 g/cc less dense than tholeiitic magmas.

The rocks of the Jemez volcanic field straddle the composition boundary between subalkaline and alkaline rocks. The older rocks tend to be alkaline, as do flows near the periphery of the field. Younger rocks near the center of the field tend to be subalkaline, suggesting than enough magma has been generated here to begin depleting the underlying mantle.

Ritito Conglomerate and Abiquiu Formation

Digital relief map of Abiquiu and Ritito Formation
        exposures in the Jemez Mountains
Relief map of the Jemez with Abiquiu and Ritito Formation outcroppings highlighted in red and green, respectively.

North of the Jemez Mountains, in the valley of the Rio Chama, lies the village of Abiquiu. This was originally founded as a community of genizaros during the Spanish era, in the mid-18th century. The Spanish word genízaro has the same roots as the English janissary, used to describe slaves (usually Christian boys) under the Ottoman Empire, who gained some measure of freedom and considerable social status by converting to Islam and becoming elite soldiers of the Turkish Army. The genízaros of Spanish New Mexico were native American boys (mostly Navajo) enslaved by the Comanche and sold to the Spanish as household servants or shepherds. They were indoctrinated in Catholicism by their Spanish masters, gained some measure of freedom, and eventually formed communities of their own, in which there was extensive intermarriage among tribes and with the Spanish. At Abiquiu, the genízaro community was awarded its own land grant or merced by the Spanish governor. The Merced del Pueblo de Abiquiu grant, originally awarded around 1750, is still a recognized legal entity today. The Abiquiu merced owns the northern part of Mesa Alta, which I have found is very well maintained, even in comparison with the adjacent National Forest lands. And, yes, they still raise some sheep, though cattle is more common now.

The village later became the eastern terminus of the Old Spanish Trail, whose western terminus was Los Angeles. The trail was first successfully traveled in 1829-1830 by Antonio Armijo, whose caravan took 142 days to make the round trip. Bode's General Merchandise, founded in 1890 as Grants Mercantile and purchased by Martin Bode in 1919, is a fixture of the modern community, which has now become somewhat gentrified. Artist Georgia O'Keefe settled permanently in the village in 1946 and produced several paintings of "The White Place", Plaza Blanca Canyon, which is underlain by dramatic white beds of ash-rich sediments now assigned to the Abiquiu Formation. This is one of the most distinctive formations of the northern Jemez.

The La Cienega eruptions were contemporaneous with the formation of the Latir volcanic field north of Taos. This included a giant caldera eruption from the Questa caldera. The first Latir eruptions occurred 30 million years ago as the rift opened, while the Questa eruption occurred five million years later. The Questa caldera has subsequently been uplifted by tectonic activity and deeply eroded, exposing valuable deposits of ores of molybdenum and other metals.

The Latir field produced extensive ash falls, while the Questa caldera produced voluminous pyroclastic flows. The pyroclastic flows formed the Amalia Tuff, which crops out as far to the southwest of Questa as the southern Tusas Mountains.

Amalia Tuff

Amalia Tuff
Amalia Tuff. 36 36.652N 105 59.000W

Quite a pretty tuff, full of quartz and sanidine phenocrysts. Such crystalline tuff is also found in Yellowstone area and in the San Juan Mountains.

A pyroclastic flow is a kind of avalanche of red hot particles of rock carried along by hot gases from the volcano. Because the hot particles are buoyed up by the hot gas, the avalanche can travel great distances (this outcrop is about 25 miles from the Questa caldera) and destroy everything in its path. The hot rock then settles out on the ground. I will have much more to say about tuffs and pyroclastic flows later in this book.

The ash from the Latir field is an important component of the Abiquiu Formation, which crops out throughout much of the northern Jemez. This formation originally included three members: a lower member of somewhat unremarkable, poorly consolidated light tan sandstone and conglomerate beds; a thin layer of limestone and chert (the Pedernal Chert); and the white upper layers, which are rich with volcanic ash and clasts of Amalia Tuff. This division and assignment has recently undergone revision. The lower Abiquiu Tuff turns out to correlate with the Ritito Conglomerate, which was originally described to the north of Abiquiu, and newer maps and papers of this area now assign exposures that were formerly described as lower Abiquiu Tuff to the Ritito Formation. Only the distinctive white upper beds are now assigned to the Abiquiu Formation, while the Pedernal Chert remains an informal unit name. This sort of thing is neither terribly uncommon nor particularly unhealthy in geology. Formation designations change as further study improves our understanding.

The white beds of the Abiquiu Formation can be traced far to the west and south, cropping out in the Jemez as far to the southwest as lower Canon de San Diego. For the traveler approach Abiquiu from the south, the first dramatic exposures of the formation are visible on the southwest flank of Sierra Negra

Sierra Negra
Sierra Negra. Looking north from 36 11.809N 106 14.396W

The white cliffs halfway down the left side of the mountain are Abiquiu Formation. There is a fault cutting right through the mountain, at about the location of the right side of the Abiquiu outcrop. To the east (right) are younger beds of the Santa Fe Group, and the top of the mountain east of the fault is capped by a small basalt lava flow that had been dated to about 4.8 million years old.

Sierra Negra
Sierra Negra. Looking north from 36 12.231N 106 14.805W

The fault is part of the western edge of the Rio Grande Rift. It displaces the basalt flow by just a a few meters, showing that activity has been limited on this fault in the last four million years.

The Abiquiu Formation is particularly well exposed in Plaza Blanca Canyon, Georgia O'Keefe's "White Place", north of Abiquiu. 

Plaza Blanca Canyon
Abiquiu Formation at Plaza Blanca Canyon. 36 14.233N 106 18.366W

The canyon is presently owned by Dar al Islam Mosque, but is open to the public.

Plaza Blanca
          Canyon
Abiquiu Formation at Plaza Blanca Canyon. 36 14.048N 106 18.180W

The darker bed in the cliffs in the center of the panorama contains numerous clasts and may mark an ancient stream bed or a debris flow. Here's a closer view.

Debris flow
          in Abiquiu Formation
Possible debris flow in Abiquiu Formation at Plaza Blanca Canyon. 36 14.048N 106 18.180W

The ridge here lies along the trace of the Plaza Colorado fault. The fault plane itself is dramatically exposed.

Fault plane of
          Plaza Colorado Fault at Plaza Blanca Canyon
Fault plane of Plaza Colorado Fault. 36 14.073N 106 18.235W

Slickenside is present along the fault plane.

Slikenside at Plaza Colorado fault
Slickenside at Plaza Colorado Fault. 36 14.073N 106 18.235W

Slickenside is a distinctively polished rock surface on the plane of a fault, caused by the motion of the rock faces past each other. The striations here are consistent with this being a normal fault, with primarily up/down motion.

A peculiar feature of some of the Abiquiu Formation beds at Plaza Blanca is the presence of well-developed conjugate jointing.

Conjugate
          jointing in Abiquiu Formation
Conjugate jointing in Abiquiu Formation at Plaza Blanca Canyon. 36 14.117N 106 18.416W

It is not uncommon for rock to have joints (fissures) that meet at right angles, with one of the angles commonly being the bedding angle. Joints that do not meet at right angles are described as conjugate joints. They're something of a geological curiosity, for which I've not come across a good explanation.

Some surfaces show well-developed manganese dendrites.

Manganese
          dendrites at Plaza Blanca
Manganese dendrites at Plaza Blanca Canyon. 36 14.294N 106 18.388W

The black patterns are manganese oxides formed by weathering of the rock.

In the area west of Abiquiu, geologists have carefully studied the sediments making up the Abiquiu Formation, and find a transition from sediments derived mostly from the San Juan volcanic field to the northwest to sediments derived mostly from the Latir volcanic field to the northeast. The former is relatively poor in quartz while the latter is rich in quartz. The feldspar composition also shifts, from mostly plagioclase to mostly sanidine. The transition from one to the other suggests changes in drainage patterns associated with the very earliest opening of the Rio Grande Rift in the Taos area.

Further west, past the ridge of Cerrito Blanco, the Abiquiu Formation is visible just north of the road.

Cerrito Blanco
West side of Cerrito Blanco showing fault offset of Tesuque and Abiquiu Formations.

Just west of here, across a fault boundary that is thrown down to the east, the Ritito Conglomerate is well exposed.

Ritito Conglomerate in road cut
Ritito Conglomerate in road cut west of Abiquiu. 36 13.855N 106 22.846W

The Ritito Conglomerate contains an abundance of rock clasts, embedded in a light ash-rich sandy matrix. This is rarely very well cemented, except just below the Pedernal Chert at the top of the formation. Areas underlain by the Ritito Conglomerate thus tend to form gentle slopes and rounded hills, like this one. The clasts are dominated by quartzite and granitoid rocks of Precambrian age, likely weathered off the Tusas Mountains to the north.

Superb exposures of the Ritito Conglomerate are found along Red Wash Canyon.

Ritito
          Conglomerate along Red Wash Canyon

Ritito Conglomerate along Red Wash Canyon. 36 14.354N 106 22.532W

The area also features a slot canyon.

Slot
          canyon in Ritito Conglomerate


Slot
          canyon in Ritito Conglomerate

Slot canyon in Ritito Conglomerate. 36 14.103N 106 22.562W

The clasts in the conglomerate here are varied and interesting, including epidote and garnet-bearing metamorphic rocks. These very thick deposits of very coarse conglomerate indicate that the source rock was very close to this location.

The Abiquiu Formation can be traced west to Canones Mesa

Canones Mesa
Canones Mesa

and the Ritito Conglomerate is beautifully exposed north of the village of Canones, where it lies atop El Rito Formation.

Ritito Conglomerate at Canones
Ritito Conglomerate north of Canones. 36 11.369N 106 26.796W

Exposures occur on the lower slopes of Cerro Pedernal. Cliffs of Abiquiu Formation are prominent along Forest Road 100 in the valley of the Rio Encino as it climbs towards the La Grulla Plateau, southwest of Cerro Pedernal.

Cliffs of Abiquiu
        Formation
Abiquiu Formation. 36 08.678N 106 30.805W

Nearby, the Ritito Conglomerate is well exposed in the road cut.

Ritito Formation in
          road cut
Ritito Formation. 36 08.342N 106 30.784W

Some exposures of the Ritito Conglomerate in this area contain large clasts of Amalia Tuff at the top of the formation, but I found none here. The clasts were all quartzite, gneiss, and other Precambrian rocks likely weathered from the Tusas Mountains.

The Pedernal Chert is particularly well exposed here, forming a prominent ledge separating the Ritito Conglomerate and Abiquiu Formation:

Pedernal
          Chert
Pedernal Chert forming the prominent ledge. Looking north from 36 08.418N 106 30.630W

It's a bit of a scramble to reach the ledge, beneath which are unusually well cemented beds of Ritito Conglomerate. Here's the chert layer close up.

Pedernal
          Chert
Pedernal Chert. Car keys for scale. 36 08.486N 106 30.622W

The Pedernal Chert is a bed of amorphous silica and limestone. The limestone probably formed as a calcium-rich soil atop the Ritito Conglomerate under arid conditions. Later, the Abiquiu Formation was deposited on top of the limestone layer, and groundwater circulating through the ash-rich Abiquiu Formation picked up silica that was redeposited in the limestone layer. Deposition of amorphous silica in limestone beds is very common, though the chemistry seem not to be fully understood. The chert varies in color from milky white to reddish brown to pitch black.

Pedernal
          Chert
Pedernal Chert. 36 08.486N 106 30.622W

Pedernal
          Chert
Pedernal Chert. 36 08.486N 106 30.622W

Large clasts of chert can be found around the base of Cerro Pedernal, where it was exploited by early native Americans for tool making. Pedernal Chert artifacts have been found as far east as the eastern slopes of the Sangre de Cristo Mountains, and Spanish archives mention chert quarries in the mid-18th century.

Another possibility for the origin of the Pedernal Chert is that it formed in an alkaline lake that covered this region. Chert is forming by this mechanism today in parts of the East African Rift Valley. This could be tested by searching for traces of rare sodium silicate minerals in the chert, but to my knowledge, this has not been done.

Further south along Forest Road 100, the ascending road cuts through Abiquiu Formation.

Abiquiu
        Formation
More Abiquiu Formation. 36 07.850N 106 30.746W

The higher elevation at which the Abiquiu Formation is exposed in the northern Jemez, compared with the exposures in the Abiquiu area to the northeast, reflects the displacement along the Canones Fault Zone, which lies between the two areas.

There are deeply weathered but extensive outcrops of Pedernal Chert in the San Pedro Mountains.

Pedernal Chert in
          the San Pedro Mountains
More Abiquiu Formation. 36 07.850N 106 30.746W

These outcrops are particularly significant because they provide strong evidence that the San Pedro Mountains were uplifted less than 25 million years ago. These beds are 240 m (800') higher than corresponding beds east of Jarosa.

The Ritito Formation is exposed again in the northwest wall of Valles Caldera, in Valle San Antonio, which can be reached by passenger vehicle with a back country permit from Valles Caldera National Preserve. Viewing the exposure requires hiking some distance along San Antonio Creek.

Abiquiu Formation in Valle San Antonio

Ritito Formation in northwest caldera rim. 35 58.102N 106 36.742W

The exposure here is somewhat poor, and easily confused with nearby lacustrine beds and Bandelier Tuff. However, it establishes that the base of the topographic rim of the caldera is located here.

Further south along the caldera rim, in upper San Antonio Canyon, are some better exposures.

Abiquiu Formation in Valle San Antonio

Abiquiu Formation in northwest caldera rim. 35.9635763N 106.6265337W

These definitely appear to be the ash-rich beds of the Abiquiu Formation proper rather than the Ritito Conglomerate. Spotty exposures of the Abiquiu Formation continue south along the caldera rim, of which the most accessible is in a road cut along State Road 126.

Abiquiu Formation
More Abiquiu Formation. 35 53.056N 106 39.781W

The Abiquiu Formation is also present in the southern Jemez, in upper San Juan Canyon. The easiest exposures for the casual visitor to reach are in landslide deposits along the road.

Abiquiu Formation
        in landslide

Abiquiu Formation
Abiquiu Formation in the southern Jemez. 35 46.262N 106 36.697W

This is located at about the same elevation as in the northern Jemez. Under the loupe, this looks like a mass of small quartz grains with considerable lithic fragments and abundant white cement. This makes the rock a lithic wacke, rather than a lithic arenite.

More pristine exposures on the east side of San Juan Canyon can be reached by a short but strenuous hike from Forest Road 10.

Abiquiu
          Formation
Abiquiu Formation on east side of San Juan Canyon. 35 45.293N 106 37.256W

The exposures here are quite thick, with a particularly spectacular exposure in a drainage into the east side of the canyon.

Abiquiu
          Formation
Abiquiu Formation on east side of San Juan Canyon. Looking east from 35 45.293N 106 37.256W

There are also exposures on the west side of the canyon, which can be reached by a less strenuous but longer hike down the canyon along an abandoned forest road branching off of Forest Road 10.

The southernmost exposures of Abiquiu Formation in the Jemez are in southern Canon de San Diego, occurring in two isolated erosional remnants just south of Red Rocks and in more substantial exposures across the Jemez River on tribal lands.

Abiquiu
          Formation
Abiquiu Formation remnants near Red Rocks. 35 38.251N 106 43.429W

The exposure here is the thin layered white beds, sandwiched between red Abo Formation on which the Abiquiu was deposited in this area and dark beds of much younger terrace gravels laid down by the Jemez River before it cut down to its current level.

The Gilman Conglomerate

Also present in the southwestern Jemez is the Gilman Conglomerate. The type section is prominent in the cliffs west of the former logging camp of Gilman.

Gilman
          Conglomerate
Gilman Conglomerate type section. Looking west from 35 42.977N 106 45.568W

At bottom are red beds of the Permian Abo Formation, some 300 million years old. Above are the beds of the Gilman Conglomerate. Above those are younger gravel beds, then a mesa of Tsherige Member, Bandelier Tuff. We know that the gravels must be older than 1.25 million years, since that’s the age of the Tsherige Member lying above them, but it’s very hard to date gravels. The individual clasts in the gravel are Paleozoic but that just tells you how old the rocks they eroded from are; it doesn’t tell you when the beds formed.The same is true  for the Gilman Conglomerate, though it’s clearly older than the gravels. It has the look of volcaniclastics, and is mapped as such in the quadrangle map for this area. The clasts in its uppermost beds have a minimum age of 25 million years, similar to the Ritito Conglomerate. Some geologists now believe that the Gilman Conglomerate correlates with the Ritito Formation and should probably be regarded as a member of the Ritito Formation. Others note that, while it is roughly the same age as the Ritito Conglomerate, it must be derived from an Oligocene volcanic field that is now buried under the southern Jemez. I have included it in the map at the start of this section as part of the Ritito Formation.

I've devoted a lot of space to the Abiquiu Formation, because its distinctive white beds nicely trace the approximate boundary of the west side of the Rio Grande Rift through the Jemez area. Further east, the Abiquiu Formation was either never deposited, was eroded off the Pajarito Uplift before the Rio Grande Rift opened, or is now buried under younger rift fill sediments of the Santa Fe Group. There is one prominent exception: an exposure along La Bajada.

Abiquiu Formation at La Bajada

Abiquiu Formation on La Bajada. 35 32.898N 106 13.971W

This is not only the finest exposure of this formation south of the Jemez; it's very nearly the only exposure south of the Jemez. The reasons why the formation is so well preserved here are unclear, but may be related to its location along the La Bajada Fault, which is thrown down to the west. This is the area where the Rio Grande Rift jumps to the west, and so this exposure marks the east edge of the Rift at this latitude.

Or it's possible this is not Abiquiu Formation at all, but is Espinaso Formation that has been misidentified. The ages and character are similar. They could be distinguished by careful clast analysis and charting of paleocurrent directions, but to my knowledge this has not yet been done.

Dikes

The Ritito Formation is intruded by the Cerrito de la Ventana dike system. One dike is crossed by the highway west of Abiquiu.

Basalt dike

Basalt dike

Basalt dike
Basalt dike of the Cerrito de la Ventana dike system. 36 13.046N 106 21.142W 

The basalt in the dike ihas most recently been dated at 18.9 million years old, making it part of an episode of volcanic activity predating the current Jemez volcanic field. It also helps us put a lower limit on the age of the Ritito Formation, which is poor enough in fossils that its age has been hard to pin down. You can see a band of the surrounding Ritito Formation baked red by the heat of the dike though, curiously, only on one side of one part of the dike.

Basalt dike
        showing baked zone in country rock
Basalt dike with baked zone in country rock

The basalt is a dense black rock with occasional veins of almost pure quartz that indicate the dike has undergone some hydrothermal alteration.

Close up of
        basalt

Quartz
        veins
Close up of basalt with quartz veins

Basalt is quite low in silica, while quartz is more or less pure silica. The quartz must have been injected long after the basalt cooled, probably as a solution in water formed in the very last part of the underlying magma chamber to cool.

This dike continues north past the western end of Plaza Blanca, where it splits into multiple strands.

Cerritos
          de la Ventana dike west of Plaza Blanca
Cerrito de la Ventana dike west of Plaza Blanca. Looking west from 36 14.178N 106 18.680W

This is but one (particularly prominent and accessible) dike of the Cerrito de la Ventana system. An example of a smaller dike is found near Red Wash Canyon.

Small dike of
          Cerrito de la Ventana system

Small dike of Cerrito de la Ventana system. 36 14.097N 106 22.257W

Another dike, or perhaps this same one in a different exposure, is found in the west wall of Arroyo del Cobre.

Here is the paleomap for the Abiquiu Formation.

Amalia Tuff
Paleotopographic map of the Abiquiu Formation

The map shows the limits of the Abiquiu depositional basin reasonably well. This reached to about the latitude of the Santa Fe River canyon, although the Abiquiu Formation is notably absent between the Santa Fe Group and Galisteo beds near the San Miguel Mountains. However, the Santa Fe River beds may not actually be Abiquiu Formation.

And this will be the last of the paleotopographic maps in this book. As we approach the present day, formations begin to overlap in time in a way that makes such maps decreasingly useful.

The beginning of the Neogene and the Santa Fe Group

During his 1867 expedition, Hayden noted that Santa Fe was surrounded by badlands -- he likened them to the "Mauvais Terres" of the Dakotas -- likely of Tertiary age. These consisted of sandstones and "marls" (unconsolidated limy clay) with some conglomerates, and he named them the Santa Fe Marls. This was the first formal name assigned to a geologic unit in New Mexico. He found that the Santa Fe Marls extended north to Black Mesa, that they dipped gently to the west, and that they were of enormous thickness -- at least 1500' (500m). He reported that the marls formed mesas on the west side of the Rio Grande, with two distinct levels. In this he was mistaken; likely he did not actually cross the river to the Pajarito Plateau, else he would have discovered that the upper level was not sedimentary rock but tuff. Then again, he would not have been the last visitor to the Pajarito Plateau to mistake the pink finger mesas of the Bandelier Tuff for sandstone.

As with the Galisteo Sands, the name Hayden gave to the Santa Fe Marls has remained attached to these beds to the present day.


Digital relief map of Santa Fe Group exposures in the
      Jemez Mountains
Relief map of the Jemez with Santa Fe Group outcroppings highlighted in red.

23 million years ago, the world was beginning to look very much like it does today, although Iraq and New Guinea were still under water and Central America was not yet a continuous ribbon of land.  In the western United States, the compression of the Laramide orogeny had reversed itself, and instead of the crust being compressed, it began to be pulled apart. At the same time, the asthenosphere under the western United States became hot enough to lift the entire region as the western United States drifted onto the northernmost part of the East Pacific Rise. This was the period when rifting along the Rio Grande Rift really began to get going.

Geologists have chosen this moment in time as the start of the Neogene. The Neogene is further divided into the Miocene and Pliocene Epochs. During the Miocene, the severe global cooling at the end of the Oligocene moderated somewhat, but cooling resumed by 15 million years ago and temperatures plunged at the end of the Miocene, about 5.33 million years ago.

Only a small part of the crustal thinning in the Rift is accounted for by lower surface elevations along the Rift. The rest is due to the mantle rising into the rift. The rift is bounded by deep faults on one or both sides, and the combination of deep faulting, a shallow, hot mantle, and the zone of fertile rock along the Jemez Lineament creates conditions for sustained volcanic activity.

The faults bounding the rift are not single large faults, but numerous parallel faults forming one or more fault zones. East of Santa Fe the Pecos-Picuris Fault Zone marks the west face of the Sangre de Cristo Mountains and plunges to the west. To the west, which is the deeper side of the Rift in the Jemez area, there are three systems of faults. These are the Pajarito Fault Zone, the Canada de Cochiti Fault Zone, and the Jemez Fault Zone. Further north, the Pajarito Fault Zone turns east to join the Embudo Fault Zone, while the Jemez and Cochiti Fault Zones disappears under the Jemez Mountains before possibly joining with the Canones Fault Zone.

The Jemez Fault Zone runs roughly along Cañon de San Diego. It is difficult to trace over much of its length, but one strand (the Sierrita Fault) is fairly obvious at Guadalupe Box.

Fault surface
          at Guadalupe Box
Fault scarp at Guadalupe Box. Looking northeast from near 35 43.882N 106 45.754W

The darker rock to the left is weathered Precambrian hornblende-biotite quartz monzonitic gneiss ("granite" is close enough for most purposes.) To the right are Permian sedimentary beds. The local topography suggests, and geological mapping confirms, that this is actually a large fault. The younger beds have been thrown down by the fault a considerable distance, bringing them into contact with the Precambrian rocks.

The Canada de Cochiti Fault Zone is largely located on tribal lands in some of the less accessible areas of the southern Jemez, but one of the major strands of the fault, the Camada Fault, passes just west of the Veterans Memorial Overlook at Kashe-Katuwe Tent Rocks.

Trace of Camada
          Fault west of Veterans Memorial Overlook
Trace of Camada Fault, Canada de Cochiti Fault Zone. 35 40.330N 106 27.430W

The white beds to the left are Peralta Tuff, while the reddish beds to right are thin surge beds of the Tsherige Member, Bandelier Tuff. The right side is thrown down by the fault.

Neither the Jemez nor the Canada de Cochiti fault zones have been very active in the recent past, as shown by the lack of much displacement of Bandelier Tuff beds overlying the faults.

"Every valley shall be exalted, and every mountain and hill made low"

The Pajarito Fault Zone is much easier to trace, probably because it is now taking up most of the extension in this part of the Rift and has been significantly displaced since the Tsherige Member of the Bandelier Tuff was erupted, 1.2 million years ago. This fault zone joins the La Bajada Fault Zone to the south and the Embudo Fault Zone to the north. The Pajarito Fault Zone is most obvious southwest of Los Alamos, along the foothills of the Sierra de los Valles, where it forms a prominent escarpment that shows up well on topographic maps:

Pajarito fault
Pajarito and Bajada faults. Downthrown sides marked with red dots.

The Pajarito Escarpment is quite obvious as one approaches it from the east along State Road 4.

Pajarito Escarpment at
          ascent of State Road 4
Pajarito Escarpment. 35 50.030N 106 21.351W

You can see that the highway winds up the face of the escarpment. Looking north and south from one of the hairpin turns in the road, we get an excellent view of the Pajarito Escarpment in both directions.

Pajarito Escarpment
          south of hairpin turn
Pajarito Escarpment south of hairpin at 35 49.991N 106 21.804W

The view to the south shows the Pajarito Fault dividing into two strands,  one forming the escarpment at the center of the picture, and a second strand to the east (left) forming a smaller escarpment.

Pajarito
          Escarpment south of hairpin turn
Pajarito Escarpment along Ponderosa Trail. 35.8213873N 106.3613967W

The fault zone is exposed at Upper Crossing in Frijoles Canyon.

Pajarito
          Escarpment south of hairpin turn
Upper Crossing. Lookin north from 35.8143423N 106.3630143W

The main strand of the fault cuts the canyon right of center, where the switchback trail is just visible left of a forested area. A second, probably much older, strand cuts from the notch just left of the peak of the canyon rim down to the left on the far side of the irregular darker rock mass. This is an outcrop of Tschicoma Formation dacite. Two small strands correspond with the two small notches to the right of the rim peak, and slightly displace the Bandelier Tuff beds beneath.

Further south, the Pajarito Escarpment also cuts through the San Miguel Mountains, a small outlying range of the Jemez. The fault scarp forms the east slope of St. Peter's Dome and exposes older rocks down to the Galisteo Formation.

Rabbit
        Hill area
Pajarito Escarpment between Sierra de los Valles and San Miguel Mountains. The peak is Rabbit Hill.
 Panorama of San Miguel
        Mountains from the east
Panorama of Frijoles Canyon from Frijoles Overlook. 35 46.385N 106 15.714W

One of the strands of the fault enters Capulin Canyon not far from the Stone Lions shrine.

Capulin Canyon
Pajarito Fault strand in Capulin Canyon. 35 46.385N 106 15.714W

The fault runs along the far side of the mesa at center and enters the canyon to its right. The displacement here is about 300 feet (100 meters) since the Tsherige Member forming the mesa cap was emplaced 1.25 million years ago. A much smaller antithetic fault runs to the west of the main fault and visibly displaces the white beds of Otowi Member in the side of the mesa. (We'll revisit the Tsherige and Otowi Members of the Bandelier Tuff in due course.) The displacement is down to the west, opposite of the main fault, which is what makes it an antithetic fault. A small fault associated with a larger fault that is downthrown on the same side is called a synthetic fault. Antithetic faults are not uncommon near major faults.

A smaller strand occurs to the west and displaces the Tsherige Member by a few tens of feet:

Fault near Galisteo outcrops
Small fault strand east of San Miguel Mountains. 35.7483163N 106.3434158W

Although the panorama slightly distorts the image, one can see that the rim of Bandelier Tuff is significantly higher left of the rocky outcrop of Galisteo Formation than to its right. The fault trace is just to the right of the rocky outcrop. (We will learn more about the Bandelier Tuff later in the book.)

The Pajarito Fault splits south of the San Miguel Mountains before both splays pass south out of the Jemez area. The east splay forms the La Bajada Fault, which marks the western boundary of the Cerros del Rio. This fault is thrown down to the west, unlike the main Pajarito Fault, and it marks the point where the Rio Grand Rift is displaced to the west. The fault runs roughly along the base of the La Bajada escarpment, and the exact trace is exposed in the hills west of the Galisteo dam:

La Bajada fault trace
Trace of La Bajada Fault west of Galisteo dam. 35 27.780N 106 14.241W

The reddish, nearly vertical, thin line on the side of the hill is the precise trace of the La Bajada Fault. The fault runs nearly at the foot of the La Bajada escarpment to the north and south, but in this area, Galisteo Creek has eroded the escarpment back so that the fault runs some distance west of it.

The pinkish beds to the left, on the downthrown side, are young Rio Grande Valley alluvium deposits. The reddish beds to the right are much older Chinle Group beds. The yellowish beds immediately to the right (east) of the fault, between the fault trace and the Chinle beds, may be a sliver of Entrada Formation. Or they may be fault gouge, the pulverized rock formed by motion along the fault, which has since been cemented by minerals brought in by groundwater flowing in the fault zone. The geologic map shows Entrada Formation at nearby locations immediately east of the fault trace.

There is a strand of the La Bajada Fault that branches from the main fault just south of the Cochiti exit from I-25.

La Bajada fault strand

La Bajada fault strand. 35.5151689N 106.2284233W

The smaller eastern strand of the fault turns northwest and runs along the base of the cliffs to the right. The yellowish tan beds forming the edge of the cliff are the Salt Wash Member of the Morrison Formation, composed mostly of sandstone laid down by a river system. These cliffs end on the skyline just left of center. The gray cliffs peeking over the Salt Wash Member at the center of the panorama are much young basalt of the Cerros del Rio, with an age close to three million years.

The red beds north (left in the photo) of the fault belong to the Galisteo Formation. Their presence next to the much older Salt Wash Member shows that the beds to the left have been thrown down a considerable distance by the fault. Also, if you look closely at the beds at the left, you will see that they are tilted upwards. The beds have been deformed by fault drag, but become more or less level further north.  The motion along the fault is not strictly up and down; there was some left lateral slip on the fault.

After the fault passes over the skyline, just left of center in the photograph (where the red soil gives way to tan soil), it begins to curve around more towards the north, but continues separating young Eocene and Oligocene beds from older Jurassic beds. The fault then passes under the lava flows of the Cerros del Rio and becomes untraceable a few miles further north. These flows unhappily (for geologists) conceal much of the accomodation zone where the eastern boundary of the Rio Grande Rift takes a jog to the east.

The main Pajarito Fault, to the west of La Bajada, cuts through Eagle Canyon just west of the old Dixon apple orchard.

Eagle Canyon
Eagle Canyon from  north rim overlook. 35 42.491N 106 23.080W

The escarpment at right marks the trace of the Pajarito Fault. (Click for a full resolution version.) The fault trace is well exposed in the canyon wall -- in fact, this is the most dramatic exposure of the Pajarito Fault I've seen. To the left (east) is Bandelier Tuff, while to the west is an exposure of dark andesite of the Paliza Canyon Formation capped by Bandelier Tuff. Note that the Bandelier Tuff is significantly thicker on the left side of the fault; there was already an escarpment 1.25 million years ago and the pyroclastic flow that produced the tuff ponded on the east side along the escarpment. This shows that the fault was already active before 1.61 million years ago, when the Otowi Member of the Bandelier Tuff was emplaced. The displacement in the Bandelier Tuff shows that the fault has been active in the 1.25 million years since the emplacement of the Tscherige Member of the Bandelier Tuff.

One of the findings of geologists that most surprises me is how long-lived deep fault zones like the Pajarito Fault Zone and the Jemez Fault Zone are. The area around the Sierra Nacimiento Mountains has been repeatedly uplifted since at least the time of the Ancestral Rocky Mountains, 300 million years ago. The Pecos-Picuris Fault Zone has been in existence nearly as long. The Pajarito Fault Zone was active during the Laramide Orogeny fifty million years ago; what's more, until the Rio Grande Rift opened, the Pajarito Fault Zone worked in reverse of how we see it today, with the current location of the Jemez Mountains forming a low area between the Sierra Nacimiento and a high area, the Pajarito Uplift, between the Pajarito Fault Zone and the Pecos-Picuris Fault Zone. The Pajarito Uplift is now the Espanola Basin of the Rio Grande Rift. Similar basins opened up the length of the Rio Grande Rift, including the San Luis Basin to the north and the Albuquerque Basin to the south. Between each basin, the Rift takes a sharp jog to the right.

We'll return now to that hairpin turn on State Road Four where it climbs the escarpment, and this time we'll look north.

Pajarito Escarpment north of
          hairpin turn
Pajarito Escarpment north of hairpin at 35 49.991N 106 21.804W

Here the escarpment runs along the eastern foot of the Sierra de los Valles until it joins the Santa Clara Canyon Fault north of Los Alamos.

North of the Jemez, there is extensive rift faulting associated with the opening of the Rio Grande Rift centered around the village of Canones. Deformation along this fault zone has been traced as far west as Coyote and extends east to where the zone merges with the Embudo Fault Zone. The main strand of the Canones Fault forms an impressive escarpment just north of the highway where it begins to climb up to Abiquiu Reservoir.

Canones Fault
          scarp showing Permian and Triassic beds
Canones Fault scarp. Looking northwest from 36 14.015N 106 23.294W

This is generally taken to mark the boundary between the Rio Grande Rift and the Colorado Plateau to the northwest.

Another strand of the fault dramatically offsets exposures of the Jurassic Entrada and Todilto Formations on the north end of Canones Mesa.

Displacement of Jurassic beds on Canones Mesa
Entrada and Todilto Formation offset by fault through Canones Mesa. Looking southwest from 36 14.149N 106 24.029W

Red and white beds of the Entrada Formation, capped with a thin bed of Todilto Formation, crop out on either side of a landslide down the north face of the mesa. On the east (left) the beds are thrown down about 70 meters (200 feet) by the fault passing between the two outcrops. This fault does not displace the lava on top of the mesa, showing that this particular strand of the Canones Fault Zone has not been active in at least three million years.

As our story comes closer to the present, where more geological evidence is preserved, the growing cast of characters will tend to break down any telling as a regular progression in time. We will find ourselves following one story line, then backtracking to catch up on another. The story of the Santa Fe Group extends from the beginning of the Neogene almost to the present day, and we will tell much of it here before returning to the middle Neogene and picking up the story of the Jemez volcanic field.

The Lower Santa Fe Group

As the rift opened, the Pajarito Uplift subsided to become the Espanola Basin. It is estimated that the basement rock beneath Espanola is now about 4600 meters (15,000') lower than corresponding basement in the Sangre de Cristo Mountains to the east. However, there is not today a yawning chasm four kilometers deep in this area. As the Pajarito Uplift dropped to become the Espanola Basin, the basin began to fill with sediments eroded off the nearby highlands.

These sediments formed beds that are, in some places, thousands of meters thick. Similar thick beds of sediments accumulated in the other basins forming the Rio Grande Rift. These rift sediments are known collectively as the Santa Fe Group. For the most part, the Santa Fe Group consists of weakly consolidated muddy sandstone, though some beds are locally well indurated and the clast size varies from mudstone to coarse conglomerate. Some of the formations making up the Santa Fe Group were deposited by the ancestral Rio Grande River and its tributaries, while others were deposited by wind. Continuing regional uplift caused by the hot asthenosphere under the western United States has resulted in erosion of these thick beds of sediments, which form spectacular badlands throughout the Santa Fe and Espanola area.

The nomenclature for the Santa Fe Group is unusually complicated. This is in part because the individual formations are not always easy to distinguish in the field, and in part because there were multiple sources of sediments. A sedimentary rock formation normally represents a correlated interval in time, but the the sediments laid down in a particular time interval may vary systematically from place to place because they came from a different source. A further complication is that many named members of the group are found to interfinger with each other.

The Santa Fe Group is divided into lower and upper sections. The lower Santa Fe Group are formations deposited in closed basins, called bolsons (Spanish bolsón, "purse"), that often had a central playa or intermittent saline lake, surrounded by piedmont deposits from the surrounding high ground. The upper Santa Fe Group are formations deposited in open basins integrated into an axial river system, either the ancestral Rio Grande or one of its major tributaries.

There has been much disagreeement on which formations to include in the Santa Fe Group. The oldest formation that is sometimes assigned to the Santa Fe Group is the Abiquiu Formation, which partially filled the rift in its earliest stages of formation. The slightly younger Zia Formation of the southwest Jemez is also assigned to the Santa Fe Group by some geologists but not by others. The oldest rift fill sediments generally accepted as part of the Santa Fe Group belong to the Tanos Formation.

Tanos Formation

Digital relief map of Zia Formation exposures in the
        Jemez Mountains
Relief map of the Jemez with Tanos Formation outcroppings highlighted in red.

The Tanos Formation was named relatively recently, to designate what had previously been mapped simply as basin floor sediments of the undivided Santa Fe Group. It crops out only in the Hagen Basin, where erosion of a tilted block of crust on the margin of the Rio Grande Rift has exposed the very oldest beds of the Santa Fe Group where they rest on the Espinaso Formation.

The base of the formation is a coarse conglomerate.

Tanos Formation

Tanos Formation
Tanos Formation. 35.384915N 106.3055342W

This is overlain by sandstone and siltstone.

Tanos Formation
Tanos Formation. 35.385717N 106.3064624

Under the loupe, the sandstone is found to be a poorly sorted, dirty sandstone.

The Tanos Formation is interpreted as being deposited in a closed basin produced by very early rifting, with occasional thin limestone beds from ponds and other indications of poorly established drainage to the north. A basalt flow interbedded with the lower sediments has been radiometrically dated as 25 million years old. The tilting of these beds indicates that they were laid down before the Hagan Basin itself was tilted, likely by the rise of the Sandia Mountains.

The Zia Formation

In July of 1859, on the eve of the Civil War, Captain John N. Macomb led a topographical engineer party on a expedition looking for a practical route from Santa Fe to the Great Basin. Much of the expedition followed the poorly charted Spanish Trail. While the expedition failed to find a practical route across the steep canyons of the Green and Colorado Rivers, the expedition included geologist John S. Newberry, who produced the first good geologic report on the heart of the Colorado Plateau. Alas, the outbreak of war delayed publication for a full sixteen years.

During the return to Santa Fe, Newberry visited Cañon de San Diego and the Valle Grande, becoming the first trained geologist to examine the Valles caldera. Near Jemez Pueblo, he noted "a number of buttes of white tufaceous rock ... undoubtedly Tertiary, [which] belong to the great series of freshwater deposits of Miocene and Pliocene age." Though his estimate of the age was slightly off — the beds were entirely Miocene in age — he was otherwise correct about their age and character.

Digital relief map of Zia Formation exposures in the
        Jemez Mountains
Relief map of the Jemez with Zia Formation outcroppings highlighted in red.

The Zia Formation is from 25 to 12 million years old, coinciding with the earliest period of rifting along the Rio Grande Rift. It is exposed in the southwest Jemez, in some places forming spectacular badlands. Some of the most prominent are in lower Canon de San Diego.

Zia Formation?
Zia Formation. Looking west from near 35 37.640N 106 43.466W

The Zia Formation here consists of very soft sandstone of nearly pure quartz, probably deposited as sand dunes about 13 to 20 million years ago. In this photograph, it is capped with younger gravels cemented with travertine and, as is usually the case when softer beds underlie more resistant beds, the softer Zia Formation forms cliffs and towers capped by the harder overlying sediments. Because the Zia Formation is so poorly cemented, its earliest name in the geologic literature was the Zia Sand.

Based on sedimentary structures in the formation, it is thought to have been deposited as sediments that were first transported south along the ancestral Jemez River, then carried east by the prevailing winds.

The Zia Formation can be traced further east, cropping out in the banks of Vallecitos Creek southwest of the village of Ponderosa.

Zia Formatio
Zia Formation. Looking east from 35 37.693N 106 42.095W

The Zia Formation is the white area to the left. The contrasting tan beds to the right are much younger terrace deposits of Vallecitos Creek, enriched with reddish sediments from the Paleozoic beds upstream. Zia Formation also crops out in the western slopes of Borrego Mesa.

Zia Formation
Zia Formation. Looking east from 35 37.693N 106 42.095W

The Zia Formation disappears in the area west of Cochiti Pueblo, buried under younger deposits of the Cochiti Formation. This reflects the migration of the active western margin of the Rio Grande Rift to the east.

The Tesuque Formation

Digital relief map of Tesuque Formation exposures in
        the Jemez Mountains
Relief map of the Jemez with Tesuque Formation outcroppings highlighted in yellow (Ojo Caliente Member) or red (all other members.)

The Tesuque Formation is the chief Santa Fe Group formation in the Espanola Basin. It lies atop the Abiquiu Formation and its age is between about 26 million and 11 million years old, making it contemporary with the Zia Formation. In places it is more than 3000 meters (10,000 feet) thick. Because of its great thickness, it was originally divided into five members, which (from oldest to youngest) are the Nambe Member, the Skull Ridge Member, the Pojoaque Member, the Chama-El Rito Member, and the Ojo Caliente Sandstone. More recent classifications have added the Cejita and Cuartales Members, both younger than the Pojoaque Member.

There is also overlap in age. It was recognized from the start that, while the Nambe, Skull Ridge, and Pojoaque Members form a fairly straightforward age sequence, the Chama-El Rito Member, with an age of about 17 to 11 million years, overlaps the Pojoaque Member and part of the Skull Ridge Member in time. It is identified as a distinct member because its sediments originate from the San Juan volcanic field to the north and northwest rather than the Sangre de Cristo Mountains to the east. The Ojo Caliente is mostly younger than the other members, but there is some overlap of the oldest Ojo Caliente beds with the earliest Pojoaque and Chama-El Rito beds.

To further complicate matters, the Tesuque Formation is also divided into Lithosome A and Lithosome B, and this division cuts across the more traditional division into members. The two lithosomes represent two distinct source regions for the sediments originally thought to originate in the Sangre de Cristo Mountains. Lithosome A is thought to be sediments eroded from the Sangre de Cristo Mountains to the east, and it consists of distinctly pinkish sediments rich in feldspar grains (arkose) and granite clasts deposited by a broad network of small streams flowing to the west. Lithosome B is thought to be sediments carried by larger rivers from the north and northeast, and it consists of tan sediments that contain singificant clay and a wide variety of different clasts. Thus, geological maps distinguish, for example, the Pojoaque Member of Lithosome A from the Pojoaque Member of Lithosome B.

The Santa Fe Group has proven to be a good fossil hunting ground. G.K. Gilbert visited San Ildefonso with the Hayden Survey in 1873 and found fossil mammal bones dating to the Pliocene. Some of these were sent to the great paleontologist Othniel Marsh, but his bitter rival E.D. Cope arrived at San Ildefonso the next year to collect a number of Miocene repile, bird, and mammal fossils. The American Museum of Natural History sponsored fossil digs from 1942 to 1965 that produced a vast collection of around 10,000 fossil fragments. Prehistoric mammals ranging from dire wolves (extinct relatives of gray wolves) and horses to camels, primitive elephants, and primitive hippopotamuses have been found in the various members of the Tesuque Formation. This forms much of the basis for its division into members.

Camel bone
          fragment
Commercial camel fossil from Santa Fe area

Camel footprints were discovered in an ash bed just west of Santa Fe Airport in the 1960s.

The Tesuque Formation is particular rich in fossils east and west of the highway from Santa Fe to Espanola. Palm fossils north of Santa Fe suggest that the elevation was still only 700m (2300') in the middle Miocene, much less than the current elevation of 1800m (6000').

The Nambe Member crops out close to the Sangre de Cristo Mountains, well east of the Jemez area, and we will not have much to say about it. It is mostly soft, pink sediments that weather to low hills.

Nambe Member
Nambe Member. Looking north from 35 54.851N 105 58.130W

Closer to the mountains, the Nambe Member starts to contain more coarse conglomerate beds.

Nambe Member
Nambe Member. Looking north from 35 57.217N 105 55.367W

This is a very common pattern in sedimentary beds. The coarser material is found nearer the source rocks and the finer material further away. This can be a very useful clue when reconstructing paleotopography for more ancient periods in the earth’s history, where sedimentary beds are all that are left of the original landscape. Thus, the tendency for beds of the Triassic (between 200 and 250 million years ago) to be coarser north of the Jemez than south of the Jemez is a clue that the sediments originated in a highlands near the Colorado border.

The Skull Ridge Member crops out as far west as the Pojoaque-Espanola highway and is notable for its large number of ash beds. These date to around 15 million years old, two million years older than the oldest known ash-producing eruptions in the Jemez region. Either the ash came from further afield, or it came from vents in the Jemez that are now buried under flows from later eruptions. The base of the Skull Ridge Member is defined in much of the region by a particularly distinctive ash bed.

A prominent feature of the Skull Ridge Member is the Red Wall, so named by paleontologists digging for fossils in the area.

Red Wall
The Red Wall. Looking west from 35 55.194N 105 57.965W

The Skull Ridge Member tends to form cliffs like these, probably because its numerous ash beds provide cementing silica.  Similar cliffs are impressive and readily accessible in the Camel Rock area along the main highway from Santa Fe to Espanola.

Skull
          Ridge Member
Skull Ridge Member in Camel Rock area. Looking east from 35 50.019N 105 59.048W

Numerous fossils have been collected from the Skull Ridge Member, and these help define the Barstovian faunal stage. A faunal stage is a distinctive combination of animal fossils found in beds of a fairly narrow range of age, not more than a few million years. The presence of ash beds that can be radiometrically dated has allowed paleontologists to fix the time frame of the Barstovian stage as from about 16.3 to 13.6 million years ago.

The Tesuque Formation is prominent in the Espanola Valley east of the Rio Grande, where it forms the spectacular badlands of Los Barrancos:

Los Barrancos
Los Barrancos. Looking north from near 35 52.897N 106 04.045W

Note that the beds tilt to the west (left). Geophysical data shows that the Rio Grande Rift in the Espanola area (the Espanola Basin ) is deepest to the west. These beds are mapped mostly as Lithosome A with Cuartales Member on top underlain by Pojoaque Member. To the west, the beds transition to Lithosome B, primarily of the Pojoaque Member. Numerous north-south faults cut across the beds, further complicating the geological picture.

Both of these members are exposed in road cuts along the highway from Pojoaque to Totavi.

Pojoaque Member,
          Lithosome A
Pojoaque Member, Lithosome A. 35 52.906N 106 06.958W

This is a thick siltstone interbedded with thin gravel layers.

Cejita Member,
          Lithosome B
Cejita Member, Lithosome B. 35 53.114N 106 06.417W

You can see the slightly redder color of Lithosome A versus the more tan color of Lithosome B.

The contrast between lithosomes within a single member is visible in the hills along the road from Otowi to Pojoaque.

Contact between Lithosomes A and B
Contact between Lithosome A and LIthosome B in Pojoaque Member. 35 53.172N 106 05.896W

Notice that the top of the hill is distinctly redder than the lower slopes. The geologic map confirms that this is the contact between Lithosome A (on top) and Lithosome B (at bottom) of the Pojoaque Member.

One of the best exposures of the Tesuque Formation for close examination is found in a road cut along Highway 285 southeast of Espanola.

Excellent exposure of Pojoaque Member

Road cut in Pojoaque Member, Lithosome A, Tesuque Formation. 35.984N 106.053W

The best location for appreciating the great thickness of the Tesuque Formation, though located just east of the Jemez region, is probably the south rim of the Santa Cruz river valley.

Lithosome A,
          Tesuque Formation, in splendor

Lithosome A, Tesuque Formation, north of Santa Cruz River. Looking north from 59.832N  105 54.366W

Almost the entire thickness of the cliffs north of the river is underlain by Lithosome A of the Tesuque Formation. The geologic map for this area does not divide this into members.

The Chama-El Rito Member

Along the western margin of the Rio Grande Rift, the Chama-El Rito Member makes up the bulk of the lower Tesuque Formation. It is prominent in the badlands north of the Santa Clara Mountains.

Chili badlands north of
          Santa Clara Mountains
Chili Badlands. Looking north from 36.035674N 106.24034W

The Canada del Almagre Fault runs roughly along the arroyo running north-northwest at the center of the panorama. Chama-El Rito Member is exposed west (left) of the fault at the foot of Lobato Mesa.

Chama-El Rito
          formatino north of Santa Clara Mountains
Chama-El Rito Member. Looking north from 36.035674N 106.24034W

To the right of the fault, on the downthrown side, the Chama-El Rito Member is overlain by Ojo Caliente Member. The Chama-El Rito Member reappears further east near the Chama River.

Chama-El
          Rito Member ormation near Chama River
Chama-El Rito Member near the Chama River. 36.1269438N 106.1670207W

The lenses of rounded clasts derived from the mountains to the north and northwest are characteristic of the Chama-El Rito Member and distinguish it from the Ojo Caliente Member, which is almost devoid of gravel.

The beds are covered further south by basalt of the the Santa Clara eruptive center, but reappear on the south flank of the Santa Clara Mountains.

Chama-El Rito
          formatino south of Santa Clara Mountains
Chama-El Rito Member. 36.030121N 106.2341054W

At Sierra Negra northeast of Abiquiu, the Chama-El Rito Member is exposed on the flanks of the mountain. Here a major fault has displaced the Chama-El Rito Member downwards on the right side of the fault, placing it alongside older Abiquiu Formation beds. We saw this picture earlier in the chapter.

Sierra Negra
Sierra Negra. Looking north from 36 12.231N 106 14.805W

The Chama-El RIto Member is the reddish beds to the right. Further west, thick beds of the Chama-El Rito Member are exposed on the flanks of Abiquiu Mesa.

Abiquiu Mesa

North of Abiquiu, there are tuff beds interpreted as remnants of an old tuff ring formed by localized volcanic activity within the Chama-El Rito Member. The tuff ring has not been dated.

Tuff ring
Remnants of tuff ring. Looking northeast from 36 13.361N 106 17.845W

The tuff ring remnants are the greenish-tan beds at left and right center in the photograph. Another set of tuff rings are found further to the east, on the road to El Rito.

El RIto tuff ring
Remnants of tuff ring. Looking west from 36 14.288N 106 11.537W

The darker rock is a basaltic tuff, formed from bits of silica-poor lava thrown out by an eruption. Basaltic tuffs do not form the vast outflow sheets you see with rhyolitic tuff, such as the Bandelier Formation. The silica-poor lava does not hold onto its gas content long enough to move far from the vent as suspended particles. This particular tuff was erupted in a small eruption some 15.3 years ago, based on potassium-argon dating of a basalt bomb embedded in the tuff.

Here's a closer look.

El RIto tuff
          ring
Mafic tuff. 36.2403711N 106.1922591W

This portion of the ring is composed of basaltic cinder and ash cementing together large broken fragments of basalt.

Other beds are much better sorted.

El RIto tuff ring
Mafic tuff. 36.2418693N 106.195026W

The color shift is fairly abrupt, perhaps representing the point where hydromagmatic explosions gave way to pure cinder.

An arroyo cuts through the tuff ring, exposing veins of fluorite and barite.

Small Fry
          fluorite
Mineral vein. 36.2424011N 106.1910985W

Fluorite, CaF2, and barite, BaSO4, are the principal ores of fluorine and barium respectively, as well as having minor uses in their raw form. Both elements are incompatible elements, and the veins here were deposited by hydrothermal fluids, possibly associated with a nearby fault.

The Ojo Caliente Sandstone

Fourteen million years ago, the climate of the Rio Grande Rift to began to cool. The many small streams that had brought sediments into the Rift started drying up and a field of dunes began to creep across the Rift. This dune field is preserved in the rock record as the Ojo Caliente Member of the Tesuque Formation.

I've mentioned that most of the members of the Tesuque Formation are difficult for a non-expert to distinguish in the field. However, the Ojo Caliente Sandstone stands out from the rest of the Santa Fe Formation as being even less likely to be well cemented and in being composed of relatively clean quartz grains. It is a difficult unit to accurately date; there are no ash beds for radiometric dating, and it is almost devoid of fossils. However, it is thought to be between about 10 and 13.5 million years old.

The formation is exposed throughout the Medanales area.

Ojo Caliente
        Sandstone ridge
Ridge of Ojo Calienta Sandstone of the Tesuque Formation. 36 11.287N 106 13.898W

In a number of locations, the Ojo Caliente Sandstone has been unusually well cemented along faults that allowed mineral-rich groundwater to circulate through the sandstone. The hardened sandstone stands out in stark relief when the surrounding softer sediments are eroded away. Cerrito Blanco, west of Abiquiu, is a an excellent example of such a sedimentary dike.

Cerrito Blanco
Cerrito Blanco. 36 12.867N 106 19.583W

Cerrito
          Blanco
Cerrito Blanco. 36 12.988N 106 19.368W

Hills of Ojo Caliente Member are prominent east of Lobato Mesa and west of the Rio Chama.

Santa Fe sediments
          east of Lobato Mesa

The foreground terrain is underlain by Chama-El Rito member. A fault runs along the west side of the hills and may account for the hills being better cemented than most Ojo Caliente Member.

Ash beds are visible at the southern edges of the small hills right and left of center. These are marked on the geologic map for the area, but not further identified or described.

The Ojo Caliente extends from here to the Rio Chama and beyond.

Ojo
          Caliente Member along Lemitas Trail
Ojo Caliente along Lemitas Trail. 36.121871N 106.1845773

The Ojo Caliente in this area is often well cemented, forming resistant beds.

Ojo Caliente
          Member along Lemitas Trail
Ojo Caliente along Lemitas Trail. 36.1255376N 106.1742324W

Ojo
          Caliente Member along Lemitas Trail
Ojo Caliente along Lemitas Trail. 36.1255376N 106.1742324W

This area also has some interesting sedimentary structures, such as this "pillow" structure.

Ojo Caliente
          Member along Lemitas Trail
"Pillow" stuctures in Ojo Caliente along Lemitas Trail. 36.1227921N 106.1798249W

Or these nests of concretions.

Ojo
          Caliente Member along Lemitas Trail
Concretions in Ojo Caliente along Lemitas Trail. 36.121871N 106.1845773

There are numerous "fins" of harder rock within the softer bed, all striking north to north-northeast.

Ojo Caliente
          Member along Lemitas Trail
Ojo Caliente "fin" along Lemitas Trail. 36.1228178N 106.1804238W

Ojo Caliente Sandstone also underlies much of the northwestern slopes of Black Mesa.

Hoodoos in Ojo
        Caliente beneath Black Mesa
Hoodoos in Ojo Caliente. 36 9.221N 106 5.701W

Here the soft nature of the beds is clear. This looks a lot like the Zia Formation, doesn't it? Similar environments produce similar rocks; but the Zia Formation appears to be significantly older than the Ojo Caliente Sandstone.

Nearby is yet another sedimentary dike.

Sedimentary dike
Cerrito de la Baca, a sedimentary dike north of Black Mesa. 36 7.146N 106 7.594W

There are excellent exposures of Ojo Caliente Member in Abiquiu Canyon. One such exposure is where the road turns to ascend El Alto.

Ojo Caliente Member in Abiquiu Canyon

Ojo Caliente Member in Abiquiu Canyon. 36 09.992N 106 21.486W

This must be admired from a distance, as the entire canyon is private property and visitors are restricted to the roadway.

Further down the canyon, there is another exposure of the Ojo Caliente Member, with impressive cross bedding.

Ojo Caliente Member in Abiquiu Canyon

Ojo Caliente Member in Abiquiu Canyon. 36 10.485N 106 20.713W

This exposure just screams eolian. Judging from the satellite photo, this may be a borrow pit, from which sand was extracted for road building or construction work.

Local exposures of Ojo Caliente Member are also found in the west rim of the Valles caldera, atop Abiquiu Formation and beneath the oldest volcanic rocks. These are the westernmost exposures of the unit.

Ojo Caliente
          Member in Valles west rim
Ojo Caliente Member. 35 53.578N 106 40.262W

You can see in the photograph how deeply eroded this small outcrop is. The exposure is poorly consolidated and uncemented, with loose sand everywhere.  A more substantial exposure is found further west along State Road 126 towards Fenton Lake.

Ojo Caliente Member

Ojo Caliente Member. 35 53.074N 106 41.019W

The bulk of this road cut is light sandy sediments of the Ojo Caliente Member. The area at upper left exposes younger formations that we will revisit later in this book.

Similar exposures are found in many places in the caldera west rim. Some are difficult to assign to particular units of the Santa Fe Group, and are mapped simply as undivided Tertiary sediments. The "undivided" means that the Tertiary beds have not been divided into separate units on the map. An example crops out along Forest Road 376 to San Antonio Springs.

Undivided Tertiary sediments

Undivided Tertiary sediments in caldera west rim. 35 53.757N 106 39.860W

Such sediment beds make up a fair part of the west caldera wall, and this outcrop must be at least five million years old. By appearance, it could be either Ojo Caliente Member or Zia Formation.

Marker Beds

The study of otherwise nondescript formations, such as the Tesuque Formation, is made easier by the presence of marker beds. These are beds in the formation that stand out in some way, so that they can be traced over long distances. We've seen an example of a marker bed already: the thin limestone bed near the top of the De Chelly Sandstone, near its gradation into the Glorieta Formation. Several marker beds are found in the Tesuque Formation.

Many of these are ash beds, such as the 285 Road Ash of the Skull Ridge Member.

285 Road
          Ash

285 Road Ash of the Skull Ridge Member. 35.937N 106.022W


285 Road Ash

Upper 285 Road Ash of the Skull Ridge Member. 35.937N 106.022W

The ash bed is the white layer at the top of the cliff in the first photo and the near hills in the second. It’s a bed rich in volcanic ash, which gives it its light color. The volcanic ash contains potassium feldspar crystals which have been radiologically dated to 15.1 million years old. This is older than the Jemez volcanic field, and the precise source of the ash is unknown, though it is suggestive that its age is close to that of the earliest giant caldera eruptions of the Yellowstone hot spot. Most of the ash beds of the Skull Ridge Member are thought to have originated either from the Yellowstone hot spot or from a cluster of calderas in southern Nevada. The 285 Road Ash is not the most prominent ash bed in the Skull Ridge Member, but it is probably the most accessible, being right off a major highway.

Another kind of marker bed is found between the Tesuque Formation and overlying terrace gravels throughout the area west and north of Espanola. (Terrace gravels are deposits left by a river when it cuts a deeper channel; we'll have more to say about them later in the book.) This marker bed, which is particularly well exposed at Arroyo Largo, consists of quartzite gravel.

Quartzite gravel marker bed

Quartzite gravel marker bed. 36 01.457N 106 05.527W

This bed is very distinctive, consisting of well rounded clasts of quartzite typical of the Tusas Mountains to the north. It mark a geologically brief interval sometime in the last two million years when the ancestral Rio Grande and its tributaries in the Espanola area carried a heavy load of gravel from the mountains to the north. Similar beds are found as far north as Abiquiu and as far south as Cochiti and may be correlated.

The Cerro Conejo Formation

Digital relief map of Cerro Conejo Formation exposures
        in the Jemez Mountains
Relief map of the Jemez with Cerro Conejo Formation outcroppings highlighted in red.

The upper beds of the Zia Formation, as originally defined, were assigned to the Cerro Conejo Member. However, there turns out to be a significant hiatus at the base of the Cerro Conejo, amounting to perhaps as much as 1.6 million years of missing beds. This led geologists to promote the Cerro Conejo to formation rank.

The Cerro Conejo is about 14.6 to 9.5 million years old and represents a time of transition in the western Santo Domingo Basin. The eolian deposits of the Zia Formation were beginning to give way to fluvial deposits from streams descending off the Colorado Plateau. For this reason, some geologists assign the Cerro Conejo Formation to the middle Santa Fe Group, as transitional between playa deposition and river deposition.

The formation is exposed mostly on tribal lands, but there is a good exposure in a highway roadcut between Bernalillo and San Ysidro.

Cerro Conejo
          Formation

Cerro Conejo Formation. 35.3974N 106.6396W

A number of younger formations assigned to the Santa Fe Group represent beds deposited as the Rio Grande became established as a through-running river. We'll examine these upper Santa Fe Group formations later in the book.

Next page: And, finally, we get to the birth of the Jemez volcanic field

Copyright ©2014 Kent G. Budge. All rights reserved.