<<

Developing a 3-D subsidence model for the late Paleozoic Taos trough in northern

By

Nur Uddin Md Khaled Chowdhury, MS

A Dissertation

In

Geology

Submitted to the Graduate Faculty of Texas Tech University in Partial Fulfillment of the Requirements for the Degree of

DOCTOR OF PHILOSOPHY

Approved

Dustin E. Sweet Chair of Committee

George B. Asquith

James E. Barrick

Seiichi Nagihara

Mark Sheridan Dean of the Graduate School

December, 2019

Copyright 2019, Nur Uddin Md Khaled Chowdhury

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

ACKNOWLEDGEMENTS

Numerous individuals have been instrumental in the completion of this dissertation. First, and foremost, I would like to sincerely thank my adviser, Dr. Dustin E.

Sweet, for his guidance, unwavering support, and ultimately, for making this endeavor thoroughly exciting and enjoyable. I also sincerely thank my committee members Dr.

George B. Asquith, Dr. James E. Barrick, and Dr. Seiichi Nagihara.

I would like to thank Annabelle Lopez and New Mexico Bureau of Geology &

Mineral Resources for providing me access to cores and cuttings repository and providing me available resources including well log data, well reports and strip logs relevant to this study. I am thankful to Neuralog Inc. for providing me free access to Neuralog software.

Funding for this research was generously supported by John Emery Adams

Memorial Scholarship and Concho Resources Inc. Scholarship from West Texas

Geological Society (WTGS), student research grants from Geological Society of America

(GSA), and grants-in-aid from the southwest section of American Association of

Petroleum Geologists (AAPG). I am indebted to the department of Geosciences at Texas

Tech University.

Finally, I would like to thank my loving family, especially my parents and siblings for their unwavering support and continued source of inspiration.

ii

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

TABLE OF CONTENTS

ACKNOWLEDGEMENTS ...... ii

ABSTRACT ...... vi

LIST OF TABLES ...... viii

LIST OF FIGURES ...... ix

1. INTRODUCTION AND GEOLOGIC BACKGROUND ...... 1 RATIONALE FOR THIS STUDY ...... 1 Objectives: Basin subsidence mechanism - flexure vs. wrenching ...... 1 Significance: Taos-Rainsville basin model and implications for Late Paleozoic plate tectonic model of southwestern North America ...... 2 GEOLOGIC BACKGROUND ...... 3 Late Paleozoic southwestern USA and Ancestral ...... 3 Taos and Rainsville troughs ...... 5 of the greater Taos trough region ...... 8

2. SUBSURFACE STRATIGRAPHY AND FACIES DISTRIBUTION OF THE LATE PALEOZOIC RAINSVILLE TROUGH ...... 14 METHODS AND MATERIALS ...... 15 IDENTIFYING LITHOLOGY AND BUILDING STRATIGRAPHIC SECTIONS ...... 17 Arroyo Peñasco ...... 20 ...... 21 Porvenir Formation ...... 23 ...... 24 Sangre de Cristo Formation ...... 25 Yeso Formation ...... 27 Glorieta Formation ...... 28

3. AGE CONSTRAINTS ON UPPER PALEOZOIC STRATA IN THE GREATER TAOS TROUGH REGION ...... 36 METHODOLOGY RATIONALE ...... 37 DATA ANALYSIS AND AGE CONTROLS ...... 38 Espiritu Santo Formation ...... 38 Tererro Formation ...... 40 iii

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Sandia Formation ...... 42 Porvenir Formation ...... 44 Alamitos Formation ...... 46 Sangre de Cristo Formation ...... 48 Yeso and Glorieta formations ...... 50 SUMMARY AND STRATIGRAPHIC CORRELATIONS ...... 53

4. - CLIMATE-SENSITIVE FACIES FROM LOW LATITUDE RAINSVILLE TROUGH ...... 58 LATE PALEOZOIC CLIMATIC SETTING ...... 59 CLIMATE INDICATORS: COAL AND EVAPORITE AS PALEOCLIMATIC PROXY ...... 62 METHODS AND MATERIALS ...... 65 STRATIGRAPHIC RECORDS OF PENNSYLVANIAN AND PERMIAN PALEOCLIMATE ...... 66 Pennsylvanian strata ...... 66 Permian strata...... 68 IMPLICATIONS FOR REGIONAL PALEOCLIMATE ...... 69 UTILIZATION OF CLIMATE-SENSITIVE FACIES FOR CORRELATIONS ...... 73

5. 3-D BASIN SUBSIDENCE MODEL...... 76 METHODS AND MATERIALS ...... 76 Building individual stratigraphic section for construction of subsidence curves ...... 76 Obtaining parameters for backstripping: decompaction factor and porosity ...... 77 Obtaining parameters for backstripping: assessing age ...... 78 Obtaining parameters for backstripping: water depth and sea level changes ...... 79 The Process of Backstripping: Iteratively Removing the Stratigraphic Record ...... 79 RESULTS AND DISCUSSIONS ...... 80 Basin aspect ratio ...... 80 Tectonic subsidence analysis of the Taos-Rainsville trough ...... 81 Rate of basin subsidence through time ...... 87 3-D basin model for the Rainsville trough ...... 89 Taos-Rainsville basin model ...... 94 ARM implications ...... 97

iv

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

6. CONCLUSION ...... 100

REFERENCES CITED ...... 103

APPENDICES ...... 119

A. MATLAB code for lithologic plot ...... 119

B. MATLAB code for producing subsidence curves using backstripping technique ...... 122

C. MATLAB code for producing 3-D structure map ...... 126

v

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

ABSTRACT

During the late Paleozoic, southwestern Laurentian craton experienced deformation consisting of -cored uplifts and adjacent sedimentary basins known as the ancestral Rocky Mountains (ARM). Due to unusual inland position of these uplifts from known plate boundaries and overprinting by younger tectonic activities, their tectonic model is not well understood and contrasting plate tectonic models are proposed for the orogenic event. Understanding individual basins is necessary to better decipher tectonic model(s) of the greater ARM.

The Taos-Rainsville trough (also known as the Rowe-Mora basin), located in north-central New Mexico, is one of the core ARM basins. Structural basin models for the Taos-Rainsville trough are also contradictory. One model predicts a flexural basin model where the lithosphere was loaded from the west, whereas the other model proposes a wrench basin model. Tectonic subsidence curves should differentiate between the two proposed models due to the different rates of subsidence that result from these two basin mechanisms. The main goal of this study is to construct numerous 1-D backstripped tectonic subsidence curves that, when time correlated, will allow assembling the 3-D basin subsidence architecture necessary to assess the best basin model. Evaluating contrasting basin models for the Taos-Rainsville trough will provide insight into regional strain patterns, thus helping resolve greater ARM tectonic models.

Petrophysical logs tied to outcrops along the western margin of the Rainsville trough were utilized to produce composite stratigraphic columns. A chronostratigraphic framework was built based by correlating lithofacies, regional biostratigraphic data, and vi

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

climate-sensitive facies. Tectonic subsidence analysis and basin modeling of the greater

Taos-Rainsville trough suggests that subsidence of the Rainsville trough was likely

controlled by the thrust loaded flexure of the El Oro-Rincon uplift. Subsidence of the

Taos trough was initiated by a flexural load during the Early Pennsylvanian. However,

during the Desmoinesian, the Picuris-Pecos fault initiated as a strike-slip system with significant dip slip and the basin subsidence history compares better to known transtensional systems. Thus, the greater Taos trough basin model is most consistent with a wrench basin where sediment accumulation is facilitated by both thrust loading and transtension.

vii

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

LIST OF TABLES

2.1: Wells and corresponding petrophysical logs used in the study……….…………17

5.1: Depositional age ranges of lithologic units……………………….……...………79

viii

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

LIST OF FIGURES

1.1 Late Paleozoic tectonic elements of the greater ancestral Rocky Mountains...... 4

1.2 Late Paleozoic tectonic elements of the greater Taos trough area ...... 7

1.3 Regional stratigraphy, nomenclature, and stratigraphic relationships in the greater Taos trough region ...... 9

2.1 Map of the study area with location of wells and surface sections used in this study ...... 15

2.2 A portion of the scanned log from well-1, digitized version of the log and interpreted lithology using well log data ...... 19

2.3 Interpreted lithology and constructed lithologic column from well log data using MATLAB programming ...... 20

2.4 Isopach map of Arroyo Peñasco Group ...... 21

2.5 Isopach map of Sandia Formation ...... 22

2.6 Isopach map of Porvenir Formation ...... 24

2.7 Isopach map of Alamitos Formation...... 25

2.8 Isopach map of Sangre de Cristo Formation...... 26

2.9 Isopach map of Yeso Formation ...... 27

2.10 Isopach map of Glorieta Formation ...... 29

2.11 An example of picking formation boundary and correlation of wells ...... 31

2.12 Cross-sections along east to west transects across the Rainsville trough ...... 32

2.13 Cross-section along north-south transects of the Rainsville trough...... 33

2.14 Isopach maps of (A) Pennsylvanian and (B) lower Permian strata in the Rainsville trough area...... 35

3.1 The greater Taos trough is divided into three broad regions based on paleogeographic affinity ...... 38

3.2 Biostratigraphic data from the Espiritu Santo Formation from all three regions of the study area ...... 39

ix

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

3.3 Biostratigraphic data from the Mississippian Tererro Formation ...... 41

3.4 Biostratigraphic data from the Pennsylvanian Sandia Formation...... 43

3.5 Biostratigraphic data collections from the Porvenir Formation...... 45

3.6 Biostratigraphic data assemblages from the Alamitos Formation ...... 48

3.7 Biostratigraphic data assemblages from the Sangre de Cristo Formation ...... 50

3.8 Summary of biostratigraphic age distributions assembled formation-wise from Figures 3.2 through 3.7 ...... 52

3.9 Basin-wide stratigraphy of the Taos trough region...... 55

3.10 Reconstructed basinal subsurface stratigraphy of the Rainsville trough ...... 56

4.1 Relation between sedimentological response and climate change. A. Probability of clastic sediment flux in response to climate wetness; B. Probability of chemical sediment formation in response to climate wetness ...... 64

4.2 Wells used in this study to decipher paleoclimatic information ...... 65

4.3 Plots of coals versus evaporite in different wells from the Rainsville trough using MATLAB programming ...... 67

4.4 Climate-sensitive lithologies from different wells ...... 69

4.5 Glacial episodes in Gondwana during the late Paleozoic. Summary of Permo- paleoclimatic interpretation using lithological indicator from the Rainsville trough...... 72

4.6 Utilization of climate-sensitive facies in constructing subsurface stratigraphy of the Rainsville trough ...... 74

5.1 Length and width of the Rainsville trough, Taos trough, and the greater Taos trough are plotted to compare with basin aspect ratios of flexural and strike-slip basins from known settings ...... 81

5.2 Backstripped total (A) and tectonic (B) subsidence curves from the Taos- Rainsville trough ...... 84

5.3 Tectonic subsidence curves from different ARM basins were plotted with reference data from known tectonic settings to compare between flexure due to static loads versus flexure due to migrating loads ...... 86

x

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

5.4 Chart showing the relationship of tectonic subsidence rate of the Taos- Rainsville trough during Pennsylvanian and early Permian ...... 89

5.5 Structural development of the Rainsville trough through time ...... 91

5.6 Structure maps of Missourian-Virgilian Alamitos Formation, Wolfcampian Sangre de Cristo Formation and Leonardian Yeso Formation...... 92

5.7 3-D structure maps of different lithostratigraphic units through time, generated using MATLAB programming ...... 93

5.8 West to east cross-sectional view illustrating the evolution of the Taos- Rainsville trough through time ...... 97

5.9 The Pennsylvanian and early Permian tectonic subsidence of Taos-Rainsville trough are compared with subsidence in other ARM basins ...... 99

xi

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

CHAPTER 1 INTRODUCTION AND GEOLOGIC BACKGROUND

RATIONALE FOR THIS STUDY

Objectives: Basin subsidence mechanism - flexure vs. wrenching

During late Paleozoic, southwestern Laurentia experienced intra-plate

deformation known as the ancestral Rocky Mountains (ARM) (Fig. 1.1; Kluth and

Coney, 1981; Kluth, 1986; Ye et al., 1996; Dickinson and Lawton, 2003). The greater

Taos trough is one of numerous depocenters that record ARM deformation (Fig. 1.2; e.g.,

Miller et al, 1963; Casey, 1980, 1981; Kluth and Coney, 1981; Kluth, 1986; Ye et al.,

1996; Baltz and Myers, 1999). Two different structural basin models have been proposed

for the Taos trough. Soegaard (1990) proposed a west-loaded, flexural basin where sediment was sourced primarily from rocks to the west. In contrast, Baltz and Myers (1999) suggested a basin model where intrabasinal strain is manifested as thrusts with Precambrian-cored hanging walls and a strike-slip fault bounding the basin to the west, potentially a wrench basin. Tectonic subsidence curves should differentiate between the two proposed models due to the different rates of subsidence that results from these two basin mechanisms (Xie and Heller, 2009). The main goal of this study is to construct numerous 1-D backstripped tectonic subsidence curves that when time correlated will allow assembling the 3-D basin subsidence architecture necessary to assess the correct basin model. Deciphering between a thrust loading versus a wrench basin mechanism for the Taos trough provides insight into regional strain patterns that should bear on the proposed tectonic models.

1

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Significance: Taos-Rainsville basin model and implications for Late Paleozoic plate tectonic model of southwestern North America

ARM deformation is believed to be a response to collision between

supercontinents Laurentia (North America) and Gondwana (South America-Africa) along the Appalachian-Ouachita-Marathon belt mostly based on the synchronous timing of these two events (Kluth and Coney, 1981; Algeo, 1992; Dickinson and Lawton, 2003;

Dickerson, 2003). This model predicts significant wrenching during formation of ARM basins (e.g. Algeo, 1992; Dickinson and Lawton, 2003). However, the high angle between the structural trend of the ARM compared to the Ouachita-Marathon belt does not follow the curvilinear deformation model for mountain belts and raises questions on whether collision along the Ouachita-Marathon belt is an appropriate plate tectonic model. Alternatively, Ye et al. (1996) proposed flat-slab subduction along the southwestern margin of North America, as the tectonic driver for ARM deformation, analogous to the Cenozoic Laramide Rocky Mountains deformation. According to this model, ARM basins should experience flexural subsidence due to northeast-southwest shortening (e.g., Barbeau, 2003). This model is controversial as no arc volcanism and associated shortening has been identified along the late Paleozoic southwest Laurentian margin (Kluth, 1998; Dickinson, 2000; Dickinson and Lawton, 2001, 2003). The NW-SE structural trend of the deformation appears to follow lineaments in the Precambrian basement (Marshak et al., 2000), but is also consistent with NE-SW shortening resulting from a subduction zone (Ye et al., 1996) or oblique translation (Leary et al., 2017) along the southwest margin of the Laurentia. Hence, plate tectonic models of the ARM remain debated and to distinguish the most apt model, mechanisms of basin formation need to be

2

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

understood. Deciphering between a thrust-loading versus a wrench-basin mechanism for the Taos-Rainsville trough provides insight into regional strain patterns that should bear on the proposed tectonic models.

GEOLOGIC BACKGROUND

Late Paleozoic southwestern USA and Ancestral Rocky Mountains

Structural elements of late Paleozoic ancestral Rocky Mountains (ARM), located within the North American craton and about 1500 km inland from any coeval plate boundary, are a series of Precambrian-cored uplifts (Fig. 1.1; Kluth and Coney, 1981;

Kluth, 1986; Ye et al., 1996; Dickinson and Lawton, 2003). These largely northwest- southeast trending uplifts are separated by structurally deep basins that are filled with dominantly clastic sediments, primarily coarse arkosic sedimentary strata, and nonconformably lie atop or adjacent to Precambrian basement rocks (Mallory; 1972;

Rascoe and Baars, 1972; McKee et al., 1975). Basins of the central part of the ARM

commonly form clastic wedges adjacent to Precambrian-cored uplifts, such that the wedge generally thickens asymmetrically towards a contiguous region of contemporaneous basement (Mallory; 1972; Rascoe and Baars, 1972; McKee et al.,

1975). Structural relief of as much as 12 km has been documented on some ARM structures (McConnell, 1989).

During the Pennsylvanian, paired basins and uplifts formed synchronous to one another indicating that the Precambrian-cored fault blocks were being uplifted and eroded, and that the paired basins were also subsiding at a comparable rate with sediment fill shed from those uplifts (Ye et al., 1996). However, in the core of the ARM during the

3

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

early Permian, sedimentation locally buried (up to 1 km) Precambrian-cored uplifts forming a buttress (Moore et al., 2008) and the provenance signal shows an increasing far-field source, rather than being entirely locally sourced (Soreghan et al.,

2012; Sweet et al., 2013). These relationships are consistent with the two-phase subsidence model for ARM basins as proposed by Soreghan et al. (2012): 1)

Precambrian-cored uplift and simultaneous adjacent basin development in the

Pennsylvanian, and 2) eiporogenic subsidence of yoked basins and uplifts originating from an isostatic adjustment of a high-density crustal root in the early Permian.

Figure 1.1 Late Paleozoic tectonic elements of the greater ancestral Rocky Mountains. Adapted from Baltz and Myers (1999), Sweet and Soreghan (2010), Sweet et al. (2015) and Leary et al. (2017). Location of Desmoinesian and Wolfcampian paleoequators are 4

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

from Scotese (1997). Abbreviations for uplift: A = Apishapa; AFR = Ancestral ; AW = Amarillo Wichita; CBP = central basin platform; DZ = Defiance-Zuni; F = Florida; P = Pedernal; Pi = Piute; SG = Sierra Grande; U = Uncompahgre; UP = Ute Pass uplift. Blue box indicates location of figure 1.2.

Taos and Rainsville troughs

In northern New Mexico, tectonic elements of the ARM consist of the Taos and

Rainsville troughs, also referred to as the Rowe-Mora basin, bounded by Precambrian-

cored uplifts to the west and east, the southern Uncompahgre uplift and Sierra Grande

uplift, respectively (Fig. 1.2; e.g., Sutherland, 1972; Casey et al., 1979; Soegaard, 1990;

Baltz and Myers, 1999; Sweet and Watters, 2015). Some workers portray Pennsylvanian-

aged thrusts separating the western Taos trough from the Rainsville trough (Baltz and

Myers, 1999; Sweet and Watters, 2015). Stratigraphic thinning of lower Permian strata

across Precambrian basement in the hanging wall of one of those thrusts, the El Oro-

Rincon uplift, and thickening of Pennsylvanian strata east of that uplift indicates it was likely a positive feature during the Pennsylvanian (Baltz and Myers, 1999). Thus, in this dissertation, Taos and Rainsville trough are considered as two separate basins, and the term ‘Taos trough’ refers to the western portion of the Rowe-Mora basin, ‘Rainsville trough’ indicates the eastern part of the basin, and ‘greater Taos trough’ is utilized to refer to the overall combined Taos-Rainsville trough area.

Several paleomaps have displayed the greater Taos trough connecting northward with the Central trough (e.g. Sutherland, 1972; Kluth and Coney, 1981; Hoy and Ridgway, 2002; Sweet and Soreghan, 2010). Based on well-log data and isopach maps, Baltz and Myers (1999) argued that the Cimarron arch was a positive relief feature during the Pennsylvanian that separated the greater Taos trough from the Central 5

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Colorado trough. During the late Paleozoic, the region would have been located near

equatorial latitudes (Fig. 1.1; Scotese, 1997).

Structural overprinting of the ARM occurred during the Laramide where thrust faulting that involved Precambrian basement related to Laramide shortening bisected the greater Taos trough region such that the western portion of the region now resides in the hanging wall of the Laramide uplift (Yin and Ingersoll, 1997) and forms the core of the Sangre de Cristo Mountains in north-central New Mexico. Most Permian-aged

strata has been eroded from the hanging wall since the Late -early Tertiary

Laramide uplift (Bachman, 1953; Bachman and Dane, 1962; Miller et al., 1963).

6

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 1.2 Late Paleozoic tectonic elements of the greater Taos trough area. Green lines indicate faults with inferred Pennsylvanian-Permian movement. Adapted from Baltz and Myers (1999), and Sweet et al. (2015). Abbreviation: PPF = Picuris-Pecos fault (slip- sense from Cather et al., 2006; 2011 and Wawrzyniec et al., 2007).

7

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Stratigraphy of the greater Taos trough region

A generalized regional stratigraphic column of the study area is shown in Figure

1.3.

Mississippian

Arroyo Peñasco Group: Armstrong (1955) first assigned the name Arroyo

Peñasco to this group exposed in the Sangre de Cristo Mountains. Baltz and Read (1960) subdivided the group into the Espiritu Santo Formation and the Tererro Formation.

Espiritu Santo Formation: The Espiritu Santo Formation consists of a basal unit (Del Padre Sandstone Member) and an overlying carbonate unit (Baltz and

Read, 1960). Sandstone and carbonate units of the Espiritu Santo Formation overlie

Precambrian basement rock in the greater Taos trough region (Baltz and Read, 1960;

Armstrong and Mamet, 1979; Miller et al., 1963), though this formation is locally absent due to sub-Pennsylvanian erosion or non-deposition. Armstrong and Mamet (1979) assigned an Early Mississippian, late Tournasian, age for the Espiritu Santo Formation via foraminiferal zonation of microfossils collected in the region.

8

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 1.3 Regional stratigraphy, nomenclature, and stratigraphic relationships in the greater Taos trough region. Straight red lines refer to conformable relations between stratigraphic units whereas wavy red lines indicate unconformable relations. White areas

9

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

denote hiatuses. Modified after Baltz and Myers (1999). Abbreviations: Kin. = Kinderhookian, Osa. = Osagean, Mer. = Meramecian, Che. = Chesterian.

Tererro Formation: Baltz and Read (1960) assigned the name Tererro Formation to the unit that unconformably overlies the Espiritu Santo Formation. They divided this formation into three members, in ascending order: (1) Macho Member, (2) Manuelitas

Member, and (3) Cowles Member. The Macho Member comprises of pebble- to boulder- sized brecciated . These mostly angular to slightly rounded blocks are randomly oriented in finer silt- to sand-sized limestone particles. The Manuelitas Member consists of calcarenite, pebble-cobble sized conglomerate limestone, and crystalline limestone.

This member also contains some oolite grains, and sandy to silty limestone. The Cowles

Member has two lithologic units: a basal, silty to sandy calcarenite, and upper siltstone

unit that comprises of inter-bedded calcarenite, thin crystalline limestone, and occasional

marl. Based on mostly microfossil and sparse megafossil studies, the age for the Tererro

Formation is assigned to the Late Mississippian (Armstrong and Mamet, 1967, 1979; and

Baltz and Read, 1960).

Pennsylvanian

Sandia Formation: C. L. Herrick (1900) first used the term 'Sandia series' for a suite of rocks in the Sandia and . The Sandia Formation is composed mainly of with inter-bedded sandstone of variable thicknesses, pebbly conglomerate, and thin limestone beds (Baltz and Myers, 1999). The first occurrence of in the Taos trough occurs in the upper part of this formation. The Sandia

Formation is widespread throughout the southern Sangre de Cristo Mountains and unconformably overlies either the Mississippian Arroyo Peñasco Group or crystalline

10

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019 basement rocks. This formation thickens up to several hundred meters northward from a few centimeters at southern exposures near Bernal (Baltz and Myers, 1999). One of the very common characteristics of this formation is that rocks are heterogeneous both vertically and laterally indicating a complex depositional environment.

Madera Group: Keyes (1903) first used the name ‘Madera’ to describe a limestone unit that overlies the Sandia Formation in the of central New

Mexico. The limestone-bearing conformably overlies the Sandia

Formation. The unit consists of two formations in the southern portion of the Sangre de

Cristo Mountains: the Porvenir Formation and the overlying Alamitos Formation. (Keyes,

1903; Baltz and Myers, 1984; Baltz and Myers, 1999).

Porvenir Formation: Baltz and Myers (1984) first applied the name Porvenir

Formation and assigned a type section at Johnson Mesa Road. Baltz and Myers (1984) divided this formation into three intergrading lateral facies: (1) a southern, dominantly carbonate facies, (2) a northern, sandstone-shale-limestone facies, and (3) a northwestern, shale dominated facies. The Porvenir Formation locally crops out throughout the southern

Sangre de Cristo Mountains except atop places of Paleozoic uplift (Baltz and Myers,

1999).

Alamitos Formation: Sutherland (1963) assigned the name Alamitos Formation to describe Middle and Upper Pennsylvanian rocks that overly the in the southern Sangre de Cristo Mountains. (The term Flechado has since been supplanted in favor of the Porvenir Formation). The contact with the underlying Porvenir Formation is probably unconformable in most areas (Baltz and Myers, 1999). The Alamitos 11

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Formation consists of interbedded shale, limestone, sandstone and granule to pebble conglomerate. Shale is generally the dominant rock type, but is highly variable locally

(Baltz and Myers, 1999). A characteristic lithologic type is nodular to irregularly lenticular limestone inter-bedded with silty-marly shale (Baltz and Myers, 1999).

Lower Permian

Sangre de Cristo Formation: The Sangre de Cristo Formation generally unconformably overlies the Alamitos Formation and Precambrian basement rocks, such as on the Sierra Grande uplift. However, locally the contact between the Alamitos

Formation and the Sangre de Cristo Formation may be conformable in the greater Taos

Trough region (Baltz and Myers, 1999). The Sangre de Cristo Formation usually consists of nonmarine shale interbedded with arkosic sandstone, silt, the occasional pebbly conglomerate, and a few limestone beds (Baltz and Myers, 1999).

Yeso Formation: The Yeso Formation is predominantly composed of sandstone with minor limestone and evaporite beds and is inferred as non-marine to shallow-marine deposition (Lucas et al., 2015; Mack and Dinterman, 2002). The contact with the underlying Sangre de Cristo Formation is generally inferred as conformable in the region

(Baltz and Myers, 1999). Biostratigraphic data for this formation within the greater Taos trough region is not available, but a Leonardian age is assigned based on collections from south of the Taos-Rainsville trough.

Glorieta Sandstone: The was previously considered as a member of the San Andres Formation, however, in recent work the unit is denoted as a separate formation (i.e., Woodward, 1987; Colpitts, 1989; Lucas et al., 1999; Lucas and 12

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Zeigler, 2004; Mack and Dinterman, 2002; Kues and Giles, 2004, Connell, 2008). This

unit is generally composed of fine-grained, well-sorted, yellowish brown colored quartzose sandstone. Based on sedimentologic characteristics, a dominantly shallow marine or eolian origin is suggested for these sandstone-dominated strata (Lucas et al.,

2013). Biostratigraphic data are not available from the Glorieta Sandstone in the study area. Stratigraphically, the Glorieta Sandstone lies atop the early-middle Leonardian Yeso

Formation and below the San Andres Formation, thus, a late Leonardian age is assigned based on regional correlations (Lucas et al., 2013).

13

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

CHAPTER 2 SUBSURFACE STRATIGRAPHY AND FACIES DISTRIBUTION OF THE LATE PALEOZOIC RAINSVILLE TROUGH

The western portion of the greater Taos trough is exposed in the hanging wall of

Late Cretaceous-early Tertiary Laramide-aged thrusts and forms the modern Sangre de

Cristo Mountains; however, late Paleozoic sediments of the Rainsville trough are mostly buried in the subsurface (Fig. 2.1). Due to absence of adequate exposures and unavailability of seismic data, subsurface geology of the Rainsville trough is unknown other than a few speculative interpretations of basin geometry based on knowledge from limited exposures in the southwestern part and outside of the basin and a few key well penetrations (i.e., Soegaard, 1990; Baltz and Myers, 1999). This study utilizes petrophysical well log data from those well records through the Permo-Pennsylvanian strata to interpret subsurface geology of this area. A detailed attempt to describe subsurface stratigraphy of this area is not known from the literature. The purpose of this chapter is to build lithostratigraphic columns using primarily well log data and correlate lithofacies among wells to illustrate distribution of subsurface lithofacies in the Rainsville trough.

14

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 2.1 Map of the study area with location of wells and surface sections used in this study. Green circles indicate location of wells (details of these wells are given in Table- 2.1). Green boxes refer to surface sections used in this study that were described by Baltz and Myers (1999). Red box indicates approximate location of surface section of Holman Grade area described by Sweet and Watters (2015). Inset map illustrates ARM uplifts of the central region of the greater ARM deformation during the late Paleozoic. Green box of the inset map refers to the location of Taos-Rainsville trough area.

METHODS AND MATERIALS

Petrophysical well logs are the chief data utilized to interpret lithology necessary to predicting facies relationships and unravel depositional signatures in subsurface.

Whereas geophysical techniques can provide insights into geometry of the subsurface stratigraphy, no seismic data were available for this study. Although the most precise method for evaluating subsurface depositional successions is coring, only limited cored

15

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

intervals from two wells are available (see Table-2.1). Petrophysical well log data was

calibrated to available cores, drill-cuttings, and sample description from well reports.

Specific rock types were identified utilizing different well log parameters (Table-

2.1). Shale was identified using a gamma ray value >90 API and a separation of neutron

and density porosity by at least 12%. Sand was identified by using a relative low gamma

ray (usually less than 70 API), and cross-over of neutron and density porosity in

limestone matrix. For identifying limestone, a low gamma ray (usually less than 50 API),

neutron and density overlap, and a high Pe value (usually close to 5) were used.

Generally, gamma ray values ranging from 70 to 90 API were utilized for siltstone, along with a small separation between neutron and density (smaller separation than in shale).

Neutron and density curves of dolomite also show relatively smaller separation than shale, but a lower gamma ray value than both shale and siltstone (usually less than 70

API). Occasional slightly negative density also was useful for separating dolomite from other rock types. Coal was identified by using high values of both neutron and density porosity (>30%) and usually a low gamma ray value. For identifying anhydrite, a negative value of density porosity (usually less than -10%), neutron porosity close to zero and a very low gamma ray value (usually less than 20 API) were used. When available, a

Pe value of ~5 was used for limestone and anhydrite, ~1.8 for sandstone, ~3 for both shale and dolomite. In most cases, a fixed cut-off value was not valid for identifying similar rock types from different wells across the basin. Cut-off values were instead based on correlation to available core and cutting material.

16

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Table 2.1: Wells and corresponding petrophysical logs used in the study

Well Location Available datasets used for this study Well name no Latitude Longitude Types of well logs Other datasets Salmon Ranch A 1 35.99 -105.13 GR, Pe, PhiN_D cuttings, well reports

Salmon Ranch B 2 36.03 -105.14 GR, Pe, PhiN_D well reports

Mora Ranch 3 36.1618 -104.9387 GR, Sonic, PhiD cores, cuttings, well reports cuttings, strip log, well Mares Duran 5 36.2266 -105.1052 GR, Resistivity reports Arquello 6 36.21 -105.1 GR, PhiN, PhiD

Medina 7 36.2886 -105.1947 GR, PhiN, PhiD cuttings, well reports

Shell State 8 36.1804 -104.5752 GR, PhiN, PhiD cores, cuttings, well reports

McDaniel Sons 9 36.2742 -104.8315 GR, PhiN, PhiD cuttings

Gonzales Pittman 10 36.0310 -104.6662 GR, PhiN, PhiD

Leatherwood Reed 11 35.6177 -105.1368 SP, Radioactivity cuttings, strip log

Shoemaker Reed 12 35.6792 -105.0214 cuttings, strip log JD Hancock 13 35.6779 -105.1901 SP cuttings, well reports Sedberry St. Clair Sedberry 14 35.6648 -105.2123 strip log, well reports

Cities Services 16 35.89 -104.96 GR, PhiN, PhiD well reports Abbreviations: GR = gamma ray, Pe = Photoelectric, PhiN_D = neutron and density porosity, PhiN = neutron porosity, PhiD = density porosity, SP = spontaneous potentials.

Scanned petrophysical logs were initially available through the New Mexico

Bureau of Mines and Geology. Digitization of those using Neuralog software was undertaken to help interpret lithology, make correlations, and with better visualizations

(Fig. 2.2). A new technique of interpreting lithology from the digitized logs using

MATLAB programming was created (see Appendix 1).

IDENTIFYING LITHOLOGY AND BUILDING STRATIGRAPHIC SECTIONS

Each well that penetrated through Pennsylvanian and Permian strata of the

Rainsville trough was used for this study (see Figure-2.1 for locations of these wells).

17

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Regional stratigraphic knowledge from the southwest portion of the basin, originally described by Baltz and Myers (1999) was also utilized to help identify different lithofacies on the logs.

18

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 2.2 A portion of the scanned log from well-1, digitized version of the log and interpreted lithology using well log data. See legend in Figure-2.3 for description of identified lithotypes.

19

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 2.3 Interpreted lithology and constructed lithologic column from well log data using MATLAB programming (see Appendix 1 for MATLAB codes). See Figure-2.1 for location of wells.

Arroyo Peñasco Group

The Arroyo Peñasco Group that overlies the Precambrian basement rocks is mostly limestone dominated but contains sand and shaly sand at the lower portion. The facies was identified from only six of the studied wells located mainly at the central part

20

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019 of the basin and the Mora River section, and was absent in most of the eastern portion of the basin (Fig. 2.4). The Arroyo Peñasco Group is usually thin and thickness ranges about

20 to 50 meters (Fig. 2.4).

Figure 2.4 Isopach map of Arroyo Peñasco Group. Note that, due to its thickness less than 50 meter (except one well: Salmon A), a different color code was used for constructing the isopach map of the Arroyo Peñasco Group than the other analyses.

Sandia Formation

The Sandia Formation is composed of dominantly shale and alternating sandstone. This interval also contains several coal beds and occasional siltstone.

Carbonate is not common, but thin beds of limestone occur in a few wells. The Sandia

Formation is the thickest among all units and usually thickens toward wells located close

21

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

to the El Oro-Rincon uplift and thins away to the east, north and south of the basin (Fig.

2.5).

Figure 2.5 Isopach map of Sandia Formation. Black lines are isopach lines with contour interval of 200 meters. Thick black line along the eastern edge of El Oro-Rincon indicates inferred fault line with Pennsylvanian motion.

Arroyo Peñasco-Sandia formational boundary: The Mississippian Arroyo Peñasco

Group is dominantly composed of limestone, whereas, the Sandia Formation is shale dominated. Thus, the facies boundary between these two lithostratigraphic units was

22

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019 identified by abrupt increase in shale i.e. increase of gamma ray value in the overlying

Sandia Formation.

Porvenir Formation

Among all the Pennsylvanian and Permian strata, the Porvenir Formation is the most limestone rich with minor alternating sand and shale beds. The formation thins in wells located in the eastern part of the basin. Thickness of the unit slightly increases in the central-northern part of the basin (Fig. 2.6). Well-5 contains the thickest interval of the Porvenir Formation (Figs. 2.6 and 2.7).

23

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 2.6 Isopach map of Porvenir Formation. Black lines are isopach lines with contour interval of 100 meters. Thick black line along the eastern edge of El Oro-Rincon indicates fault line with inferred Pennsylvanian motion. See the color code in Figure 2.5.

Sandia-Porvenir formational boundary: Limestone content is not common in the

Sandia Formation but when present is typically thin. Thus, the lower boundary of the

Porvenir Formation was picked by first appearance of thick limestone beds or sudden increase in abundance of thin limestone beds. The Sandia Formation also contains several coal beds in most of the wells, which was also used for distinguishing between the two units.

Alamitos Formation

The Alamitos Formation consists of dominantly alternating shale and sand beds with thin limestone beds. The formation generally thins toward the eastern, northern and southern part of the basin but maintains fairly similar thickness across the basin (Fig 2.7).

Well-5 contains the thickest interval of the Alamitos Formation (Figs. 2.6 and 2.7).

Porvenir-Alamitos formational boundary: Identifying the boundary between

Porvenir and Alamitos formations was a challenge, as there was no obvious lithologic change or a marker bed present at their boundary. However, within the Alamitos

Formation, limestone content decreases and shale content increases relative to the

Porvenir facies. Thus, these criteria were utilized for choosing facies change between these two formations. Furthermore, the Porvenir Formation was also distinguished by occurrences of anhydrite in several wells.

24

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 2.7 Isopach map of Alamitos Formation. Black lines are isopach lines with contour interval of 100 meters. Thick black line along the eastern edge of El Oro-Rincon indicates fault line with inferred Pennsylvanian motion. See the color bar in Figure 2.5.

Sangre de Cristo Formation

The Sangre de Cristo Formation is dominated by abundant silt and silty sand with alternating shale. The facies onlaps Precambrian basement rocks of the Sierra Grande uplift to the east and Cimarron arch to the north of the basin. The formation does not demonstrate well defined trends in thickness variations, however, in generally the unit

25

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019 thickens towards central part of the basinal area and on top of Sierra Grande uplift located northeast of the study area. (Fig. 2.8).

Figure 2.8 Isopach map of Sangre de Cristo Formation. These strata extend on top of the Precambrian basement-cored uplifts i.e., Sierra Grande uplift to the east and Cimarron arch to the north. Blue dashed line indicates inferred western and southern extent of Sierra Grande uplift and Cimarron Arch, respectively. Black lines are isopach lines with contour interval of 100 meters. Thick black line along the eastern edge of El Oro-Rincon indicates fault line with inferred Pennsylvanian motion. See the color bar in Figure 2.5.

Alamitos-Sangre de Cristo formational boundary: The Sangre de Cristo

Formation is dominantly composed of siltstone with alternating sand and shale. Hence,

26

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

the boundary between the formations was picked based on last occurrence of limestone

and/or a decrease in shale content expected in the Alamitos Formation.

Yeso Formation

The Yeso Formation is composed of sandstone and silty sandstone facies. Sand

volume increases towards upper part of this formation. Two thick beds of anhydrite were

identified from the Yeso interval in several wells. The unit extends over the Precambrian

basement-cored uplifts to the east and north of the basinal area (Fig. 2.9). The thickness

of the formation increases to the east and south-eastern part of the study area.

Figure 2.9 Isopach map of Yeso Formation. These strata onlap Precambrian basement cored uplifts i.e., Sierra Grande uplift to the east and Cimarron arch to the north. Black 27

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

lines are isopach lines with contour interval of 100 meters. Thick black line along the eastern edge of El Oro-Rincon indicates fault line with inferred Pennsylvanian motion. See the color code in Figure 2.5. Blue dashed line indicates inferred western and southern extent of Sierra Grande uplift and Cimarron Arch, respectively.

Sangre de Cristo-Yeso formation boundary: Shale content abruptly decreases and

sand content increases in the Yeso Formation. Thus, the boundary between the units was

identified by a change from alternating silt, shale, and sand occurrence to mostly silty

sand to sand lithology. More importantly, the Yeso Formation contains thick anhydrite

beds (in most of the wells two thick beds of anhydrite were identified) that help

distinguish from underlying Sangre de Cristo Formation and overlying Glorieta

Formation.

Glorieta Formation

The Glorieta Formation is composed almost entirely of sandstone and shows

similar thickness across the area, except for a few wells in the south, where thickness

increases to slightly more than 100 meters (Fig. 2.10). The formation onlaps Precambrian

basement rocks of Sierra Grande uplift to the east and Cimarron arch to the north of the

basin (Fig. 2.10).

28

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 2.10 Isopach map of Glorieta Formation. Strata of the unit extend on top of the Precambrian basement cored uplifts i.e., Sierra Grande uplift to the east and Cimarron arch to the north. Black lines are isopach lines with contour interval of 100 meters. Thick black line along the eastern edge of El Oro-Rincon indicates fault line with inferred Pennsylvanian motion. See the color code in Figure 2.5. Blue dashed line indicates inferred western and southern extent of Sierra Grande uplift and Cimarron Arch, respectively.

Yeso-Glorieta formational boundary: Both the Yeso and Glorieta formations are composed of dominantly sandstone but sandstone of the Glorieta Formation contains less clay and more quartz as indicated by a constant value of low gamma ray. Whereas, the

29

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Yeso Formation contains silty and shaly sandstone that “cleans” towards the top part of the interval. The boundary between the Yeso and Glorieta formations was picked where shalyness of the sandstone composition decreases abruptly. Thus, a constant low gamma ray along with neutron-density crossovers was used for identifying the Glorieta

Formation.

Once formation boundaries were recognized from logs, different wells were correlated across the basin by connecting tops and bottoms of similar units. Figure 2.11 illustrates an example of lithologic correlations between two adjacent wells utilizing gamma ray curves.

30

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 2.11 An example of picking formation boundary and correlation of wells. Correlations of wells 11 (Leatherwood Reed) and 12 (Shoemaker Reed) based on gamma ray curves (see Figure-2.1 for locations and Table-2.1 for details). 31

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Different wells were correlated to illustrate thickness changes across the basin.

Several cross sections were produced along east-west (Fig. 2.12) and north-south (Fig.

2.13) transects.

Figure 2.12 Cross-sections G-H, C-D and K-F along east to west transects across the Rainsville trough. Numbers on top of each cross-section indicate wells and surface sections (see inset map for locations; also see Figure-2.1 and Table-2.1 for details).

32

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 2.13 Cross-section along north-south transects of the Rainsville trough. Numbers on top of each cross-section indicate wells or outcrop sections used in this study (see inset map on Figure-2.12 for locations; also see Figure-2.1 and Table-2.1 for details).

Additionally, isopach maps were created for the total Pennsylvanian and Permian strata (Fig. 2.14). Lithological cross-sections and isopach maps demonstrate the distribution of lithofacies determined from logs across the basin. Pennsylvanian strata usually thin towards the eastern (Fig. 2.12) and southern (Fig. 2.13) parts of the basin,

suggesting a roughly north-south trending basin depocenter located close to the El Oro-

Rincon uplift to the west. Pennsylvanian strata also exhibit a zero isopach line that

corresponds roughly with the western extent of the Sierra Grande uplift (Figs. 2.4-2.7).

33

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

During the Permian, location of sediment depocenters shifted to the south and east, such

that, depocenters situated atop previously exposed Precambrian basement of the

Cimarron arch and Sierra Grande uplift (Figs. 2.8-2.10; 2.14). Permian strata do not

demonstrate the variation in thickness that the Pennsylvanian strata does, rather the strata

are relatively more uniform throughout the basin (Figs. 2.8-2.10). In a few wells in the

north-western part of the basin, Permian strata are absent (wells 5, 6 and 7; see Figure-2.1 for locations). Mississippian strata are absent in several wells, mostly along the eastern and northern parts of the basin.

34

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 2.14 Isopach maps of (A.) Pennsylvanian and (B.) lower Permian strata in the Rainsville trough area.

35

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

CHAPTER 3 AGE CONSTRAINTS ON UPPER PALEOZOIC STRATA IN THE GREATER TAOS TROUGH REGION

Many ARM basins have poor age constraints for the basin fill. In the Taos trough,

various beds are fossiliferous allowing for some biostratigraphic age assessment, but

biostratigraphic constraints are poor in strata of the Rainsville trough owing to lack of

outcrop and core. This chapter reviews the biostratigraphic constraints within the greater

Taos trough region to help in age correlation of the largely subsurface stratigraphic

sections of the Rainsville trough. Published literature was reviewed thoroughly to

combine and compile systematically all available biostratigraphic data from the region.

Terrigenous sediments eroded from ARM uplifts were deposited largely on

extensive tropical shelves, and the ARM can be envisioned as uplifted islands surrounded

by shallow seas (e.g., Dickinson et al., 1983; Nelson and Lucas, 2011). The

Pennsylvanian stratigraphy of New Mexico preserves ARM-derived sand interbedded

with shale and carbonate beds that contain age-diagnostic marine fauna (e.g., Brotherton et al., in press). These fossil records are remarkably extensive, and fauna used for age

calibration include fusulinids and conodonts, as well as cephalopods, , and

other invertebrates (Nelson and Lucas, 2011).

As deformation affected modern New Mexico through the Pennsylvanian, ARM

basins generally developed as north- to northwest-trending elongate troughs that received

sediment shed from uplifts, and thicknesses of these depocenters varied through time

(Brotherton et al., in press). In the greater Taos trough region, Pennsylvanian strata

36

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

generally thicken toward faults with Pennsylvanian movement. However, lower Permian

strata thin towards and progressively onlaps Precambrian basement (Figs. 2.7 and 2.8).

These variations in stratal thicknesses and temporal migration of facies make correlation

of lithostratigraphic units within and across basins challenging. Utilizing biostratigraphic

constraints will help correlation of stratigraphic sections within the greater Taos trough

region given the high degree of lateral facies changes observed in most Pennsylvanian-

Permian basins in New Mexico.

METHODOLOGY RATIONALE

All available paleontological and palynological data were collected to assess ages

of deposition for the lithostratigraphic units filling the Taos and Rainsville troughs. The

greater Taos trough area was divided into three broad regions with potentially similar subsidence histories, namely, the Taos trough, Rainsville trough, and Pecos shelf (Fig.

3.1). Published literature was reviewed to combine and compile all available

biostratigraphic data in these three regions and summarized by formations in Figures 3.2

through 3.7. Summarized data include type of biostratigraphic records collected from a

formation, their geographic location, assigned age, corresponding authors and of the

data published. To assess the potential error inherent in assigning age to lithostratigraphy

within the greater Taos trough region, I utilized two lines of thinking. First, all the

available biostratigraphic data recovered from a formation were assessed and interpreted

to distill as small an age range as the data permits. Secondly, the existence of specific

climate-controlled facies that represent minimal time transgression were utilized as

primary marker beds for lithostratigraphic correlation. This chapter will focus on the

37

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

assessment of biostratigraphic data, whereas the climate sensitive facies will be addressed

in the next chapter.

Figure 3.1 The greater Taos trough is divided into three broad regions: the Taos trough (green shaded), the Rainsville trough (yellow shaded), and the Pecos shelf (blue shaded). Divisions are based on paleogeographic affinity. ARM uplifts are shaded in brown and thick black lines are inferred fault lines.

DATA ANALYSIS AND AGE CONTROLS

Espiritu Santo Formation

Extensive work has been done on outcrops of the Espiritu Santo Formation in the

greater Taos trough regarding its age determination and correlation among different parts

of the basin. All the paleontological data yielded Osagean and Meramecian age for the

Espiritu Santo Formation (Fig. 3.2; Armstrong, 1955; Armstrong and Holcomb, 1967; 38

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Clark and Read, 1972; Armstrong and Mamet, 1974, 1979, 1990; Armstrong et al., 1979,

2004). Most of the age assignment for the formation came from microfossil assemblages, more specifically from foraminiferal zonations. Based on collections

(microfauna of Zone 9) from chert nodules, a late Osagean age is assigned by Armstrong and Mamet (1974, 1990). Armstrong and Mamet (1979) suggested a late Tournasian age

depending on following foraminiferal assemblages: abundant Calcisphaera laevis

Williamson, Endothyra sensu stricto, Latiendothyra of the group L. parakosvenis,

Septabrunsiina parakrainica Skipp, Holcomb, and Gutschick, Septatournayella

pseudocamerata Lipina and Spinoendothyra spinosa (Chernysheva). Based on Endothyra

sp. And Plectogyra sp. Armstrong (1955) assigned a Meramecian age. No were

recovered from the subsurface of the Rainsville trough.

Figure 3.2 Biostratigraphic data from the Mississippian Espiritu Santo Formation from all three regions of the study area as illustrated in Figure 3.1. Vertical lines indicate 39

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

ranges of age distribution assigned by various authors. Green shaded area refers to an optimum age assignment to the formation based on all available biostratigraphic datasets. Carboniferous numerical ages are from Davydov et al. (2012).

Tererro Formation

Like the Espiritu Santo Formation, voluminous biostratigraphic data are available for the Tererro Formation. Both megafossils and microfossils were collected from this formation. Most biostratigraphic works yielded a Late Mississippian (Chesterian) age for the Tererro Formation (Fig. 3.3; i.e., Clark, 1966; Armstrong and Mamet, 1979, 1990;

Armstrong et al., 2004). Baltz and Read (1960) assigned an Early Mississippian age for

Macho and Manuelitas members of the formation, but a Late Mississippian age for the

Cowles Member. The basal Macho Member is conglomeratic and lies atop an erosional surface, and this relationship likely allowed for mixing with the underlying Espiritu Santo

Formation and the Early Mississippian age assignation for the unit. Based on foraminiferal assemblages from different localities in the greater Taos trough area,

Armstrong and Mamet (1979) assigned a Chesterian (Late Mississippian) age. Baltz and

Myers (1999) assigned a Meramecian age to the Manuelitas Member based on megafossils originally reported by Baltz and Read (1960) from Gallinas creek area in southwestern Rainsville trough. Sutherland (1963), based on megafossil collections from all three regions assigned a Mississippian age to this formation (Fig. 3.3).

40

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 3.3 Biostratigraphic data from the Mississippian Tererro Formation from all three localities. Vertical lines indicate ranges of age distribution assigned by authors. Solid lines indicate assigned age ranges based mainly on forams, whereas dashed lines refer to ranges assigned based on megafossils. Green shaded area refers to an optimum age assignment to the formation based on all available biostratigraphic datasets. Carboniferous numerical ages are from Davydov et al. (2012).

From all available biostratigraphic data, it is conclusive that, the Arroyo Peñasco

Group consists of the Espiritu Santo and Tererro formations is entirely Mississippian age in the greater Taos trough area (Figs. 3.2 and 3.3) and this age range does not overlap with overlying Sandia Formation (Figs. 3.3 and 3.4). Furthermore, biostratigraphic age ranges suggest that there was unconformity during the Kinderhookian and late Chesterian

(Figs. 3.3 and 3.4). Differentiating the Espiritu Santo from the Tererro Formation from log data was challenging, thus, a numerical age comprising the entire Mississippian (359 to 323.2 Ma) was used for the Arroyo Peñasco Group (Fig. 3.8; Davydov et al., 2012).

41

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Sandia Formation

The Sandia Formation shows a wide range of biostratigraphic age assigned by different authors based on various fossil assemblages. A Morrowan through middle to late Desmoinesian age is assigned for this formation by several studies (Fig. 3.4; i.e.,

Sutherland, 1963; Sutherland and Harlow, 1973; Baltz and Myers, 1984, 1999; Webster and Kues, 2006). Most of the Morrowan ages came from brachiopods and trilobites

(Sutherland, 1963; Sutherland, 1972; Sutherland and Harlow, 1973; Baltz and Myers,

1984; 1999). In some cases, a Morrowan age is assigned based solely on the absence of

Morrowan fossils from the underlying Tererro Formation (Clark, 1966; Sutherland,

1972). Fusulinid assemblages from various part of the region shows a late Atokan or

Atokan age for the top portion of the formation (Brown et al, 2013; Treat, 2014) (Fig.

3.4). More recent conodont recoveries from mostly the upper part of the formation indicate an Atokan to early Desmoinesian age (Fig. 3.4; Brown et al., 2013; Treat, 2014;

Watters, 2014). Kietzke (1990) proposed an entirely Desmoinesian age based on gastropod collections. Based on gastropods, Kues (1990) suggested a Desmoinesian through late Virgilian age to this formation. The smallest age assignment that encompasses most of the data, especially those with particular Carboniferous index qualities, is Morrowan through early Desmoinesian (Fig. 3.4).

In the greater Taos trough region, depositional age assigned to the Sandia

Formation spanning to the base of the Morrowan may not be warranted because comparison of the abundant data (Henry, 1973; Sutherland and Harlow,

1973) of northern New Mexico with conodont faunas (Idiognathodus klapperi Zone,

42

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Barrick et al., 2013a) reveals no evidence for any regional deposition earlier than the late

Morrowan. Moreover, lower Morrowan strata has not been demonstrated throughout central and northern New Mexico (Brotherton et al., in press). Thus, in the greater Taos trough area, I consider the base of the Sandia Formation is upper Morrowan, numerically this age assignment would approximately correspond to 321 Ma (Fig. 3.8; Davydov et al.,

2012).

Figure 3.4 Biostratigraphic data from the Pennsylvanian Sandia Formation from all three regions of the study area as illustrated in Figure 3.1. Vertical lines indicate ranges of age distribution assigned by authors. Solid lines indicate assigned age ranges based on fusulinids and conodonts, whereas dashed lines refer to ranges assigned based on various megafossils. Green shaded area refers to an optimum age assignment to the formation based on all available biostratigraphic datasets. Carboniferous and Permian numerical ages are from Davydov et al. (2012) and Ramezani and Bowring (2018), respectively.

43

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Porvenir Formation

Fossil assemblages from the Porvenir Formation yield the most consistence

depositional age among all Permo-Carboniferous stratigraphic units in the greater Taos trough region (Fig. 3.5). Most of the biostratigraphic works suggested a Desmoinesian age for the Porvenir Formation, primarily based on fusulinid and conodont recovery (Fig.

3.5; i.e., Read and Wood, 1947; Clark, 1966; Baltz and Myers, 1984, 1999; Treat, 2014;

Watters, 2014; Lucas et al., 2015). Treat (2014) suggested that uppermost Atokan carbonate may occur at the transition from the uppermost Sandia Formation to the lowermost Porvenir Formation based on conodont faunas, as well as fusulinid occurrences reported in Baltz and Myers (1999; Beedeina insolita near the base of the

Porvenir Formation). The shortest age assignment that encompasses all the

biostratigraphic data is latest Atokan through late Desmoinesian (Fig. 3.5).

44

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 3.5 Biostratigraphic data collections from the Pennsylvanian Porvenir Formation. Vertical lines indicate ranges of age distribution assigned by authors. Solid lines indicate assigned age ranges based on fusulinids and conodonts, whereas dashed lines refer to ranges assigned based on various megafossils. Green shaded area refers to shortest duration of age ranges based on available fusulinid and conodont occurrences. Carboniferous and Permian numerical ages are from Davydov et al. (2012) and Ramezani and Bowring (2018), respectively.

Biostratigraphic constraints for the Sandia Formation indicate that the top of the unit overlaps in age with the overlying Porvenir Formation. Studies that utilize gastropods, and brachiopods typically provide the most overlap (Fig. 3.4); however, these faunas are typically not considered good index fossils of the

Pennsylvanian. Thus, using only those studies that incorporate fusulinids and conodonts

(Fig. 3.4), the Sandia Formation could still range to the upper part of the lower

Desmoinesian. Sutherland (1963, 1972) and Sutherland and Harlow (1973) collected fossils from the La Pasada (southern part of the greater Taos trough) and Flechado

(northern part of the basin) formations (Fig. 3.4), which includes the lower limestone member of the Madera Formation within their La Pasada and Flechado formations. This limestone interval is now recognized as the Porvenir equivalent. Therefore, their early

Desmoinesian age assignation may have come from fossils collected from facies equivalent to the Porvenir Formation. Although, within the axis of the Taos trough,

Watters (2014) recovered conodonts of typical Porvenir age from facies equivalent to typical Sandia Formation lithology. Thus, the lithofacies that characterize the Sandia

Formation can locally overlap with the typically assigned age of the Porvenir Formation.

This is a problem when assigning ages to the subsurface lithofacies of the Rainsville trough.

45

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

The base of the Porvenir Formation was picked based on the presence of a prominent thick carbonate interval or presence of repetitive thin limestone beds. The age of this carbonate interval is unknown in the subsurface of the Rainsville trough.

However, regionally across the midcontinent and New Mexico, a widespread carbonate of “early” Desmoinesian age is expressed (Heckel, 1999, 2002, 2008, 2013; Boardman and Heckel, 1989; Boardman et al., 2004; Heckel et al., 2002; Barrick et al., 2004, 2013a,

2013b; Wahlman, 2013). Thus, I infer that the age of this prominent carbonate interval is

“early” Desmoinesian in age, like the widespread carbonate interval across the midcontinent and the carbonate interval in the axis of the Taos trough. Furthermore, early

Desmoinesian fusulinids were recovered by Baltz and Myers (1984, 1999) from the base of the Porvenir Formation near Mora Gap (location 4, Fig. 2.1) and Fragoso Ridge

(slightly south of the Mora Gap section), both the localities are situated near the depocenter axis of the Rainsville trough. Therefore, throughout the Rainsville trough, I considered the base of the Desmoinesian as the base of the Porvenir, which numerically corresponds to 311.8 Ma (Fig. 3.8; Davydov et al., 2012).

Alamitos Formation

Most of the paleontological data suggest a range of middle to late Demoinesian through Virgilian age for the Alamitos Formation in the greater Taos trough area (Fig.

3.6; i.e., Read and Wood, 1947; Sutherland, 1963; Sutherland, 1972; Sutherland and

Harlow, 1973; Webster and Kues, 2006). However, Sutherland and Harlow (1973), and

Sutherland (1972) suggested mid-Desmoinesian through late-Desmoinesian or

Missourian age in Taos trough area (Fig. 3.6). Baltz and Myers (1984, 1999) and Krainer

46

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019 et al. (2004) assigned a late Demoinesian through early Wolfcampian age for this formation (Fig. 3.6). Fusulinids provide the primary age constrains. Other significant fossil assemblages include crinoids, brachiopods, gastropods and pelecypods. Fossil collections from this formation are mostly on southwestern Rainsville trough and Pecos shelf. Available biostratigraphic data from the Taos trough area is very sparse (Fig. 3.1).

The shortest duration of time allowable by the biostratigraphic data (primarily based on fusulinid zonation) is the middle part of the Desmoinesian through earliest Wolfcampian

(Fig. 3.6). Some age assignments range into the middle to late Desmoinesian, which overlaps with the inferred end of the Porvenir depositional age.

Most of the biostratigraphic data from the top part of the Porvenir and bottom part of the Alamitos show an overlap in the late Desmoinesian. I consider this overlap perhaps due to a gradational transition of facies, which appears to have completely switched to facies indicative of the Alamitos Formation in the upper Desmoinesian. Thus, numerical ages assigned for this study for the duration of the Porvenir carbonate interval are 311.8

Ma to 307 Ma. (Fig. 3.8; Davydov et al., 2012).

The oldest beds of the Alamitos Formation are late Desmoinesian based on fusulinids (Beedeina sulphurensis, in Baltz and Myers, 1999), and conodonts of the

Swadelina neoshoenesis Zone of Barrick et al. (2013a). Fusulinid faunas of Missourian and Virgilian age occur sporadically in different sections of the Alamitos Formation. The youngest faunas reported from the Alamitos Formation includes Leptotriticites and

Triticites creekensis, taxa originally assigned an earliest Permian age, but now known to range from the latest Pennsylvanian into the earliest Permian (Wahlman, 2013).

47

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Therefore, I suggest a latest Desmoinesian through latest Virgilian depositional age for the Alamitos Formation. Numerical age assignments for the Alamitos Formation is 307

Ma to 298.9 Ma (Fig. 3.8; Davydov et al., 2012).

Figure 3.6 Biostratigraphic data assemblages from the Alamitos Formation from all three regions of the study area as illustrated in Figure 3.1. Vertical lines indicate ranges of age distribution assigned by authors. Solid lines indicate assigned age ranges based on fusulinids primarily, whereas dashed lines refer to ranges assigned based on various megafossils and plant fossils. Green shaded area refers to an optimum age assignment to the formation based on all available biostratigraphic datasets. Carboniferous and Permian numerical ages are from Davydov et al. (2012) and Ramezani and Bowring (2018), respectively.

Sangre de Cristo Formation

The majority of the age assignments for the Sangre de Cristo Formation in the

Taos trough area came from trace fossils (primarily ichnofauna) and plant fossils

(Fig. 3.7). Most of the authors suggested a Wolfcampian age (Fig. 3.7; Baltz and O’Neill,

1986; Hunt et al., 1990; Berman, 1993; Baltz and Myers, 1999; Krainer et al., 2004; 48

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Berman et al., 2013; Lucas et al., 2015; Reinhart et al., 2015; Voigt and Lucas, 2015).

Except for a single location in the Rainsville trough near Mora, Baltz and Myers (1999) also suggested Wolfcampian age for this formation (Fig. 3.7). Based on plant fossil Read and Wood (1947), and Clark (1966) assigned the base of the formation at the latest

Virgilian (Fig. 3.7). The smallest duration of time permissible by the data spans the

Wolfcampian (Fig. 3.7). The earliest Wolfcampian part of this interval overlaps with the age range of deposition of the underlying Alamitos Formation.

Based on plant fossils, Clark (1966) and Read and Wood (1947) suggested the base of the Sangre de Cristo to be at the latest Virgilian. However, none of the Virgilian assignments are based on Carboniferous index fossils. Moreover, except for one locality in Baltz and Myers (1999), more recent abundant biostratigraphic data from overall greater Taos trough area agrees on an entirely Wolfcampian age for the Sangre de Cristo

Formation (Fig. 3.7; i.e., Hunt et al., 1990; Krainer et al., 2004; Berman et al., 2013;

Lucas et al., 2015; Reinhart et al., 2015; Voigt and Lucas, 2015). Thus, I infer a

Wolfcampian age for the Sangre de Cristo Formation (Fig. 3.8). Zhang and Wang (2019) recently indicated that the base of the Leonardian corresponds to the mid-lower

Artisnkian, which shortens the duration of the Wolfcampian. This study utilizes that definition for the top of the Wolfcampian. Numerical ages assigned to this formation ranges from 288 Ma to 272.3 Ma (Fig. 3.8; Ramezani and Bowring, 2018).

49

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 3.7 Biostratigraphic data assemblages from the Sangre de Cristo Formation from all three regions of the study area as illustrated in Figure 3.1. Vertical lines indicate ranges of age distribution assigned by authors. Solid lines indicate assigned age ranges based on fusulinids primarily, whereas dashed lines refer to ranges assigned based on various plant fossils and vertebrate faunas. Green shaded area refers to an optimum age assignment to the formation based on all available biostratigraphic datasets. Carboniferous and Permian numerical ages are from Davydov et al. (2012) and Ramezani and Bowring (2018), respectively.

Yeso and Glorieta formations

No biostratigraphic data were available for the Yeso Formation from the Taos trough area. However, a Leonardian age is assigned for these formations based on biostratigraphic data south of the basin (Mack and Dinterman, 2002; Lucas et al, 2005,

2013, 2015).

The contact between the Wolfcampian Sangre de Cristo Formation and the overlying mostly non-marine to shallow-marine Yeso Formation is generally conformable in the region (Baltz and Myers, 1999), but biostratigraphic data for Yeso

50

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

and Glorieta formations within the Taos trough region is unavailable. Regionally, the

Glorieta Formation overlies the Yeso Formation and both the formations are considered

Leonardian (Mack and Dinterman, 2002; Lucas et al, 2005, 2013, 2015). Similar to

above, I utilize the base of the Leonardian as portrayed in Zhang and Wang (2019). Thus,

in this study, the Yeso and Glorieta formations are considered early and late Leonardian, respectively. Numerical age assignments are 288 Ma to 280 Ma for the Yeso Formation

and 280 Ma to 272.3 Ma for the Glorieta Formation (Fig. 3.8; Ramezani and Bowring,

2018). Note, data allowing for these units to span the Leonardian are unavailable.

51

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 3.8 Summary of biostratigraphic age distributions assembled formation-wise from Figures 3.2 through 3.7. Red bars indicate age distribution based on megafossils (i.e., 52

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

brachiopods, gastropods, trilobites, plant fossils and/or vertebrate fossils). Orange bars indicate age assignment based on fusulinids and conodonts. Green bars indicate inferred age assignment utilized in this study. Blue bar indicates age control data came from outside the greater Taos trough area. Carboniferous and Permian numerical ages are based on Davydov et al. (2012) and Ramezani and Bowring (2018), respectively. Abbreviations: P. = Peñasco, C. = Cristo.

SUMMARY AND STRATIGRAPHIC CORRELATIONS

Within the greater Taos trough region, more biostratigraphic data are available

from the southern part of the region than the northern part of the basin. East of the El

Oro-Rincon within the Rainsville trough, most of the Paleozoic sediments are subsurface

and available biostratigraphic data are limited mostly to the southwestern part of the

basin. However, justifiable lithological correlations among wells is made based on

available age controls from different parts of the basin.

Assembled biostratigraphic data shows age assignations of the various formations

vary widely based on types of fossil assemblages. From oldest to youngest, the following

lithostratigraphic assemblages yielded the following ages. An Early Mississippian

(Tournasian) age is assigned to the Espiritu Santo Formation by most of the authors. The

Tererro Formation is commonly assigned a Chesterian age. Thus, there was probably a

Meramecian aged unconformity between the Espiritu Santo and Tererro formations of the

Arroyo Peñasco Group. In the Taos trough region, these two formations can be differentiable from surface exposures (Fig. 3.9). However, it was difficult to distinguish them from well log signatures and they are undifferentiated in the Rainsville trough (Fig

3.10). Abundance and widely distributed fossil assemblages from carbonate-dominated

Arroyo Peñasco Group indicates an entirely Mississippian age, so the basal limestone

dominated units of the wells are considered Mississippian. The Sandia Formation 53

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

contains the most variable biostratigraphic age range in the region, but is typically late

Morrowan through Atokan. The Porvenir Formation demonstrates a consistent

Desmoinesian age. A late Desmoinesian through Virgilian age is assigned to the

Alamitos Formation. Most authors assigned a Wolfcampian age to the Sangre de Cristo

Formation. For the Yeso and Glorieta formations published literature do not show any fossil collection from the greater Taos trough region, but a Leonardian age is assigned primarily based on fossils collected from southern New Mexico.

Based on biostratigraphic synthesis discussed in this chapter and lithostratigraphy

of the area described by Miller et al. (1963) and Watters (2014), a generalized basin-wide

stratigraphy was built for the Pecos-Taos trough area (Fig. 3.9). A subsurface stratigraphic framework of the Rainsville trough was constructed based on biostratigraphic assemblages from the southwestern portion and outside of the basin, and lithofacies correlations build in chapter two (Fig. 3.10).

54

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 3.9 Basin-wide stratigraphy of the Taos trough region. Wavy and straight lines indicate unconformable and conformable relations, respectively, between overlying and underlying strata. Solid lines indicate similar relation throughout the basin whereas dashed lines indicate relation confirmed at some localities, but not certain if the same relation exists throughout the basin. Blank space indicates missing strata due to erosion or non-deposition. Carboniferous and Permian numerical ages are based on Davydov et al. (2012) and Ramezani and Bowring (2018), respectively. Abbreviations: Kin. = Kinderhookian, Osa. = Osagean, Mer. = Meramecian, Ches. = Chesterian. 55

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 3.10 Reconstructed basinal subsurface stratigraphy of the Rainsville trough. Wavy and straight lines indicate unconformable and conformable relations, respectively, between overlying and underlying strata. Solid lines indicate similar relation throughout 56

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019 the basin whereas dashed lines indicate relation confirmed at some localities, but not certain if the same relation exists throughout the basin. Blank space indicates missing strata due to erosion or non-deposition. Carboniferous and Permian numerical ages are based on Davydov et al. (2012) and Ramezani and Bowring (2018), respectively. Abbreviation: Miss. = Mississippian.

57

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

CHAPTER 4 PENNSYLVANIAN-PERMIAN CLIMATE-SENSITIVE FACIES FROM LOW LATITUDE RAINSVILLE TROUGH

The late Paleozoic ARM basins predominantly record thick sequences of terrigenous sediments derived from adjacent uplifts (McKee, 1975, Kluth and Coney,

1981, Ye et al., 1996), which in turn can yield information on the weathering and climatic history of source regions (Suttner and Dutta, 1986, Grantham and Velbel, 1988; Johnsson et al., 1988; Johnsson, 1990, 1992; Heins, 1993; Cecil et al., 2003). More importantly, some of these ARM basins recorded marine carbonate and other chemical and biogenic sedimentary rocks, which are considered as significant archives of paleoclimatic information (Cecil, 1990; Soreghan, 1994, 1997; Parrish, 1998; Cecil and Edgar, 2003;

Cecil et al., 2013 and many others). Published Paleozoic paleoclimatic studies are limited largely to the Permian interval in north-central New Mexico, (Mack et al., 1979; Mack,

2003; Tabor and Poulsen, 2008; Tanner and Lucas, 2017, 2018); whereas attempts to study Pennsylvanian paleoclimatic records in the area are unknown.

Facies that are related to regional climate change are usually basin-wide events.

Thus, climate-sensitive facies can be considered to have minimal time-transgressions and used as marker beds for correlating strata among wells. This chapter aims to identify climate sensitive facies from the Permo-Carboniferous strata of the subsurface Rainsville trough to help build a robust chronostratigraphic correlation by augmenting litho- and bio-stratigraphic correlations made in the previous chapters.

58

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

LATE PALEOZOIC CLIMATIC SETTING

The climate of a region is defined by the variation of precipitation, temperature,

wind and sunlight (Leeder, 1999; Cecil, 2003). Stratigraphic records of chemical

sedimentation as well as characteristics of clastic influx in a basin can archive

paleoclimate data of that region. Occurrences and distribution of certain facies, such as

coal, evaporite, paleosol, laterite, and eolianite intervals, in geological records serve as a

proxy for paleoclimatic records (Cecil, 1990; Parrish, 1998; Cecil and Edgar, 2003; Cecil

et al., 1985, 2003; Parrish and Peterson, 1988; Giles et al., 2013). These indicators have

been widely used for Paleozoic climatic reconstruction.

The Late Paleozoic earth was characterized by widespread glaciation and

deglaciation events in Gondwanaland known as the Late Paleozoic Ice Age (LPIA)

(Isbell et al., 2003, 2012; Fielding et al., 2008a, 2008b, 2008c; and references therein).

Although, the LPIA is commonly presented as one long protracted episode that waxed

and waned (i.e., Veevers and Powell, 1987; Frakes and Francis, 1988; Frakes et al.,

1992), recent findings indicate multiple short-time (1-8 m.y.) discrete episodes of glaciations that were separated by non-glacial intervals (Isbell et al., 2003, 2012; Fielding et al., 2008a, 2008b, 2008c; Franks et al., 2008; Birgenheier et al., 2009). However, the extent and timing of the LPIA and their impacts on low-latitude paleoclimate remains debated (Ziegler et al., 1987; Cecil, 1990; Heckel, 1994; West et al., 1997; Isbell et al.,

2003, 2008; Jones and Fielding, 2004; Angiolini et al., 2007; Poulsen et al., 2007;

Soreghan et al., 2007, 2008, 2009; Fielding et al., 2008a, 2008b; Tabor and Poulsen,

2008; Birgenheier et al., 2009; Bishop et al., 2010; Giles, 2012).

59

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Inferred far-field (paleoequatorial and paleotropical Pangea) effects of waxing

and waning of Gondwanan glaciation is recorded from cyclic sedimentation due to sea

level oscillation (Wanless and Shepard, 1936; Heckel, 1977, 1986, 1990, 1994, 2008;

Ramsbottom, 1979; Busch and Rollins, 1984; Ross and Ross,1985; Veevers and Powell,

1987; Frakes and Francis, 1988; Frakes et al., 1992; Soreghan,1994; Rankey, 1997;

Soreghan and Giles, 1999; Smith and Read, 2000; Wright and Vanstone, 2001;

Olszewski and Patzkowsky, 2003; Gibling and Rygel, 2008; Sweet and Soreghan, 2012).

Paleoclimatic records from Pangean low latitudes indicate that during the Permo-

Carboniferous the paleotropics were also characterized by significant changes, including

variations in atmospheric pCO2 (Montañez et al., 2007, 2016; Franks et al., 2014;

McElwain et al, 2016), widespread continental aridification (Parish, 1993; Kessler et al.,

2001), and major shifts in floral assemblages (Gastaldo et al., 1996; DiMichele et al.,

2001, 2009, 2010, 2016; Cleal and Thomas, 2005).

The long-term Permo-Pennsylvanian climate evolution of equatorial Pangea is

interpreted to reflect gradual drying from the Late Pennsylvanian through the Permian

(Parrish, 1993; Benison and Goldstein, 2001; Kessler et al., 2001; Tabor and Montanez,

2002, 2004; Ziegler et al., 2002; Cecil, 2003; Cecil et al., 2003; Schneider et al., 2006;

Tabor and Poulsen, 2008; Tabor et al., 2008; DiMichele et al., 2009; Retallack and

Huang, 2010; Boucot et al., 2013; Giles et al., 2013; Sweet et al., 2013; Pike and Sweet,

2018). Numerous coal beds and abundant siliciclastic flux has been identified from

Pennsylvanian strata in the Appalachian Basin (Cecil, 1985). Substantial coal beds are

also identified from the Pennsylvanian sections of the Midland basins (Eble, 2003).

60

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Eolian sandstone deposition and evaporite beds are common in Permian strata (Soreghan,

1992, 1997; Soreghan et al., 1997, 2002; Cecil et al., 2003; Sweet et al., 2013; Pike and

Sweet, 2018). Although conventionally the Permian and Late Pennsylvanian climate in western tropical Pangea has been interpreted as arid to semi-arid, several studies indicate significantly wetter local environments and/or more seasonal climate than previously thought (Sweet et al., 2013; Soreghan et al. 2002; Tramp et al., 2004, Soreghan et al.,

2014; Sweet, 2017; Pike and Sweet, 2018). The long-term aridification trend is rather dynamic and temporary reversals occur at many scales.

Paleoclimatic changes in paleotropics during the late Paleozoic are also attributed to several other factors other than high latitude glaciations, including tectonic drift (Gibbs et al., 2002; Tabor and Poulsen, 2008; Tabor et al., 2008), formation of Central Pangean

Mountains (CPM; Rowley et al., 1985; Kessler et al., 2001; Tabor and Montañez, 2002;

Otto-Bliesner, 2003; Parrish, 1993; Tabor and Montañez, 2004), and withdrawal of epeiric seaways (Tabor and Montañez, 2002; Ziegler et al., 2002). During the Early to

Late Pennsylvanian, a northward tectonic drift of western Pangea may have affected the regional and temporal pattern of palaeoclimate in tropical Pangaean basins (Tabor and

Poulsen, 2008; Tabor et al., 2008). Continued assembly of Pangea disrupted the zonal climate system and created a monsoonal climatic circulation like present day southeast

Asia (Parrish and Peterson, 1988; Parrish, 1993; Soreghan et al., 2002; Tabor and

Montanez, 2002).

61

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

CLIMATE INDICATORS: COAL AND EVAPORITE AS PALEOCLIMATIC PROXY

Paleoclimatic indicators from older rock records can be divided into three groups:

geochemical, paleontological and lithological (Parrish, 1998). The stable isotopes of

oxygen and carbon are the primary geochemical indicators of paleoclimate, although

isotopes of strontium, hydrogen, nitrogen and other isotopes can also reveal paleoclimatic

information (Railsback, 1990; Dettman and Lohman, 1993; Hayes, 1993; Grossman,

1992, 1994; Parrish, 1998 and references therein). Plants and respond to climate

and, thus, fossils can be used as excellent paleontological recorders of paleoclimate

(Hallam, 1973; Dodd and Stanton, 1981; McKerrow and Scotese, 1990; DiMichele and

Aronson, 1992; Parrish, 1998; DiMichele et al., 2001, 2009, 2010, 2016; Davydov,

2014). Paleosols and climate-sensitive sedimentary rocks are also used as lithological

indicators of paleoclimate (e.g., Parrish, 1998; Cecil, 2003; Tramp et al., 2004).

The stratigraphic distribution of coal and evaporite may reflect the temporal and

geographic variations of climate (Fig. 4.1; Cecil et al., 2003). Coal beds or peat formation

are indicative of readily available water and typically assigned to a wet climate (Fig. 4.1;

Briden and Irving, 1964; Robinson, 1973; Cecil et al., 1981, 1982; Ziegler et al., 1987;

Cecil, 1990; Scotese and Barret, 1990; Cecil and Dulong, 2003). The presence of

evaporite beds in geologic records indicates seasonally arid and probably warm climates

over a period of time (Fig. 4.1; Cecil, 1990; Parrish, 1998). Evaporite usually tends to be

deposited in low to mid latitudes (Parrish et al., 1982; Railsback, 1992). Although

evaporite can potentially be formed at high latitude (Wilson, 1979), they are

volumetrically insignificant (Parrish, 1998). Shallow-water carbonate deposition is 62

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

commonly considered warm water and low paleolatitudes (Parrish, 1998). Sandstone

compositions also can be used for paleoclimatic interpretation, but additional types of

paleoclimatic information are needed for this type of study (Suttner and Dutta, 1986,

Grantham and Velbel, 1988; Johnsson et al., 1988; Johnsson, 1990, 1992; Heins, 1993).

The volume of clastic sediment flux to the basin can also be correlated with the climate of

the source regions (Fig. 4.1; Cecil, 1990; Parrish, 1998; Cecil et al., 2003), although accommodation space for these sediments are created by subsidence and sea level changes.

63

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 4.1 Relation between sedimentological response and climate change. A. Probability of clastic sediment flux in response to climate wetness; B. Probability of chemical sediment formation in response to climate wetness. Adapted from Wilson (1973) and Cecil (1990).

64

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

METHODS AND MATERIALS

Coal, evaporite, carbonate intervals, and the rate of clastic sediment flux inferred

by large intervals of sand-rich strata are utilized to reconstruct late Paleozoic

paleoclimate of northern New Mexico and these climate-sensitive facies are used to

facilitate subsurface stratigraphic correlations in the Rainsville trough. In this study, I

utilized the presence of coal and evaporite layers to infer wet versus dry paleoclimatic

conditions, respectively. Secondly, the presence of carbonate was used as an indication of warm climate. Thirdly, thick packages of sand supplied to the basin was utilized as an indication of highly seasonal rainfalls.

Figure 4.2 Wells used in this study to decipher paleoclimatic information are denoted by green circles (see Table-2.1 for list of wells). Abbreviation: PPF = Picuris-Pecos Fault. 65

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

STRATIGRAPHIC RECORDS OF PENNSYLVANIAN AND PERMIAN PALEOCLIMATE

Pennsylvanian strata

Numerous coal beds were identified from well logs of the late Morrowan-Atokan

Sandia Formation (Fig. 4.3). Well-developed coals were not found in the available cored intervals of two wells, but numerous bands of carbonaceous materials were recognized in both cores and well cuttings. In addition, presence of coal beds was reported by Baltz and

Myers (1999) in outcrop of the Sandia Formation exposed farther to the west at the Mora

River section (location-4 of Figure-2.1). Evaporite intervals were not present in the

Sandia Formation and only a few thin carbonate beds were identified from the logs. The

Sandia Formation is dominantly composed of numerous thick shale intervals and alternating beds of sandstone (thickness about 1500 meters in the well ‘Salmon A’) suggesting siliciclastic flux to the basin was relatively high during the Early

Pennsylvanian.

66

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 4.3 Plots of coals versus evaporite in different wells from the Rainsville trough using MATLAB programming. Frequency of coal deposition was higher in the Sandia and Sangre de Cristo formations. Frequency of evaporite deposition was higher in the Yeso and Porvenir formations. See Figure-4.2 for well locations.

Repetitive and thick carbonate beds were identified in the Porvenir intervals of all studied wells. Several thick evaporite beds were recognized from well logs of the

Porvenir Formation (Fig. 4.3). Additionally, evaporite material was observed in cuttings and cores from wells 1 and 2 (Salmon – A and – B; see Figure-4.2 and Table-2.1).

Carbonaceous bands also were identified, but are uncommon and typically located near the base of top of the interval. Presence of these carbonaceous bands were not found in cuttings of the Porvenir intervals.

67

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Carbonate sequences were common in the Alamitos Formation. Few gypsious and anhydritic beds were reported from cuttings of these intervals. Two thick evaporite beds were identified from logs and cuttings of the Leatherwood Reeds well (Well-11; Table-

2.1 and Figure-4.2). Occurrences of evaporite in the Alamitos Formation from this well were also reported by Baltz and Myers (1999). Some carbonaceous bands were identified from a few wells utilizing well log data, but generally were rare in this formation.

Permian strata

The presence of coal in the Sangre de Cristo Formation was observed in several wells (Figs. 4.3 and 4.4). Baltz and Myers (1999) reported existence of gypsum in wells 1 and 2 (Salmon Wells ‘A’ and ‘B’), but no evidence of evaporite was found during this study. The gypsum beds reported by Baltz and Myers (1999) may be dolomitic beds as identified from well log signatures (Pe values close to 3, slightly negative density porosity in limestone matrix, GR values are intermediate – not as low as expected in evaporite). The thickness of this dominantly clastic formation is about 900 meters (in wells 1 and 2), and, thus, clastic sediment flux was likely high.

At least two thick sequences of evaporite deposits were identified from several well logs of the Yeso Formation (Figs. 4.3 and 4.4). Gypsum and anhydrite beds in this formation have been reported from nearby outcrop localities. No coal or carbonate beds were identified from well logs of the Yeso Formation, although occasional carbonate was reported from neighboring areas located further northwest (i.e., Chama basin; Tanner and

Lucas, 2018). Sandstone and siltstone of the upper and lower part of the Yeso Formation is interpreted to be eolian origin (Mack, 2003; Lucas et al., 2013, 2015).

68

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 4.4 Climate-sensitive lithologies from different wells. Well numbers indicated by numbers in circles (see Figure-4.2 for locations). Abbreviations: N.A. = North American, Leo = Leonardian, Wol = Wolfcampian, Vir = Virgilian, Mis = Missourian, Des = Desmoinesian, Ato = Atokan, Mor = Morrowan.

IMPLICATIONS FOR REGIONAL PALEOCLIMATE

The ARM basins were located near equatorial during the Pennsylvanian-Permian

(Scotese, 1997, 1999; Boucot et al., 2013). Correspondingly, a warm humid climate should have prevailed in this region. The Rainsville trough was located within 10°S of the equator during the Pennsylvanian and within 15°N during the Permian (Scotese,

1997, 1999), thus the region migrated across the paleoequator sometime in the Late

Pennsylvanian to early Permian.

69

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

The variability in occurrences of climate sensitive lithologies in the Lower

Pennsylvanian through lower Permian strata of the greater Taos trough region may

provide insights into regional and temporal climatic change in western Pangea.

Significant repetitive coal beds along with thick clastic intervals characterize the Sandia

Formation. Sediment flux is usually low in both ever dry and ever wet conditions;

maximum sediment flux favors a wet-dry seasonal condition (Fig. 4.1). Whereas, coal formation favors a tropical rainy or ever wet condition (Fig. 4.1). Thus, during the

Morrowan – Atokan ( – Early Moscovian) in northern New Mexico, a predominantly seasonal to wet climate prevailed that locally was wet enough to produce coal (Fig. 4.5).

Evaporite associated with abundant marine carbonate is indicative of arid to

semiarid climatic conditions during deposition of the Desmoinesian Porvenir and late

Desmoinesian-Virgilian Alamitos formations (Fig. 4.5). In Missourian-Virgilian strata,

anhydrite was found in only one well and abundance of carbonate decreases upward

through the interval, while coal intervals become more common (Fig. 4.4). Taken

together, the Desmoinesian records more aridity that the Missourian-Virgilian strata.

Notable coal beds and thick clastic deposits in the Wolfcampian Sangre de Cristo

Formation suggest a wet to seasonal climate during this time (Fig. 4.5). The Leonardian

Yeso Formation, containing gypsum beds in the Rainsville trough as well as throughout

the New Mexico, indicates deposition in an arid climate during the Leonardian

(). Eolian sandstone common to the Glorieta Formation indicates

70

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

semi-arid to arid climatic conditions during the late Leonardian, especially since no

coeval marine deposits are known in the area.

Climatic records show spatial and temporal variability during the Pennsylvanian-

Permian time in the low latitude North American continent. Traditionally, the Early

Pennsylvanian has been interpreted as a wet climatic period that gradually became drier

during Late Pennsylvanian through early Permian (Cecil and Edgar, 2003; Cecil et al.,

2013). However, evaporite in the Porvenir and Alamitos formations and repetitive coals in the Sangre de Cristo Formation within the Rainsville trough suggests a significantly drier period during the Middle to Late Pennsylvanian, which became relatively wetter and more seasonal during portions of the Wolfcampian in northern New Mexico.

Although, early Permian climate is referred mostly as semi-arid to arid in the western US continent (Parrish, 1993; Benison and Goldstein, 2001; Kessler et al., 2001;

Tabor and Montanez, 2002, 2004; Ziegler et al., 2002; Cecil, 2003; Schneider et al.,

2006; Tabor and Poulsen, 2008; Tabor et al., 2008; Retallack and Huang, 2010; Boucot et al., 2013), indication of seasonality and wet conditions is suggested from several recent studies. Allen et al. (2011) suggested highly seasonal climate in tropical settings and these seasonal intervals coincide with glacial episodes; thus, they infer seasonality is due to far-field impacts of glacial changes in southern hemisphere. Similarly, the seasonality

indicated by high sediment flux of the Early Pennsylvanian Sandia Formation and the

Wolfcampian Sangre de Cristo Formation inferred from the Rainsville trough strata may

be correlated to glacial episodes of southern hemisphere (Fig. 4.5). Fossil forests in the

71

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019 early Permian Sangre de Cristo Formation also provide evidence of wet conditions

(Reinhart et al., 2015).

Figure 4.5 Glacial episodes (indicated by blue bars) in Gondwana during the late Paleozoic. Summary of Permo-Carboniferous paleoclimatic interpretation using lithological indicator from the Rainsville trough.

72

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

UTILIZATION OF CLIMATE-SENSITIVE FACIES FOR CORRELATIONS

Even though limitations exist, reasonable correlations of coeval subsurface

stratigraphy using climatic markers across the Rainsville trough is possible. Abundant

and repetitive coals of the Sandia Formation, evaporite and thick carbonate intervals of

the Porvenir Formation, and evaporite in the Yeso Formation were likely regional

climate-induced events and useful for basin-wide correlations. Tying these climate- sensitive lithologic indicators to a formation and coupling with regional age assignation provides a subsurface chronostratigraphy for the Rainsville trough (Fig. 4.6).

73

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 4.6 Utilization of climate-sensitive facies in constructing subsurface stratigraphy of the Rainsville trough. Wavy and straight red lines indicate unconformable and

74

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019 conformable relations, respectively, between overlying and underlying strata. Solid red lines indicate similar relation throughout the basin whereas dashed red lines indicate relation confirmed at some localities, but not certain if the same relation exist throughout the basin. Blank space indicates missing strata due to erosion or non-deposition. Carboniferous and Permian numerical ages are based on Davydov et al. (2012) and Ramezani and Bowring (2018), respectively.

75

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

CHAPTER 5 3-D BASIN SUBSIDENCE MODEL

To develop a 3-D basin subsidence model, produced and available composite sections of the Taos-Rainsville tough, available biostratigraphic data and age correlations from climate sensitive facies were utilized to construct tectonic subsidence curves from different parts of the basin. The purpose of this chapter is to construct numerous 1-D backstripped tectonic subsidence curves that, when time correlated, would allow reconstruction of the 3-D basin subsidence architecture necessary to assess the correct basin model. Determining between a flexural mechanism versus a mixed flexural and strike-slip mechanism has implications for strain patterns across the orogen and for correct paleogeographic depiction of the area.

METHODS AND MATERIALS

Building individual stratigraphic section for construction of subsidence curves

Composite sections constructed from wells and surface sections (Fig. 2.1) were

subjected to backstripping (Steckler and Watts, 1978; Angevine et al., 1990) to assess

isostatic and tectonic components of subsidence. Correlation of these 1-D backstripped

curves by time surfaces allowed for the development of a 3-D basin model.

Sweet and Watters (2015) studied a series of road-cut sections along Highway

518 and constructed a detailed composite stratigraphic column for the Holman grade area

located approximately at the Taos trough depocenter (Fig. 2.1, red box area). Watters

(2015) used this composite section that was entirely lower Desmoinesian (data came from

both the base and top of the section) as part of a regional composite stratigraphic column. 76

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

The rest of the Pennsylvanian section was compiled from thickness and lithology of brachiopod and fusulinid zones of the region (Miller et al., 1963). Thickness and lithology of Permian-aged strata were compiled from Baltz and Myers’ (1999) Mora

River section. The composite section constructed by Watters (2015) for the axis of Taos trough was used for comparison between Taos and Rainsville troughs (red lines, Fig. 5.2; see Figure 2.1 for location).

Obtaining parameters for backstripping: decompaction factor and porosity

To assess total subsidence within a basin utilizing this method, the sedimentary column first needs to be iteratively decompacted to ascertain the true component of isostatic sedimentary loading. Decompaction involves removing the loss of porosity through burial to arrive at a cumulative decompacted stratigraphic column, which can be calculated using the following equation:

Here Zd is the thickness of a decompacted sedimentary interval, Zi is the thickness of compacted strata, Φi is the observed compacted porosity (Bond and Kominz, 1984), and Φd is the predicted porosity after decompaction at a certain depth. To estimate porosity at certain depth for specific stratigraphic layer within the basin, the following equation was used:

77

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Here, ф is the lithology specific surface porosity (Bond and Kominz, 1984), z is depth of interest (usually ½ thickness of decompacted interval is used) for porosity prediction, and ck is the lithologic constant. Within a stratigraphic interval, values for ф and ck were determined based on the relative percentages of lithologic composition i.e., shale, siltstone, sandstone, micrite, and fossiliferous limestone etc. (e.g., Hegarty et al.,

1988).

Obtaining parameters for backstripping: assessing age

Determination of the depositional ages of each stratigraphic section along with its thickness is very crucial during construction of basin subsidence curves, yet in this study assessing age of sedimentary strata was a challenge. Available paleontological data was the primary age calibration, but in most subsurface stratigraphic sections biostratigraphic data is not available. In those cases, age of stratigraphic intervals was estimated by regional lithostratigraphic correlation. Details on age calibration and correlations discussed in Chapters 3 and 4 and a summary of age ranges of lithostratigraphic units used in this study is given in Table-5.1.

Table 5.1: Depositional age ranges of lithologic units approximate numerical lithologic units depositional age age assignment (Ma) Glorieta late Leonardian 280-272.3 Yeso early Leonardian 288-280 Sangre de Cristo Wolfcampian 298.9-288 latest Desmoinesian- Alamitos Virgilian 307-298.9 Porvenir Desmoinesian 311.8-307 Sandia late Morrowan-Atokan 321-311.8 Mississippian Arroyo Peñasco 359-323.2

78

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Obtaining parameters for backstripping: water depth and sea level changes

Subsidence is the chief process driving accommodation space in basins and is

defined by an isostatic and tectonic component (Allen and Allen, 1990). Accommodation

space is also controlled by sedimentation rate and rise and fall of global sea level (e.g.,

Catuneanu, 2006). Facies analysis of constructed stratigraphic sections, both from

outcrops and subsurface were utilized to interpret local water depths. To estimate eustatic

change, the average magnitude of estimated eustatic change from other basins was

incorporated (Haq and Schutter, 2008; Eros et al., 2012).

The Process of Backstripping: Iteratively Removing the Stratigraphic Record

Backstripping involves decompacting the youngest interval (lithologic layer)

within a stratigraphic column, and then removing that thickness and decompacting the

underlying intervals (Steckler and Watts, 1978; Angevine et al., 1990). The process is

repeated iteratively for each successive older interval. The final product is a decompacted

sedimentary column and represents the total amount of recorded subsidence (Stot).

Tectonic subsidence (Stec) is the amount of subsidence left over from the Stot after

removing the isostatic component due to sedimentary loading (Siso) and water loading.

The following equation is used to calculate tectonic subsidence (Stec):

Here, Wd is the estimated water depth for the interval (using water depth ranges for depositional facies), ΔSL is the change in sea level form the previous interval, ρM is

79

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

the density of the mantle and ρW is the density of the water. The right half of the

equation is the isostatic effect of water loading within the basin.

The isostatic subsidence (Siso) due to a sediment load is calculated using the

following equation:

Here ρS is the density of the sedimentary column, which is defined by the ratio of

the density of rock volume to the density of pore volume, Zd is the decompacted

thickness of a sedimentary interval, ρM is the density of the mantle and ρW is the density

of water.

RESULTS AND DISCUSSIONS

Basin aspect ratio

Length versus width of different basins from known tectonic settings were plotted

to compare with the Taos-Rainsville trough (Fig. 5.1). Basin aspect ratios show a linear

relationship between length and width for both flexural and strike-slip basins. However, the ratio is usually higher for flexural basins indicating relatively larger basin dimensions

(lengths and widths) than strike-slip (wrench) basins, thus they plot in distinct regions

(Fig. 5.1). The blue dashed line in Figure 5.1 separate these two regions. An average length and width of the greater Taos trough, and separately the Taos and Rainsville troughs were plotted with basins from known tectonic settings. The Taos trough plots in

80

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

the strike-slip regime, the greater Taos plots in flexural regime, and the Rainsville trough falls in the transition zone between strike-slip and flexural basins (Fig. 5.1).

Figure 5.1 Length and width of the Rainsville trough, Taos trough, and the greater Taos trough are plotted to compare with basin aspect ratios of flexural and strike-slip basins from known settings. Both length and width are plotted in log scales. Reference data compiled from Barbeau (2003) and references cited therein.

Tectonic subsidence analysis of the Taos-Rainsville trough

Composite sections produced from all thirteen wells (see Figure-2.1 and Table-

2.1) were subjected to backstripping for tectonic subsidence analysis. Two composite sections (locations 4 and 15 of Fig. 2.1; Mora Gap section and Section-G of Baltz and

Myers (1999), respectively) from surface exposures of the Rainsville trough were produced by compiling data presented by Baltz and Myers (1999). Decompacted total 81

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

and tectonic subsidence of all fifteen sections from the Rainsville trough (13 wells and 2

surface sections; see Figure-2.1 for locations) for the Pennsylvanian through the early

Permian are plotted in Figure-5.2. Both total and tectonic subsidence curves show similar

shapes. Prior to the Early Pennsylvanian, subsidence was minimal and Mississippian subsidence rates are not shown here. Both the rate and magnitude of subsidence greatly increased during the Early Pennsylvanian (Bashkirian/Morrowan-Atokan; Fig. 5.2).

However, during the Middle and Late Pennsylvanian (Desmoinesian through Virgilian)

the slope of most of the subsidence curves becomes shallower and may indicate a relative tectonic quiescence period in the Rainsville trough. However, constant or increasing

slopes of subsidence curves in a few wells (wells 3, 5 and 7; Fig. 5.2) indicate that

subsidence continued during this time. The slopes of the subsidence curves steepen again

during the earliest Permian (Wolfcampian) and become shallower again during the

Leonardian. The magnitudes of subsidence from locations proximal to the El Oro-Rincon are greater (light green lines; Fig. 5.2) than those of subsidence farther away from the El

Oro-Rincon uplift (dark green lines; Fig. 5.2).

Both magnitude and timing of peak subsidence pattern in the Taos trough greatly differs from subsidence in the Rainsville trough. In the Taos trough, during the Early

Pennsylvanian magnitude of subsidence is relatively lower than curves from the

Rainsville depocenter (see Figure 2.1 for locations). The magnitude increases abruptly during the Desmoinesian but decreases during the Late Pennsylvanian. The subsidence curve shows an uptick again at the beginning of the Wolfcampian.

82

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

83

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 5.2 Backstripped total (A) and tectonic (B) subsidence curves from the Taos- Rainsville trough. Red lines indicate subsidence of the Holman Grade area of the Taos trough and green lines indicate subsidence curves from the Rainsville trough (see Figure 2.1 for locations). Light green curves refer to subsidence of basin depocenter area and locations proximal to the El Oro-Rincon uplift. Dark green lines represent subsidence of relatively distal locations from the El Oro-Rincon. A MATLAB program was developed to produce these subsidence curves (Appendix 2).

Duration, magnitude, and rate of subsidence obtained from the resultant tectonic

subsidence curves (Fig. 5.2B) were compared with subsidence curves from known

tectonic settings to assess appropriate (flexure vs wrenching) basin formation mechanism.

The Taos trough is most similar to basins from known strike-slip regimes (red lines, Fig.

5.2). However, the overall magnitude and duration of subsidence does not comply perfectly with either strike-slip or flexural basins. In the Taos trough, the Morrowan-

Atokan portion of the subsidence curve (Fig. 5.2) appears most consistent with known basins undergoing flexural subsidence; however, the steepened slope of subsidence during the Desmoinesian, probably in response to dextral movement along the Picuris-

Pecos fault (Cather et al., 2011). In contrast, tectonic subsidence curves from the

Rainsville trough more closely resemble flexural basin subsidence patterns. Among the fifteen studied locations of the Rainsville trough, sections located near the sediment depocenter and proximal to the El Oro-Rincon (Figs. 2.1 and 5.2) show similar patterns of tectonic subsidence with each other. In contrast, locations distal to the El Oro-Rincon show a different trajectory of subsidence with lesser magnitude (Fig. 5.2). This relationship indicates that the load of the El Oro-Rincon may have not migrated to the farther east in the Rainsville trough (Fig. 5.3). The lack of migration in subsidence patterns probably represents flexure from a static load, as suggested by Sturmer et al.

84

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

(2018) for several other ARM basins (Fig. 5.3). The early Permian portion of tectonic

subsidence curves demonstrate a Wolfcampian uptick in subsidence magnitude relative to

the Late Pennsylvanian, but indicate negligible subsidence in the Leonardian. Moreover,

the Wolfcampian sediment depocenter did not occur in similar depocenters as the

Pennsylvanian, but rather demonstrated relatively minor variance in stratigraphic

thickness across the region (Figs. 2.5-2.8). The early Permian portions of the tectonic subsidence curves from both basins are shown as dashed lines (Fig. 5.2B), which would

not be attributable to ‘classic’ ARM deformation as Permian strata onlap ARM uplifts

(i.e., Sierra Grande uplift and Cimarron arch) in this region (see Chapter 2). These

Permian subsidence patterns may be attributable to the two-phased subsidence model of

Soreghan et al. (2012) or a different tectonic model needs to be considered for broad

regional subsidence (Sweet et al., 2019).

85

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 5.3 Tectonic subsidence curves from different ARM basins were plotted with reference data from known tectonic settings to compare between flexure due to static loads versus flexure due to migrating loads. (time in Millions of ). Green and red lines represent tectonic subsidence curve of well Salmon-A in the Rainsville and Holman Grade area of the Taos trough, respectively. Gray lines indicate reference data from

86

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

basins with known tectonic settings (data compiled from Sinclair and Naylor, 2012). Basins with subsidence rate less than 0.05 km/m.y. (kilometers per million years) are considered flexure from slow and protracted (migrating) loads and more than 0.05 km/m.y. are flexure created by static loads (Sinclair and Naylor, 2012; Sturmer et al., 2018).

Rate of basin subsidence through time

The rate of tectonic subsidence from each basin also demonstrates differences in subsidence style. Considering probable different timing of ARM deformation, tectonic subsidence rate was calculated for different interval of basin subsidence (initiation and termination of basin subsidence). For each interval, the rate of tectonic subsidence was plotted for the duration of subsidence (Fig. 5.4). The rate of tectonic subsidence plotted against different locations indicate maximum rate of tectonic subsidence in the Rainsville trough during the late Morrowan-Atokan. The rate decreased during the Desmoinesian and continued to lessen to the end of the Pennsylvanian. However, in the Taos trough, the maximum rate of subsidence was during the Desmoinesian. These relationships demonstrate the differences in timing of peak subsidence between the two closely related adjacent basins. During the Wolfcampian, the rate of subsidence increased significantly compared to the Late Pennsylvanian subsidence. During the Leonardian, subsidence rate reduced again and may indicate the final demise of the late Paleozoic tectonic influence in this region.

87

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

88

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 5.4 Chart showing the relationship of tectonic subsidence rate of the Taos- Rainsville trough during Pennsylvanian and early Permian. Numbers on X-axes refers to wells and surface sections used for calculating tectonic subsidence. Light green bars indicate subsidence rates calculated from locations proximal to the El Oro-Rincon whereas darker green bars refer to distal locations (see Figure 2.1 for locations). The letter ‘H’ refers to Holman Grade area of the Taos trough.

3-D basin model for the Rainsville trough

All 1-D subsidence curves from fifteen sections of the Rainsville trough were

time-correlated and structure maps created for each time correlated surface. Among all

lithostratigraphic units studied here, the Leonardian Glorieta Formation was the easiest to

trace in subsurface. Therefore, taking the top of the Glorieta Formation as the datum, the

depths to all time-correlated surfaces were calculated to create structure maps.

Structure maps show the evolution of the Rainsville trough from the

Mississippian through the early Permian (Figs. 5.5 and 5.6). The structure of the top of

the Precambrian basement surface and the structural surface of the Mississippian Arroyo

Peñasco Group almost mimic each other, indicating minimal or no tectonic subsidence.

Thus, the basin did not start to develop until the Early Pennsylvanian. A sudden increase

of the depth to the surface of the Arroyo Peñasco from the late Morrowan-Atokan Sandia

structure indicates abrupt acceleration of basin subsidence during the Early and early

Middle Pennsylvanian. Modest change in structural elevation between the Sandia and

Porvenir surfaces indicates that the rate of subsidence reduced significantly during the

Desmoinesian. There is negligible to no change between Porvenir and Alamitos structural

surfaces, suggesting a tectonic quiescent period during the Missourian-Virgilian. During the Pennsylvanian, the sediment depocenter remained constant at a north-northeast to

southwest direction (Figs 5.5 and 5.6). However, abrupt changes in elevation between the 89

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Alamitos and Sangre de Cristo structural surfaces suggest a renewed subsidence during

the early Permian (Wolfcampian). The roughly north-south trending Pennsylvanian

sediment depocenter also shifted to a more east-west trending located more southerly in the basinal area during the early Permian (Fig. 5.6), indicating a change in tectonic regime of the basin development.

90

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 5.5 Structure development of the Rainsville trough through time. Surface maps of Precambrian basement rocks (top-left), Mississippian Arroyo Peñasco Group (top-right), Morrowan-Atokan Sandia (bottom-left), Desmoinesian Porvenir (bottom-right). Top of the Glorieta surface is taken as the datum and structure maps are created by calculating

91

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019 depth from the datum to corresponding surfaces. See color bar in Figure 5.6 for depth of structural surfaces. Contours are in meters.

Figure 5.6 Structure maps of Missourian-Virgilian Alamitos Formation (top-left), Wolfcampian Sangre de Cristo Formation (top-right) and Leonardian Yeso Formation (bottom-left). Top of the Glorieta Formation is considered as the datum and structure 92

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019 maps represent depth from this datum to top of corresponding formation. Contours are in meters.

Figure 5.7 3-D structure maps of different lithostratigraphic units through time, generated using MATLAB programming (see Appendix 3). Top of the Glorieta

93

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Formation is considered as the datum and structure maps represent depth from this datum to top of each corresponding formation.

Taos-Rainsville basin model

Tectonic subsidence curves indicate onset of basin subsidence during the late

Morrowan (Figs. 5.5, 5.7 and 5.8) in the greater Taos trough area. The subsidence

continued during the Atokan in both the Taos and Rainsville troughs. Coarse arkosic

sediments in the upper part of the Sandia Formation probably indicate denudation of

adjacent ARM uplifts. Very thick clastic strata and high accumulation rate in the

Rainsville trough also indicate accommodation space created rapidly by tectonic

subsidence in this basin (Fig. 5.8). In the Taos trough, late Morrowan-Atokan sediments are much thinner than in the Rainsville trough (Fig. 5.8). Thus, despite both the Taos and

Rainsville troughs started subsiding synchronously during the Early Pennsylvanian, the

Taos trough was tectonically less influenced than the Rainsville trough during this time.

Subsidence patterns across the Rainsville trough demonstrate the expected pattern

for a basin that was loaded from the west where a foredeep was located in front of the

load, but accommodation space decreases more distally (Fig. 5.3) (e.g., DeCelles and

Giles, 1996). However, the lack of migration in the depocenter (Figs. 2.5-2.8) indicates

that the load was static similar to ARM basins elsewhere (Sturmer et al., 2018).

Tectonic subsidence pattern from the Taos trough demonstrates similar west

loading during the late Morrowan-Atokan, albeit of lesser magnitude than the Rainsville

trough (Fig. 5.3). However, in the Desmoinesian, the rate and magnitude of tectonic

subsidence greatly increased for the Taos trough more closely resembling subsidence

94

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

patterns from transtensional settings. Thus, it seems likely that both the El Oro-Rincon

and Picuris-Pecos faults manifested as east-verging thrusts during the late Morrowan-

Atokan, but slip on the Picuris-Pecos fault expressed a large component of strike-slip movement during the Desmoinesian. Independent assessments using provenance, paleomagnetic, and field data indicate a dextral component of Desmoinesian slip on the

Picuris-Pecos fault (Woodward et al., 1999; Cather et al., 2006, 2011; Wawrzyniec et al.,

2007). However, interaction of the Picuris-Pecos and El Oro-Rincon faults at depth is unknown. The decrease in tectonic subsidence in the Middle to Late Pennsylvanian within the Rainsville trough likely indicates that once dextral slip occurred on the Picuris-

Pecos fault, displacement along the El Oro-Rincon fault diminished. Baltz and Myers

(1999) speculated that the Picuris-Pecos fault may have curved to the west at the northern edge of the Brazos uplift, which, if true, would have created a releasing bend in the vicinity of the Taos trough, potentially accounting for the substantial component of dip- slip required to accumulate the Desmoinesian-aged sediment. Confirmation of the trajectory of the Picuris-Pecos fault is difficult since it would be buried beneath the

Cenozoic . Alternatively, rotation of the El Oro-Rincon block could have produced dip-slip movement along the Picuris-Pecos fault.

During the Missourian-Virgilian, slopes of subsidence curves from all studied

wells and outcrop locations become shallower. The thickness of strata and rate of

sediment accumulation decreased in both Taos and Rainsville troughs. During this time,

thrusting of the El Oro-Rincon uplift and transcurrent movement of the Picuris-Pecos

Fault to the west may have stopped (Fig. 5.8). All these features suggest a tectonic

95

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

quiescence period in this region and probably indicate cessation of the regional

Pennsylvanian tectonic driver.

All tectonic subsidence curves show an increase in subsidence rate during

Wolfcampian, suggesting renewed tectonic subsidence. The rate of subsidence was fairly

constant throughout the region unlike the Pennsylvanian (Fig. 5.4). Lower Permian strata onlap the Sierra Grande uplift to the east and the Cimarron arch located north of the

Rainsville trough (Fig. 5.8). Lower Permian strata may also have buried the El Oro-

Rincon uplift in the center of the basin; however, Laramide-aged reactivation of the intrabasinal thrusts on the eastern side of the El Oro-Rincon preclude formal assessment of this relationship. A roughly north-south trending sediment depocenter located proximal to the El Oro-Rincon during the Pennsylvanian that shifted more easterly and southeasterly locations during the Permian. Onlapping of previously exposed crystalline basement by early Permian strata and an uptick in tectonic subsidence rates is consistent with epeirogenic subsidence hypothesized to prevail during the early Permian (Soreghan et al., 2012; Sweet et al., 2019).

A constant but very low rate of sediment accumulation, shown by flatten subsidence curves (Fig. 5.2) and a reduced rate of subsidence (Fig. 5.4) during the

Leonardian suggests the final demise of Permian deformation in this region.

96

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 5.8 West to east cross-sectional view illustrating the evolution of the Taos- Rainsville trough through time. The inset map shows the location of the cross-section.

ARM implications

The differences in tectonic subsidence styles and timing within the greater Taos trough region indicate that Taos and Rainsville trough experienced two different styles of

Pennsylvanian-aged subsidence. Furthermore, Pennsylvanian-aged versus Permian-aged

97

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

subsidence also varies in style within the greater Taos-Rainsville trough. Results obtained

from this study were compared to available data from other ARM basins to assess the

broader tectonic setting. A wrench-fault model of the Taos trough lends support to the

Dickinson and Lawton (2003) model, whereas a thrust-loaded flexure model aptly supports the Ye et al. (1996) and Leary et al. (2017) structural models of the ARM. If the

Picuris-Pecos fault did curve west at the northern edge of the Brazos uplift, then in the subsurface beneath the San Juan basin the fault should show north-east verging thrusting,

which would provide kinematic consistency with both the Ye et al. (1996) and Leary et

al. (2017) models.

Both the Taos and Rainsville trough subsidence style are consistent with flexure due to static load and comply with subsidence style of the Paradox and Eagle basins

(Johnson et al., 1988, 1992; Barbeau, 2003; Sturmer et al., 2018). Sturmer et al. (2018)

contrasted late Paleozoic basins of western Laurentia (Ely-Bird, Wood River, Wyoming shelf) to ARM basins and found that ARM basins behaved as static loads. Results from this study are consistent with those results (Fig. 5.3) and indicate that tectonic activity off the Nevadan margin probably had minimal influence on the basin.

98

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Figure 5.9 The Pennsylvanian and early Permian tectonic subsidence of Taos-Rainsville trough are compared with subsidence in other ARM basins (data compiled from Geslin, 1998; Soreghan et al., 2012; Seals and Soreghan, 2013; Sturmer et al., 2018).

99

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

CHAPTER 6 CONCLUSION

A subsurface stratigraphic framework of the late Paleozoic Rainsville trough was

built utilizing petrophysical well log data (chapter 2). Lithostratigraphic columns were

constructed by interpreting scanned well log data. Scanned logs were manually digitized, and MATLAB programming was utilized to introduce a new technique to interpret lithology from petrophysical logs. The produced subsurface sections from well log data were correlated to show the facies characteristics of the basin. Stratigraphic cross- sections and isopach maps demonstrate thickening of Pennsylvanian strata towards east of the El Oro-Rincon and thinning towards eastern and southern edge of the basin (Figs.

2.5-2.7). These facies relationships suggest a roughly north-south trending sediment depocenter during the Pennsylvanian that located close to the El Oro-Rincon uplift. Early

Permian strata demonstrate a shift of sediment depocenter from the Pennsylvanian, and onlap Precambrian basement-cored rocks of the Sierra Grande uplift and Cimarron Arch suggesting a major change in depositional style at the Pennsylvanian-Permian boundary.

A synthesis of available biostratigraphic data provides a scheme for basin-wide stratigraphic correlations (chapter-3). The basal, mostly carbonate strata of the Arroyo

Peñasco Group is Mississippian in age and the Sandia formation is late Morrowan through latest Atokan in age. Biostratigraphic data suggest that the carbonate dominated

Porvenir facies distribute from the earliest Desmoinesian through the late Desmoinesian in the greater Taos trough area. The Alamitos Formation ranges from the latest

100

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Desmoinesian through the Virgilian in age. The Sangre de Cristo Formation is

Wolfcampian in age in the Taos-Rainsville trough.

Stratigraphic records of climate-sensitive facies were utilized to reconstruct

Permo-Pennsylvanian paleoclimate of northern New Mexico, and these facies were used

as marker beds for correlating strata among wells assuming that climate-sensitive facies had minimal time-transgression (chapter-4). Abundant coals and the high clastic flux of the Sandia interval suggest an ever wet to seasonal climate during the Early

Pennsylvanian that became drier during Middle Pennsylvanian as evidenced from evaporite and carbonate deposition of the Porvenir Formation. High clastic flux and occasional coal layers in the Sangre de Cristo Formation indicate a seasonal to wet conditions during the Wolfcampian. The presence of wide-spread evaporite layers in the

Yeso Formation indicate a dry climate during the Leonardian.

A 3-D basin subsidence model was constructed through correlating time- equivalent surfaces of backstripped 1-D tectonic subsidence curves produced from subsurface sections of the Rainsville trough. Tectonic subsidence analysis and basin modeling suggest that the onset of basin subsidence was during the late Morrowan for both the Taos and Rainsville troughs. Peak subsidence of the Rainsville trough was in the late Morrowan-Atokan, whereas in the Taos trough peak subsidence was during the

Desmoinesian (Fig. 5.4). Subsidence of the Rainsville trough was largely controlled by the thrust-loaded flexure of the El Oro-Rincon uplift. Both the El Oro-Rincon and

Picuris-Pecos faults manifested as east-verging thrusts during the Morrowan-Atokan, but slip on the Picuris-Pecos fault expressed a large component of strike-slip during the

101

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Desmoinesian. During the Missourian-Virgilian, a relatively tectonic quiescent period in the region occurred with minimal subsidence throughout the greater Taos trough area.

Onlapping of early Permian strata and a renewed tectonic activity, consistent with a

different tectonic mechanism prevailed during the Permian than the Pennsylvanian

tectonic driver (Fig. 5.8).

102

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

REFERENCES CITED

Algeo, T. J., 1992, Continental-scale wrenching of southwestern Laurussia during the Ouachita-Marathon orogeny and tectonic escape of the Llano block: West Texas Geological Society Guidebook, v. 92–92, p. 115–131.

Allen, P. A., and Allen, J. R., 1990, Basin Analysis: Principles and Application to Petroleum Play Assessment: Blackwell Scientific Publications, 451p.

Angevine, C. L., Heller, P. L., and Paola, C., 1990, Quantitative Sedimentary Basin Modeling: Continuing Education Course Note Series, Tulsa, OK: American Association of Petroleum Geologists, v. 32.

Armstrong, A. K., 1955, Preliminary observations on the Mississippian system of Northern New Mexico: New Mexico Bureau of Mines and Mineral Resources Circular, v. 39, 42p.

Armstrong, A. K., and Mamet, B. L., 1967, and carbonate facies of the Mississippian Arroyo Peñasco Formation, north-central New Mexico: New Mexico Bureau of Mines and Mineral Resources Memoir 20, 80 p.

Armstrong, A. K., and Mamet, B. L., 1974, Biostratigraphy of the Arroyo Peñasco Group, Lower Carboniferous (Mississippian), north-central New Mexico Ghost Ranch: New Mexico Geological Society Guidebook, v. 25, p. 145-158.

Armstrong, A. K., and Mamet, B. L., 1979, The Mississippian System of north-central New Mexico; in Ingersoll, Raymond V., Woodward, Lee A., and James, H. L., eds., Santa Fe Country: New Mexico Geological Society Guidebook, v. 30, p. 201-210.

Armstrong, A. K., and Mamet, B. L., 1990, Stratigraphy, facies and paleotectonics of the Mississippian system, Sangre de Cristo Mountains, New Mexico and Colorado and adjacent areas: New Mexico Geological Society Guidebook, v. 41, 450 p.

Armstrong, A. K., Kottlowski, F. E., Siemers, W. T., Thompson III, S., and Baltz Jr, E. J., 1979, Mississippian and Pennsylvanian (Carboniferous) systems in the US- New Mexico: US Geological Survey Professional Paper (United States), 1110p.

Armstrong, A. K., Mamet, B. L., and Repetski, J. E., 2004, Mississippian system of New Mexico and adjacent areas: New Mexico Geological Society Special Publication, v. 11, p.77-93. 103

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Bachman, G. O., 1953, Geology of part of northwestern Mora County, New Mexico: U.S. Geological Survey Oil and Gas Investigation Map, OM-137.

Bachman, G. O., and Dane, C. H., 1962, Preliminary geologic map of the northeastern part of New Mexico: U.S. Geological Survey Miscellaneous Geological Investigation Map, I-358.

Baltz, E. H., and Read, C. B., 1960, Rocks of Mississippian and probable age in Sangre de Cristo Mountains, New Mexico: American Association of Petroleum Geologists Bulletin, v. 44, p. 1749-1774.

Baltz, E. H., and Myers, D. A., 1984, Porvenir Formation (new name) – and other revisions of nomenclature Mississippian, Pennsylvanian, and Lower Permian rocks, southeastern Sangre de Cristo Mountains, New Mexico: U. S. Geological Survey Bulletin, v. 1537-B, 39p.

Baltz, E. H., and O'Neill, J. M., 1986, Geologic map and cross sections of the Sapello River area. Sangre de Cristo Mountains, Mora and San Miguel Counties, New Mexico: US Geological Survey Miscellaneous Investigations Series Map I-1575.

Baltz, E. H., and Myers, D. A., 1999, Stratigraphic framework of upper Paleozoic rocks, southeastern Sangre de Cristo Mountains, New Mexico, with a section on speculations and implications for regional interpretation of Ancestral Rocky Mountains paleotectonics: New Mexico Bureau of Mines and Mineral Resources Memoir, v. 48, 269p.

Barbeau, D. L., 2003, A flexural model for the Paradox Basin: implications for the tectonics of the Ancestral Rocky Mountains: Basin Research, v. 15, no. 1, p. 97- 115.

Barrick, J.E., Lambert, L.L., Heckel, P.H., and Boardman, D.R., 2004, Pennsylvanian conodont zonation for Midcontinent North America: Revista Espanola de Micropaleontologia, v. 36, p. 231-250.

Barrick, J.E., Lambert, L.L., Heckel, P.H., Rosscoe, S.J., and Boardman, D.R., 2013a, Midcontinent Pennsylvanian conodont zonation: Stratigraphy, v. 10, p. 55-72.

104

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Barrick, J.E., Lucas, S.G., and Krainer, K., 2013b, Conodonts of the (uppermost Middle to Upper Pennsylvanian), Cerros de Amado region, central New Mexico, U.S.A., in Lucas, S.G., Nelson, W.J., DiMichele, W.A., Spielmann, J.A., Krainer, K., Barrick, J.E., Elrick, S., and Voigt, S., eds., The Carboniferous- Permian Transition in Central New Mexico: New Mexico Museum of Natural History and Science Bulletin 59, p. 239-252.

Benison, K.C., and Goldstein, R., 2001, Evaporites and siliciclastic of the Permian Nippewalla Group of , USA — a case for nonmarine deposition in saline lakes and saline pans: Sedimentology, v. 48, p. 165–188.

Berman, D. S., 1993, Lower Permian vertebrate localities of New Mexico and their assemblages: New Mexico Museum of Natural History and Science Bulletin, v. 2, p. 11-21.

Berman, D. S., Henrici, A. C., and Lucas, S. G., 2013, (Synapsida, Ophiacodontidae) from the Lower Permian Sangre de Cristo Formation of New Mexico: New Mexico Museum of Natural History and Science Bulletin, v. 60, p. 36-41.

Boardman, D.R., and Heckel, P.H., 1989, Glacial-eustatic sea-level curve for early Late Pennsylvanian sequence in northcentral Texas and biostratigraphic correlation with curve for midcontinent North America: Geology, v. 17, p. 802-805.

Boardman, D.R., Heckel, P.H., and Marshall, T.R., 2004, Preliminary report on lower Desmoinesian (mid-Moscovian) conodonts from lower and middle Cherokee Group of southern Midcontinent North America. I.U.G.S. Subcommission on Carboniferous Stratigraphy: Newsletter on Carboniferous Stratigraphy, v. 22, p. 41-47.

Bond, G. C., and Kominz, M. A., 1984, Construction of tectonic subsidence curves for the early Paleozoic miogeocline, southern Canadian Rocky Mountains: Implications for subsidence mechanisms, age of breakup, and crustal thinning: Geological Society of America Bulletin, v. 95, no. 2, p. 155-173.

Brotherton, J.L., Chowdhury, N.U.M.K., and Sweet, D.E., In Press, A synthesis of late Paleozoic sedimentation in central and eastern New Mexico: Implications for timing of Ancestral Rocky Mountains deformation: SEPM special publications.

105

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Brown, L. M., Rexroad, C. B., and Zimmerman, A.N., 2013, Conodont Biostratigraphy of the Porvenir Formation (Pennsylvanian, Desmoinesian) in the Southeastern Sangre de Cristo Mountains, New Mexico: The Mountain Geologist, v. 50, no. 3, p. 99-119.

Casey, J.M., 1980, Depositional systems and paleogeographic evolution of the late Paleozoic Taos trough, northern New Mexico: Rocky Mountain Section (SEPM).

Casey, J.M., 1981, Depositional systems and basin evolution of the late Paleozoic Taos trough, northern New Mexico: Unpublished PhD dissertation, University of Texas at Austin.

Casey, J.M., Scott, A.J., Ingersoll, R.V., Woodward, L.A., and James, H.L., 1979, Pennsylvanian coarse-grained fan deltas associated with the Uncompahgre uplift, Talpa, New Mexico: New Mexico Geological Society Guidebook, v. 30, p. 211- 218.

Cather, S. M., Karlstrom, K. E., Timmons, J. M., and Heizler, M. T., 2006, Palinspastic reconstruction of basement-related aeromagnetic features in north- central New Mexico: Implications for Mesoproterozoic to late Cenozoic tectonism: Geosphere, v. 2, p. 299–323.

Cather, S.M., Read, A.S., Dunbar, N.W., Kues, B.S., Krainer, K., Lucas, S.G., and Kelley, S.A., 2011, Provenance evidence for major post–early Pennsylvanian dextral slip on the Picuris-Pecos fault, northern New Mexico: Geosphere, v. 7, no. 5, p. 1175-1193.

Catuneanu, O., 2006, Principles of sequence stratigraphy (first edition): Amsterdam, Elsevier, 375p.

Cecil, C.B., 1990, Paleoclimate controls on stratigraphic repetition of chemical and siliciclastic rocks: Geology, v. 18, no. 6, p. 533-536.

Cecil, C.B., 2003, The concept of autocyclic and allocyclic controls on sedimentation and stratigraphy, emphasizing the climatic variable. In: Cecil, C.B., Edgar, T.N. (Eds.), Climate Controls on Stratigraphy: SEPM Special Publication, v. 77, p. 13–20.

Cecil, C.B., and Dulong, F.T., 2003, Precipitation models for sediment supply in warm climates. In: Cecil, C.B., Edgar, T.N. (Eds.), Climate Controls on Stratigraphy: SEPM Special Publication, v. 77, p. 21–27.

106

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Cecil, C. B., Stanton, R. W., Dulong, F. T., Rupert, L. F., and Renton, J. J., 1981, A geochemical model for the origin of low ash and low-sulfur coal: In Mississippian- Pennsylvanian Boundary in the Central part of the Appalachian Basin, Geological Society of America Field Trip No. 4: 175-7.

Cecil, C.B., Stanton, R.W., Neuzil, S.G., Dulong, F.T., Ruppert, L.F., and Pierce, B.S., 1985, Paleoclimate controls on late Paleozoic sedimentation and peat formation in the central Appalachian Basin (USA): International Journal of Coal Geology, 5(1- 2), p.195-230.

Cecil, C.B., Dulong, F.T.,West, R.R., Stamm, R.,Wardlaw, B., and Edgar, N.T., 2003, Climate controls on the stratigraphy of a Middle Pennsylvanian cyclothem in North America. In: Cecil, C.B., Edgar, N.T. (Eds.), Climate Controls on Stratigraphy. SEPM Special Publication, v. 77, p. 151–180.

Clark, K. F., 1966, Geology of the Sangre de Cristo Mountains and adjacent areas between Taos and Raton, New Mexico: New Mexico Geological Society Guidebook, v. 17, p. 57-65.

Davydov, V.I., Korn, D., and Schmitz, M.D., 2012, The Carboniferous period, in: Gradstein, F.M., Ogg, J.G., Schmitz, M.D., Ogg, G. (Eds.), The Geologic Time Scale 2012: Elsevier, Amsterdam, p. 603–651.

DeCelles, P. G., and Giles, K. A., 1996, Foreland basin systems: Basin research, v. 8, no. 2, p. 105-123.

Dickerson, P. W., 2003, Intraplate mountain building in response to continent–continent collision—the Ancestral Rocky Mountains (North America) and inferences drawn from the Tien Shan (Central Asia): Tectonophysics, v. 365, no. 1, p. 129-142.

Dickinson, W. R., 2000, Geodynamic interpretation of Paleozoic tectonic trends oriented oblique to the Mesozoic Klamath-Sierran continental margin in California; in Soreghan, M.J., and Gehrels, G.E., eds., Paleozoic and paleogeography and tectonics of western Nevada and northern California: Geological Society of America Special Paper, v. 347, p. 209–245.

Dickinson, W. R., and Lawton, T. F., 2001, Carboniferous to Cretaceous assembly and fragmentation of Mexico: Geological Society of America Bulletin, v. 113, p. 1142–1160.

107

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Dickinson, W. R., and Lawton, T. F., 2003, Sequential intercontinental suturing as the ultimate control for Pennsylvanian Ancestral Rocky Mountains deformation: Geology, v. 31, no. 7, p. 609-612.

Dickinson, W.R., Beard, L.S., Brakenridge, G.R., Erjavec, J.L., Ferguson, R.C., Inman, K.F., Knepp, R.A., Lindberg, F.A., and Ryberg, P.T., 1983, Provenance of North American Phanerozoic in relation to tectonic setting: Geological Society of America Bulletin, v. 94, p. 222-235.

DiMichele, Pfefferkorn, H.W., and Gastalod, R.A., 2001, Response of Late Carboniferous and Early Permian plant communities to climate change: Annual Review of Earth and Planetary Sciences, v. 29, p. 461–487.

DiMichele, W.A., Kerp, H., and Chaney, D.S., 2004, Tropical floras of the Late Pennsylvanian–Early Permian transition: Carrizo Arroyo in context. In: Lucas, S.G., Zeigler, K.E. (Eds.), Carboniferous–Permian Transition at Carrizo Arroyo, Central New Mexico: New Mexico Museum of Natural History and Science Bulletin, v. 25, p. 105–110.

DiMichele,W.A., Tabor, N.J., Chaney, D.S., and Nelson,W.J., 2006, From wetlands to wet spots: environmental tracking and the fate of Carboniferous elements in Early Permian tropical floras. In: Greb, S.F., DiMichele,W.A. (Eds.),Wetlands Through Time: Geological Society of America Special Paper, v. 399, p. 223–248.

DiMichele, W.D., Montañez, I.P., Poulsen, C.J., and Tabor, N.J., 2009, Feedbacks and regime shifts in the Late Palaeozoic ice-age earth: Geobiology, v. 7, p. 200–226.

Eble, C.F., 2002. Palynology of late Middle Pennsylvanian coal beds in the Appalachian Basin: International Journal of Coal Geology, v. 50, p. 73–88.

Eros, J. M., Montañez, I. P., Osleger, D. A., Davydov, V. I., Nemyrovska, T. I., Poletaev, V. I., and Zhykalyak, M. V., 2012, Sequence stratigraphy and onlap history of the Donets Basin, Ukraine: insight into Carboniferous icehouse dynamics: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 313, p. 1-25.

Geslin, J. K., 1998, Distal ancestral Rocky Mountain tectonism: Evolution of the Pennsylvanian-Permian Oquirrh-Wood River basin, southern Idaho: Geological Society of America Bulletin, v. 110, no. 5, p. 644–663.

108

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Giles, J.M., Soreghan, M.J., Benison, K.C., Soreghan, G.S., and Hasiotis, S.T., 2013, Lakes, loess, and paleosols in the Permian of Oklahoma, USA: implications for paleoclimate and paleogeography of the Midcontinent: Journal of Sedimentary Research, v. 83, no. 10, p.825-846.

Haq, B.U., and Schutter, S.R., 2008, A chronology of Paleozoic sea-level changes: Science, v. 322, no. 5898, p.64-68.

Heckel, P. H., 1999, Overview of Pennsylvanian (Upper Carboniferous) stratigraphy in Midcontinent region of North America. In: Heckel, P. H., Ed., Middle and Upper Pennsylvanian (Upper Carboniferous) cyclothem succession in Midcontinent Basin, USA: Kansas Geological Survey, Open-File Report 99–27, p. 68-102.

Heckel, P. H., 2002, Genetic stratigraphy and conodont biostratigraphy of upper Desmoinesian-Missourian (Pennsylvanian) cyclothem succession in Midcontinent North America. In: Hills, L. V., Henderson, C. M. and Bamber, E. W., Eds., The Carboniferous and Permian of the World, Calgary: Canadian Society of Petroleum Geologists, Memoir 19, p. 99–119.

Heckel, P. H., 2008, Pennsylvanian cyclothems in Midcontinent North America as far- field effects of waxing and waning of Gondwana ice sheets. In: Fielding, C. R., Frank, T. D. and Isbell, J. L., Eds., Resolving the Late Paleozoic Ice Age in time and space: Geological Society of America, Special Paper 441, p. 275–289.

Heckel, P.H., 2013, Pennsylvanian stratigraphy of Northern Midcontinent Shelf and biostratigraphic correlation of cyclothems: Stratigraphy, v. 10, no. 1–2, p.3-39.

Heckel, P.H., Boardman, D.R., and Barrick, J. E., 2002, Desmoinesian-Missourian regional stage boundary reference position for North America. In: Hills, L.V., Henderson, C.M. and Bamber, E.W., Eds., The Carboniferous and Permian of the World, Calgary: Canadian Society of Petroleum Geologists, Memoir 19, p. 710- 724.

Hegarty, K. A., Weissel, J. K., and Mutter, J. C., 1988, Subsidence history of Australia’s southern margin: constraints on basin models: American Association of Petroleum Geologists Bulletin, v. 72, no. 5, p. 615-633.

Henry, T.W., 1973, Brachiopod biostratigraphy and faunas of the Morrow Series (Lower Pennsylvanian) of northwestern Arkansas and northeastern Oklahoma: [Unpublished PhD dissertation], The University of Oklahoma, Norman, 515 p.

109

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Herrick, C. L., 1900, The Geology of the White Sands of New Mexico: Journal of Geology, v. 15, p. 805-816.

Hoy, R. G., and Ridgway, K. D., 2002, Syndepositional thrust-related deformation and sedimentation in an Ancestral Rocky Mountains basin, Central Colorado trough, Colorado, USA: Geological Society of America Bulletin, v. 114, no. 7, p. 804- 828.

Hunt, A. P., Lucas, S. G., and Huber, P., 1990, Early Permian footprint fauna from the Sangre de Cristo Formation of northeastern New Mexico: New Mexico Geological Society Guidebook, v. 41, p. 291-303.

Johnson, S. Y., Schenk, C. J., and Karachewski, J. A., 1988, Pennsylvanian and Permian depositional systems and cycles in the Eagle basin, northwest Colorado In G. S. Holden (Ed.), Geological Society of America Field Trip Guidebook, 1988: Professional Contributions, Colorado School of Mines, v. 12, p. 156–175.

Johnson, S. Y., Chan, M. A., and Konopka, E. A., 1992, Pennsylvanian and Early Permian paleogeography of the Uinta-Piceance basin region, northwestern Colorado and northeastern Utah: U.S. Geological Survey Bulletin, 1787-CC, CC1-CC35.

Kessler, J.L., Soreghan, G.S., and Wacker, H.J., 2001, Equatorial aridity in western Pangea: Lower Permian loessite and dolomitic paleosols in northeastern New Mexico, USA: Journal of Sedimentary Research, v. 71, no. 5, p.817-832.

Keyes, C. R., 1903, Geological sketches of New Mexico: Ores and Metals, v. 12, 48 p.

Kietzke, K. K., 1990, Microfossils from the Flechado Formation (Pennsylvanian, Desmoinesian) near Talpa, New Mexico: New Mexico Geological Society Guidebook, v. 41, p.259-276.

Kluth, C. F., 1986, Plate tectonics of the ancestral Rocky Mountains, in Peterson, J.A., eds., Paleotectonics and Sedimentation in the Rocky Mountain Region, United States: American Association of Petroleum Geologists Memoir, v. 41, p. 353-369.

Kluth, C. F., 1998, Late Paleozoic deformation of interior North America: The greater Ancestral Rocky Mountains: Discussion: American Association of Petroleum Geologists Bulletin, v. 82, p. 2272-2276.

110

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Kluth, C. F., and Coney, P. J., 1981, Plate tectonics of the Ancestral Rocky Mountains: Geology, v. 9, p. 10-15.

Krainer, K., Lucas, S. G., and Kues, B. S., 2004, Tectonically induced clasticcarbonate depositional sequences of the Pennsylvanian-Permian transition in the Rowe- Mora basin, northern New Mexico: New Mexico Geological Society Guidebook, v. 55, p. 314-325.

Kues, B. S., 1990, New and little known Middle Pennsylvanian Gastropods from the Flechado Formation, Taos County, New Mexico: New Mexico Geological Society Guidebook, v. 41, p. 251-258.

Kues, B. S., 2004, Pennsylvanian trilobites from the Sangre de Cristo and , north-central New Mexico in Geology of the Taos Region: New Mexico Geological Society Guidebook, vol. 55,326p.

Leary, R.J., Umhoefer, P., Smith, M. E., and Riggs, N., 2017, A three-sided orogen: A new tectonic model for Ancestral Rocky Mountain uplift and basin development: Geology, v. 45, p. 735-738.

Leeder, M. R., 1999, Sedimentology and Sedimentary Basins: Blackwell Science Ltd., 592 p.

Lucas, S. G., Spielmann, J. A., and Lerner, A. J., 2009, The Abo Pass tracksite: a Lower Permian tetrapod footprint assemblage from central New Mexico: New Mexico Geological Society Guidebook, v. 60, p. 285-290.

Lucas, S.G., Krainer, K., and Brose, R.J., 2013, The Lower Permian Glorieta Sandstone in central New Mexico. New Mexico Museum of Natural History and Science, Bulletin 59, p. 201-211.

Lucas, S.G., Krainer, K., Vachard, D., Voigt, S., DiMichele, W.A., Berman, D.S., Henrici, A.C., Schneider, J.W., and Barrick, J.E., 2015, Progress report on correlation of nonmarine and marine Lower Permian strata, New Mexico, USA: Permophiles.

Lucas, S. G., Krainer, K., Dimichele, W. A., Voigt, S., Berman, D. S., Henrici, A. C., Tanner, L. H., Chaney, D. S., Elrick, S. D., Nelson, W. J., and Rinehart, L. F., 2015, Lithostratigraphy, Biostratigraphy and Sedimentatology of the Upper Paleozoic Sangre de Cristo Formation, southwestern San Miguel County, New Mexico: New Mexico Geological Society Guidebook, v. 66, p. 211-228. 111

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Mack, G.H., and Bauer, E.M., 2014, Depositional environments, sediment dispersal, and provenance of the Early Permian (Leonardian) Glorieta Sandstone, central New Mexico: New Mexico Geological Society Guidebook, v. 65, p. 261-271.

Mack, G.H., and Dinterman, P.A., 2002, Depositional environments and paleogeography of the Lower Permian (Leonardian) Yeso and correlative formations in New Mexico: The Mountain Geologist.

Mallory, W. W., 1972, Pennsylvanian arkose and the ancestral Rocky Mountains; in Mallory, W. W. (ed.), Geologic atlas of the Rocky Mountain region: Rocky Mountain Association of Geologists, p. 131-132.

McConnell, D.A., 1989, Determination of offset across the northern margin of the Wichita uplift, southwest Oklahoma: Geological Society of America Bulletin, v. 101, no. 10, p. 1317-1332.

McElwain, J.C., Montañez, I., White, J.D., Wilson, J.P., and Yiotis, C., 2016, Was atmospheric CO2 capped at 1000 ppm over the past 300 million years?: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 441, p.653-658.

McKee, E. D., and Crosby, E. J., coordinators, 1975, Paleotectonic investigations of the Pennsylvanian System in the United States; Part 1, Introduction and regional analyses of the Pennsylvanian System, Chapters A through R: U.S. Geological Survey Professional Paper, v. 853, 349p.

Miller, J. P., Montgomery, A., and Sutherland, P. K., 1963, Geology of part of the southern Sangre de Cristo Mountains, New Mexico: New Mexico Bur. Mines and Min. Res., Mem. 11, 106 p.

Nelson, W.J., and Lucas, S.G., 2011, Carboniferous geologic history of the Rocky Mountain region, in Sullivan, R.M., Lucas, S.G., and Spielmann, J.A., eds., Fossil Record 3: New Mexico Museum of Natural History and Science Bulletin 53, p. 115-142.

Parrish, J. T., 1993, Climate of the supercontinent Pangea: The Journal of Geology, v. 101, no.2, p. 215-233.

Parrish, J. T., 1998, Interpreting Pre- climate from the geologic record: Columbia University Press, 338p.

112

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Pike, J. D., and Sweet, D. E., 2018, Environmental drivers of cyclicity recorded in lower Permian eolian strata, Manitou Springs, Colorado, western United States: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 499, p. 1-12.

Ramezani, J., and Bowring, S.A., 2018, Advances in numerical calibration of the Permian timescale based on radioisotopic : Geological Society, London, Special Publications, v. 450, no, 1, p.51-60.

Rascoe Jr, B., and Baars, D. L., 1972, Permian system: Geologic Atlas of the Rocky Mountain Region, USA: Rocky Mountain Association of Geologists, p. 143-165.

Read, C. B., and Wood, G. H., 1947, Distribution and correlation of Pennsylvanian rocks in late Paleozoic sedimentary basins of northern New Mexico: The Journal of Geology, v. 55, p. 220-236.

Retallack, C.J., and Huang, C., 2010, Depth to gypsic horizon as a proxy for paleoprecipitation in paleosols of sedimentary environments: Geology, v. 38, no. 5, p. 403–406.

Rinehart, L. F., Lucas, S. G., Tanner, L., Nelson, W. J., Elrick, S. D., Chaney, D. S., and DiMichele, W. A., 2015, Plant architecture and spatial architecture of an early Permian woodland buried by flood waters, Sangre de Cristo Formation, New Mexico: Paleoclimatology, Paleogeography, Paleoecology, v. 424, p. 91-110.

Rygel, M. C., Fielding, C. R., Frank, T. D., and Birgenheier, L. P., 2008, The magnitude of Late Paleozoic glacioeustatic fluctuations: a synthesis: Journal of Sedimentary Research, v. 78, no. 8, p. 500-511.

Scotese, C.R., 1999. PALEOMAP animations, paleogeography, PALEOMAP project: Arlington. University of Texas at Arlington, Department of Geology, Arlington.

Scotese, C. R., 1997, Paleogeographic Atlas, PALEOMAP Progress Report 90-0497: Arlington, Texas, Department of Geology, University of Texas at Arlington, 45p.

Seals, S.C., and Soreghan, G.S., 2013, Late Pennsylvanian sedimentation in the western Orogrande Basin (New Mexico): “Icehouse” sedimentation and rapid basin subsidence in the southern Ancestral Rocky Mountains, in Lucas, S.G., Nelson, W.J., DiMichele, W.A., Spielmann, J.A., Krainer, K., Barrick, J.E., Elrick, S., and Voigt, S., eds., Carboniferous-Permian Transition in Central New Mexico: New Mexico Museum of Natural History and Science Bulletin, v. 59, p. 331-346.

113

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Sinclair, H. D., and Naylor, M., 2012, Foreland basin subsidence driven by topographic growth versus plate subduction: Geological Society of America Bulletin, v. 124, no. 3-4, p. 368–379.

Soegaard, K., 1990, Fan-delta and braid-delta systems in Pennsylvanian Sandia Formation, Taos Trough, northern New Mexico: Depositional and tectonic implications: Geological Society of America Bulletin, v. 102, p. 1325-1343.

Soegaard, K., and Caldwell, K.R., 1990, Depositional history and tectonic significance of alluvial sedimentation in the Permo-Pennsylvanian Sangre de Cristo Formation, Taos trough, New Mexico: 41st Annual Field Conference Guidebook, p. 277-289.

Soreghan, G.S., 1992, Preservation and paleoclimatic significance of eolian dust in the Ancestral Rocky Mountains province: Geology, v. 20, no. 12, p.1111-1114.

Soreghan, G. S., 1994, The impact of glacioclimatic change on Pennsylvanian cyclostratigraphy: AAPG datapages, p. 523-543.

Soreghan, G.S., 1997, Walther's Law, climate change, and upper Paleozoic cyclostratigraphy in the Ancestral Rocky Mountains: Journal of Sedimentary Research, v. 67, no. 6, p.1001-1004.

Soreghan, G.S., Elmore, R.D., Katz, B., Cogoini, M., and Banerjee, S., 1997, Pedogenically enhanced magnetic susceptibility variations preserved in Paleozoic loessite: Geology, v. 25, no. 11, p.1003-1006.

Soreghan, G.S., Elmore, R.D., and Lewchuk, M.T., 2002, Sedimentologic-magnetic record of western Pangean climate in upper Paleozoic loessite (lower Cutler beds, Utah): Geological Society of America Bulletin, v. 114, no. 8, p.1019-1035.

Soreghan, M.J., Soreghan, G.S. and Hamilton, M.A., 2002, Paleowinds inferred from detrital-zircon geochronology of upper Paleozoic loessite, western equatorial Pangea: Geology, v. 30, no. 8, p.695-698.

Soreghan, G. S., Keller, G. R., Gilbert, M. C., Chase, C. G., and Sweet, D. E., 2012, Load-induced subsidence of the Ancestral Rocky Mountains recorded by preservation of Permian landscapes: Geosphere, v. 8, no. 3, p. 654-668.

114

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Soreghan, M. J., Heavens, N., Soreghan, G. S., Link, P. K., and Hamilton, M. A., 2014, Abrupt and high-magnitude changes in atmospheric circulation recorded in the Permian , tropical : Geological Society of America Bulletin, v. 126, no. 3-4, p. 569-584.

Steckler, M. S., and Watts, A. B., 1978, Subsidence of the Atlantic-type continental margin off New York: Earth and Planetary Science Letters, v. 41, no. 1, p. 1-13.

Sturmer, D.M., Trexler Jr, J.H. and Cashman, P.H., 2018, Tectonic Analysis of the Pennsylvanian Ely‐Bird Spring Basin: Late Paleozoic Tectonism on the Southwestern Laurentia Margin and the Distal Limit of the Ancestral Rocky Mountains: Tectonics, v. 37, no. 2, p. 604-620.

Sutherland, P. K., 1963, Paleozoic rocks; in Miller, J. P., Montgomery, A., and Sutherland, P. K., Geology of part of the southern Sangre de Cristo Mountains, New Mexico: New Mexico Bureau of Mines and Mineral Resources Memoir, no. 11, p. 22-44.

Sutherland, P. K., 1972, Pennsylvanian stratigraphy, southern Sangre de Cristo Mountains, New Mexico; in Mallory, W. W., (ed.), Geological atlas of the Rocky Mountain region: Rocky Mountain Association of Geologists, p. 139-142.

Sutherland, P. K., and Harlow, F. H., 1967, Late Pennsylvanian brachiopods from north- central New Mexico: Journal of Paleontology, v. 41, p. 1065-1089.

Sutherland, P. K., and Harlow, F. H., 1973, Pennsylvanian brachiopods and biostratigraphy in southern Sangre de Cristo Mountains, New Mexico: New Mexico Bureau of Mines and Mineral Resources Memoir, v. 27, 173 p.

Sweet, D.E., 2013, Load-induced subsidence model of the core Ancestral Rocky Mountains: The hypothesis and preliminary tests: Geological Society of America Abstracts with Programs, v. 45, no. 3.

Sweet, D. E., and Soreghan, G. S., 2010, Late Paleozoic tectonics and paleogeography of the ancestral Front Range: Structural, stratigraphic, and sedimentologic evidence from the (Manitou Springs, Colorado): Geological Society of America Bulletin, v. 122, no. 3-4, p. 575-594.

115

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Sweet, A.C., Soreghan, G.S., Sweet, D.E., Soreghan, M.J., and Madden, A.S., 2013, Permian dust in Oklahoma: source and origin for middle Permian (Flowerpot- Blaine) redbeds in western tropical Pangaea: Sedimentary Geology, v. 284, p. 181–196.

Sweet, D. E., Carsrud, C. R., and Watters, A. J., 2015, Proposing an Entirely Pennsylvanian Age for the Fountain Formation through New Lithostratigraphic Correlation along the Front Range: The Mountain Geologists, v. 52, no. 2, p. 43- 70.

Sweet, D. E., and Watters, A. J., 2015, Middle Pennsylvanian composite stratigraphic of Holman Grade records cyclic fluvio-deltaic sedimentation within the Taos trough axis: New Mexico Geological Society Guidebook, v. 66, p. 229-239.

Sweet, D.E., Chowdhury, N.U.M.K., and Brotherton, J.L., 2019, A fundamental change in subsidence style exists across the ancestral Rocky Mountains at the Pennsylvanian-Permian boundary: GSA Abstracts with Programs, Vol. 51, No. 5.

Tabor, N.J., and Montanez, I.P., 2002, Shifts in late Paleozoic atmospheric circulation over Western Equatorial Pangaea: insights from pedogenic mineral d18O compositions: Geology, v. 30, p. 1127–1130.

Tabor, N.J., and Montanez, I.P., 2004, Morphology and distribution of fossil soils in the PermoPennsylvanian Wichita and Bowie Groups, north-central Texas, USA: implications for western equatorial Pangean palaeoclimate during icehouse– greenhouse transition: Sedimentology, v. 51, p. 851–884.

Tabor, N.J., and Poulsen, C.J., 2008. Paleoclimate across the Late Pennsylvanian–Early Permian tropical paleolatitudes: a review of climate indicators, their distribution, and relation to paleophysiographic climate factors. Palaeobiology, Palaeoclimatology, Palaeoecology 268, p. 293–310.

Tabor, N.J., Montañez, I.P., Scotese, C.R., Poulsen, C.J., and Mack, G.H., 2008, Paleosol archives of environmental and climatic history in paleotropical western Pangaea during the latest Pennsylvanian through Early Permian. In: Fielding, C.R., Frank, T.D., Isbell, J.L. (Eds.), Resolving the Late Paleozoic Ice Age in Time and Space: Geological Society of America Special Paper, v. 441, p. 291–303.

Tabor, N.J., DiMichele, W.A., Montañez, I.P., and Chaney, D.S., 2013, Late Paleozoic continental warming of a cold tropical basin and floristic change in western Pangaea: International Journal of Coal Geology. 116

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Tramp, K.L., Soreghan, G.S. and Elmore, R.D., 2004, Paleoclimatic inferences from paleopedology and magnetism of the Permian Maroon Formation loessite, Colorado, USA: Geological Society of America Bulletin, v. 116, no. 5-6, p.671- 686.

Treat, C. A., 2014 Middle Pennsylvanian Conodonts of the Southern Sangre de Cristo Mountains, New Mexico: Unpublished MS thesis, Texas Tech University, 93p.

Voigt, S., and Lucas, S. G., 2015, On a diverse tetrapod ichnofauna from early Permian Red Beds in San Miguel County, North-Central New Mexico: New Mexico Geological Society Guidebook, v. 66, p. 241-252.

Wahlman, G.P., 2013, Pennsylvanian to Lower Permian (Desmoinesian-Wolfcampian) fusulinid biostratigraphy of Midcontinent North America: Stratigraphy, v. 10, p. 73-104.

Watters, A. J., 2014, Middle Pennsylvanian Paleogeographic and Basin Analysis of the Taos trough, northern New Mexico: Unpublished MS thesis, Texas Tech University, 126p.

Wawrzyniec, T. F., Ault, A. K., Geissman, J. W., Erslev, E. A., and Fankhauser, S. D., 2007, Paleomagnetic dating of fault slip in the southern Rocky Mountains, USA, and its importance to an integrated Laramide foreland strain field: Geosphere, v. 3, p. 16–25.

Webster, G. D., and Kues, B. S., 2006, Pennsylvanian crinoids of New Mexico: New Mexico Geology, v. 28, no. 1, p.3-36.

Woodward, L.A., Anderson, O.J., and Lucas, S.G., 1999, Late Paleozoic right-slip faults in the Ancestral Rocky Mountains: New Mexico Geological Society, Guidebook 50, p. 149-154.

Xie, X., and Heller, P. L., 2009, Plate tectonics and basin subsidence history, GSA Bulletin, v. 121, no. 1-2, p. 55-64.

Ye, H., Royden, L., Burchfiel, C., and Schuepbach, M., 1996, Late Paleozoic Deformation of Interior North America: The Greater Ancestral Rocky Mountains: American Association of Petroleum Geologists Bulletin, v. 80, no. 9, p. 1397- 1432.

117

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Yin, A., and Ingersoll, R. V., 1997, A Model for Evolution of Laramide Axial Basins in the Southern Rocky Mountains, U.S.A.: International Geology Review, v. 39, p. 1113-1123.

118

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

APPENDICES Appendix A

MATLAB code for lithologic plot clear all; close all; clc; %axis properties pba=[5 1 1];

%color brown=[0.6 0.2 0]; purple=[0.75 0 0.75]; %% %loading parameters % data=load("./data/19417_CNFD2_1.txt"); % data=load("./data/25804_CNFD_1.txt"); data=load("./data/30403_cnld_run1and3_Merged_1.txt"); depth=data(:,1); %ft GR=data(:,2); NPHI=data(:,3); PEF=data(:,4); DPHI=data(:,5);

%% %claculation %shale ind_sh=1; for k=1:length(GR)

if GR(k)>90 Shale(ind_sh)=2; depth_sh(ind_sh)=depth(k); ind_sh=ind_sh+1; end end %%Limestone ind_l=1; for k=1:length(GR) if GR(k)<60 && abs(abs(NPHI(k))-abs(DPHI(k)))<0.02 if PEF(k)>=4 Limestone(ind_l)=2; depth_l(ind_l)=depth(k); ind_l=ind_l+1;

end 119

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

end end

%sandstone ind_sn=1; for k=1:length(GR) if GR(k)<=70 if NPHI(k)0 if PEF(k)<2.4 Sandstone(ind_sn)=2; depth_sn(ind_sn)=depth(k); ind_sn=ind_sn+1; end

end end end

%%Sitstone ind_s=1; for k=1:length(GR) if GR(k)>70 && GR(k)<=90 if NPHI(k)>DPHI(k) Siltstone(ind_s)=2; depth_s(ind_s)=depth(k); ind_s=ind_s+1; end end end

%%Coal ind_c=1; for k=1:length(GR)

if NPHI(k)>=0.295 && DPHI(k)>=0.295 Coal(ind_c)=2; depth_c(ind_c)=depth(k); ind_c=ind_c+1; end end

%%Anhydrite ind_a=1; for k=1:length(GR)

if GR(k)<40 if DPHI(k)<=-0.08 Anhydrite(ind_a)=2; depth_a(ind_a)=depth(k); ind_a=ind_a+1; 120

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

end end end

%% fig1=figure(1); plot(depth(1:2:end),NPHI(1:2:end),'-') pbaspect(pba); for k=1:length(depth_s) a1=area(depth_s(k),Siltstone(k)); hold on; set(a1,{'facecolor'},{brown},{'edgecolor'},{brown}); pbaspect(pba); end for k=1:length(depth_sh) area(depth_sh(k),Shale(k),'facecolor','g','edgecolor','g'); hold on; pbaspect(pba); end for k=1:length(depth_sn) area(depth_sn(k),Sandstone(k),'facecolor','y','edgecolor','y'); hold on; pbaspect(pba); end for k=1:length(depth_c) area(depth_c(k),Coal(k),'facecolor','k','edgecolor','k'); hold on; pbaspect(pba); end for k=1:length(depth_l) b1=area(depth_l(k),Limestone(k)); hold on; set(b1,{'facecolor'},{purple},{'edgecolor'},{purple}); pbaspect(pba); end for k=1:length(depth_a) area(depth_a(k),Anhydrite(k),'facecolor','c','edgecolor','c'); pbaspect(pba); end set(gca,'ytick',[]); % ylim([0 0.3]) xlim([500 10000]); xlabel('Depth (ft)','fontsize',20); title('Well-1 SalmonA','fontsize',25); set(gca, 'FontSize', 15, 'LineWidth', 1); view([-90 -90])

121

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Appendix B

MATLAB code for producing subsidence curves using backstripping technique clear all; close all; clc;

%% paramerers %thickness=[45.72 42.672 890.6256 178.0032 156.6672 1222.8576 33.528]; % Glorieta, YEso, Sangre de cristo, Alamitos, Porvenier, Sandia, Arroyo Penasco % shale_mat=[0 0 19.4 35.7 25.5 39.6 12]; % silt_mat=[0 5 17.8 0 3 11.2 0]; % fosilm_mat=[ 0 7 2.1 0 4 3.4 68]; % marl_mat= [ 0 0 4.3 12.5 12.5 8 0]; % sand_mat=[100 88 56.5 51.8 55 37.8 20]; % micrite_mat=[0 0 0 0 0 0 0]; [numbers, TEXT, all]=xlsread("./excel/Salmon_Ranch_B.xlsx"); sand=all(:,7); sand_mat=cell2mat(sand(2:8)); shale=all(:,8); shale_mat=cell2mat(shale(2:8)); silt=all(:,9); silt_mat=cell2mat(silt(2:8)); marl=all(:,10); marl_mat=cell2mat(marl(2:8)); fosilm=all(:,11); fosilm_mat=cell2mat(fosilm(2:8)); micrite=all(:,12); micrite_mat=cell2mat(micrite(2:8));% % %% for k=1:length(sand_mat)

shale_ck=1429; silt_ck=3000; fosilm_ck=1800; marl_ck=5000; sand_ck=2500; micrite_ck=2457;

shale_rhm=2.7; silt_rhm=2.65; fosilm_rhm=2.8; marl_rhm=2.87; sand_rhm=2.66; micrite_rhm=2.86;

shale_phi=0.52; silt_phi=0.5; 122

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

fosilm_phi=0.42; marl_phi=0.41; sand_phi=0.34; micrite_phi=0.3;

shale1=shale_mat(k); silt1=silt_mat(k); fosilm1=fosilm_mat(k); marl1= marl_mat(k); sand1=sand_mat(k); micrite1=micrite_mat(k);

% % %% variables % ck_ratio(k)=(shale1*shale_ck+silt1*silt_ck+fosilm1*fosilm_ck+marl1*marl _ck+sand1*sand_ck+micrite1*micrite_ck)/100; rhm(k)=(shale1*shale_rhm+silt1*silt_rhm+fosilm1*fosilm_rhm+marl1*marl_r hm+sand1*sand_rhm+micrite1*micrite_rhm)/100; phi(k)=(shale1*shale_phi+silt1*silt_phi+fosilm1*fosilm_phi+marl1*marl_p hi+sand1*sand_phi+micrite1*micrite_phi)/100; end thickness=all(:,5); % Glorieta, YEso, Sangre de cristo, Alamitos, Porvenier, Sandia, Arroyo Penasco thickness=cell2mat(thickness(2:8)); % rhm_M=3.33; rhm_W=1.028; % layer=7; %number of layers you want to use for m=1:layer

if m==1 for k=1:length(thickness) cum=cumsum(thickness(1:k))-thickness(k)/2; z(k)=cum(end); phi_i=phi(k)*exp(-z(k)/ck_ratio(k)); phii(k,m)=phi_i; phi_d(k)=phi_i; z_d(k,m)=thickness(k)*(1-phi_i)/(1-phi(k)); end phi_d=phi_d(2:end); else thick=z_d(2:end-(m-2),m-1);

for k=1:length(thick) cum=cumsum(thick(1:k))-thick(k)/2; z(k)=cum(end); phi_i=phi_d(k); 123

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

phii(k,m)=phi_i; phi_d(k)=phi(k+1+m-2)*exp(-z(k)/ck_ratio(k+1+m-2)); %m=2 k=k+1// m=3 k=k+2, and it follows z_d(k,m)=thick(k)*(1-phi_i)/(1-phi_d(k)); end phi_d=phi_d(2:end); end % % end % final_zd=z_d(1,:); % % save("decompacted","final_zd"); % xlswrite('decompacted_30403_Salmon Ranch_A_V2.xlsx',final_zd); % s_tot=final_zd; for k=1:length(s_tot) if isnan(s_tot(k)) s_tot(k)=0;

end end phi_i_cal=phii(1,:); rho_s=(1-phi_i_cal).*rhm+phi_i_cal*rhm_W; s_iso=s_tot.*(rho_s-rhm_W)./(rhm_M-rhm_W); delsl=[0 0 0 37.5 25 0 15]; wd=[0 19 35 35 35 35 10]; s_tec=s_tot-s_iso+wd-delsl.*rhm_M./(rhm_M-rhm_W); % % % % s_tec=s_tec; s_tec_cum=cumsum(fliplr(s_tec)); s_tot_cum=cumsum(fliplr(s_tot)); % % Age=fliplr([266 277 290.5 302.5 306 312 323]); % % %% plot fig1=figure(); % % plot(Age,s_tec_cum,'r-', 'linewidth',2); hold on; plot(Age,s_tot_cum,'b-', 'linewidth',2) set(gca,'xcolor','black','ycolor','black'); set(gca, 'FontSize', 15, 'LineWidth', 1,'fontweight','normal'); % set(gca,'YDir','reverse'); set(gca, 'XAxisLocation', 'top'); set(gca, 'YAxisLocation', 'right'); set(gca,'XDir','reverse'); % ylabel('Subsidence (m)'); xlabel('Millions of years'); title({"Geohistory Curves-Rainsville Trough",'Salmon Ranch B'}); 124

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019 legend("S-tec","S-tot"); % % % axis square; xlim([260 340]) ylim([0 4000]); xticks([260 280 300 320 340]); yticks([0 1000 2000 3000 4000]);

% view(-90,-90) %

% saveas(fig1,'./figure_1213/Salmon Ranch A.pdf'); saveas(fig1,'./figure/Salmon Ranch B.png'); % saveas(fig1,'./figure_1213/Salmon Ranch A.tiff');

125

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

Appendix C

MATLAB code for producing 3-D structure map

The following code is for Sandia structure map. clear all; close all; clc; %% [numbers, TEXT, all]=xlsread("./data/decompacted surface depth.xlsx"); lat=all(:,3); lat_new=cell2mat(lat(2:33)); long=all(:,4); long_new=cell2mat(long(2:33)); sandia=all(:,10); sandia_new=cell2mat(sandia(2:33));

% fig1=figure(); for k=1

X=lat_new; Y=long_new; Z=sandia_new; createFit1(Y, X, Z); hold on; fig1=figure(1); % txt=sprintf(''); % text(X,Y,Z,txt,'fontsize',15); % ylim([34 36]); % xlim([-106 -104]); % zlim([-200 2200]); % ylabh = get(gca,'ylabel'); % set(ylabh,'Units','normalized'); % % set(ylabh,'position',[0 0.8 0]); xlabel('lat','fontsize',12); % ylabel('long','fontsize',15); ylabel('long','HorizontalAlignment','right','FontSize',12); zlabel('surface depth','fontsize',12); title({" 3-D Structure map ",'Sandia'},'FontSize',15) set(gca,'ZDir','reverse'); colormap (flipud(jet)); h=colorbar; ylabel(h,"Blue: deep, Red: shallow",'fontsize',12) view(-70,65) %camroll(270) % axis square; 126

Texas Tech University, Nur Uddin Md Khaled Chowdhury, December 2019

% saveas(fig1,'./figure_1213/Sandia.pdf'); saveas(fig1,'./figure_1213/Sandia.png'); %saveas(fig1,'./figure_1213/Yeso.tiff'); end

127