3 S.Lynn Peyton, Barbara Carrapa, 2013, An overview of low-temperature thermochronology in the Rocky Mountain and Its application to petroleum system analysis, in C. Knight and J. Cuzella, eds., Application of structural methods to Rocky Mountain hydrocarbon exploration and development: AAPG Studies in Geology 65, p. 37–70. An Overview of Low-temperature Thermochronology in the Rocky Mountains and Its Application to Petroleum System Analysis S. Lynn Peyton Coal Creek Resources Inc., 1590 S. Arbutus Pl., Lakewood, Colorado, 80228, U.S.A. (e-mail: slpeyton@coalcreekresources.com) Barbara Carrapa Department of Geosciences, University of Arizona, 1040 E. 4th St., Tucson, Arizona, 85721, U.S.A. (e-mail: bcarrapa@email.arizona.edu) ABSTRACT A synthesis of low-temperature thermochronologic results throughout the Laramide foreland illustrates that samples from wellbores in Laramide basins record either (1) detrital Laramide or older cooling ages in the upper ~1 km (0.62 mi) of the wellbore, with younger ages at greater depths as temperatures increase; or (2) Neogene cooling ages. Surface samples from Laramide ranges typically record either Laramide or older cooling ages. It is apparent that for any particular area the complexity of the cooling history, and hence the tectonic history interpreted from the cooling history, increases as the number of studies or the area covered by a study increases. Most Laramide ranges probably experienced a complex tectono-thermal evolution. Deriving a regional timing sequence for the evolution of the Laramide basins and ranges is still elusive, although a compilation of low-temperature thermochronology data from ranges in the Laramide foreland suggests a younging of the ranges to the south and southwest. Studies of subsurface samples from Laramide basins have, in some cases, been integrated with and used to constrain results from basin burial -history modeling. Current exploration for unconventional shale-oil or shale-gas plays in the Rocky Mountains has renewed interest in thermal and burial history modeling as an aid in evaluating thermal maturity and understanding petroleum systems. We suggest that low-temperature thermochronometers are underutilized tools that can provide additional constraints to burial -history modeling and source rock evaluation in the Rocky Mountain region. Copyright ©2013 by The American Association of Petroleum Geologists. DOI:10.1306/13381689St653578 37 10711_ch03_ptg01_hr_037-070.indd 37 6/5/13 7:59 AM 38 Peyton and Carrapa INTRODUCTION Low-temperature thermochronologic techniques have many uses relevant to petroleum exploration. Ageelevation profiles can provide the timing and rate of cooling; inverse and forward modeling of thermochronometer ages can help to constrain the thermal history of an area; ages can be used to constrain basin models; the base of a partial annealing zone (PAZ) or partial retention zone (PRZ) can be used as a structural marker for mapping; and detrital thermochronology can illuminate sediment provenance and unroofing history. These techniques and methods are discussed in detail in Chapter 2. We strongly recommend that readers unfamiliar with thermochronologic techniques and terminology read Chapter 2 (Peyton and Carrapa, 2013). In the last decade, industry focus on unconventional petroleum plays has led to the emergence of shale formations such as the Niobrara Formation, Mowry Shale, Mancos Shale, and Bakken Formation, to name a few, as important exploration targets within the Rocky Mountain region. Exploration success in these shale plays is often dependent upon a clear ­understanding of source-rock maturation and thermal history. Vitrinitereflectance (%Ro) and Rock-Eval pyrolysis studies have typically been used to evaluate the thermal maturity of source rocks, and used as input into basin modeling in the Rocky Mountain region (e.g., Hagen and Surdam, 1984; Dickinson, 1986; Nuccio, 1994a; Roberts et al., 2008). We suggest that low-temperature thermochronology should be used to provide additional constraints on the thermal evolution of these source rocks because it has the potential to provide information on the maximum temperature reached, the time when that maximum temperature was reached, and the timing, rate, and amount of cooling. Note however that source rocks are probably too fine grained to yield usable apatite or zircon, and samples should be taken from coarser grained rocks in the stratigraphic section. To date, a few papers have examined the thermal and petroleum history of the Sevier fold-thrust belt using fission-track thermochronology (Burtner and Nigrini, 1994; Osadetz et al., 2004), but there are no published studies that comprehensively document the use of low-temperature thermochronology for petroleum exploration in the Laramide foreland (Figure 1) or that integrate results with burial-history modeling and other thermal indicators (such as %Ro and fluid inclusions) to understand the thermal evolution and maturity of resource plays in this area. Thus, lowtemperature thermochronology techniques, such as apatite fission-track dating (AFT), zircon fission-track dating (ZFT), apatite (U-Th)/He dating (AHe), and zircon (U-Th)/He dating (ZHe), are underutilized 10711_ch03_ptg01_hr_037-070.indd 38 tools in the exploration for oil and gas in the region. AFT and AHe dating have been used elsewhere throughout the world as tools for petroleum exploration, and in particular for basin modeling studies (e.g., Duddy et al., 1991; Duddy et al., 1994; Duddy et al., 1998; Crowhurst et al., 2002; Osadetz et al., 2002; Osadetz et al., 2004; Parnell et al., 2005). Although AHe and ZHe dating have been successfully used to elucidate shallow crustal processes in many parts of the North American cordillera (e.g., House et al., 1997; Stockli et al., 2000; Cecil et al., 2006; Flowers et al., 2007; Flowers et al., 2008), there are only two published studies within the Laramide foreland (Figure 1) (Crowley et al., 2002; Peyton et al., 2012). In contrast, there have been many AFT and ZFT studies in the Laramide foreland, because the earlier development of these techniques resulted in their widespread use (Figure 1) (Giegengack et al., 1986; Naeser, 1986; Shuster, 1986; Naeser, 1989; ­C erveny, 1990; Kelley and Blackwell, 1990; Naeser, 1992; ­C erveny and Steidtmann, 1993; Omar et al., 1994; Chen and Lubin, 1997; Kelley and Chapin, 1997; Pazzaglia and Kelley, 1998; Kelley, 2002; ­Naeser et al., 2002; ­Kelley and Chapin, 2004; Kelley, 2005). The goals of this chapter are to review and summarize all low-temperature thermochronology studies of the Laramide Rocky Mountain region to date, and then to illustrate how these techniques may be applied to petroleum exploration using examples from elsewhere. After a brief overview of the geologic history of the Rocky Mountain region, we review the vitrinitereflectance technique and discuss the reliability and relevance of old thermochronologic data. We divide our summary of low-temperature thermochronologic work in the Laramide Rocky Mountains into two sections. First, we look at low-temperature thermochronologic studies of the Laramide basins, which date subsurface samples from petroleum wellbores (Figure 1). Most of these samples reached maximum burial after the Laramide Orogeny but before Neogene exhumation, and provide information about Neogene cooling and basin exhumation. Using AFT ages from the Green River ­Basin (Naeser, 1986, 1989; Naeser et al., 1989b) we illustrate how forward modeling of AFT ages and vitrinite reflectance data can be used to evaluate published results and interpretations. Second, we discuss studies that have dated samples from the Laramide ranges (Figure 1). Samples are largely from the Precambrian crystalline cores of the ranges, which probably reached maximum burial during the Cretaceous, followed by exhumation and cooling during the Laramide Orogeny, resulting in Late Cretaceous and Paleogene rather than Neogene cooling ages. The high topographic relief of many of the ranges led workers to analyze samples from elevation transects. In areas with less topographic 6/5/13 7:59 AM An Overview of Low-temperature 39 112 108 Crazy Mountains Basin 104 SOUTH DAKOTA d BT a a TR BH a,b e OC Wind River a Basin a WR a,c,l,p g f c front LR Greater Green River Basin thru st NEBRASKA a GM HU c a c c,k LARAMIDE ROCKY MOUNTAINS UM MB PR COLORADO c,k c m w Denver Basin Uinta Basin r n GR WRU Piceance Basin r q r FR q,r,s,t,u SR 112 WM AHe subsurface samples SC AFT subsurface samples Precambrian crystalline basement TA NEW MEXICO t COLORADO PLATEAU 112 t,u v AFT surface samples 100 s,u o UU AHe surface samples 0 BL i c c,l,p c,h Powder River Basin g 40 40 Bighorn Basin 44 Peyton et al. (2012) Crowley et al. (2002) Cerveny (1990) Omar et al. (1994) Roberts & Burbank (1993) Naeser (1986,1989); Naeser et al. (1989) g: Beland (2002) h: Strecker (1996) i: Naeser (1992) k: Kelley (2005) l: Cerveny & Steidtmann (1993) m: Painter (p.c., 2012) IDAHO n: Kelley & Blackwell (1990) UTAH o: Kelley (2002) p: Shuster, 1986 q: Bryant & Naeser (1980) r: Naeser et al. (2002) s: Kelley & Chapin (1997) Sevier t: Pazzaglia & Kelley (1998) u: Kelley & Chapin (2004) v: Lindsey et al. (1986) w: Constenius & Kelley (p.c. 2012) WYOMING 4 44 4 44 LP a: b: c: d: e: f: MONTANA a,c a c,d CB MR t N SN 200 km ARIZONA 108 104 Figure 1. Regional map showing approximate locations of AFT (red) and AHe (blue) studies, along with basins and ranges discussed in this chapter. See individual papers for more specific sample location information. Gray shading depicts region of Sevier foreland basin deformed during the Laramide Orogeny, referred to as the Laramide Rocky Mountains. Stippled areas show crystalline basement outcrop. Thick dashed line is outline of the Colorado Plateau. Thin dashed line represents the Cheyenne Belt. Abbreviations as follows: BL, Black Hills; BH, Bighorn Range; BT, Beartooth Range; CB, Cheyenne Belt; FR, Front Range; GM, Granite Mountains; GR, Gore Range; HU, Hartville Uplift; LP, Lima Peaks; LR, Laramie Range; MB, Medicine Bow Mountains; MR, Madison Range; OC, Owl Creek Mountains; PR, Park Range; SC, Sangre de Cristo Range; SN, Sierra Nacimiento; SR, Sawatch Range; TA, Taos Range; TR, Teton Range; UM, Uinta Mountains; UU, Uncompahgre Uplift; WM, Wet Mountains; WR, Wind River Range; WRU, White River Uplift; p.c., personal communication. 100 km (62 mi). 10711_ch03_ptg01_hr_037-070.indd 39 6/5/13 7:59 AM 40 Peyton and Carrapa relief, studies have used a wide areal distribution of samples to interpret the structure of the Precambrian basement. The Colorado Plateau (Figure 1) also experienced deformation during the Laramide Orogeny, resulting in a series of monoclines and flexures with much less structural relief than the foreland ranges. Although the Colorado Plateau has been studied using both AFT and AHe dating (Naeser et al., 1989a; Kelley et al., 2001; Naeser et al., 2001; Flowers et al., 2007; Flowers et al., 2008), this chapter focuses on results from the Laramide foreland (Figure 1). Finally, we discuss the workflow used by many workers for integrating low-temperature thermochronology results into petroleum studies, and summarize some studies from other parts of western North America, such as the Canadian thrust belt (Osadetz et al., 2004), and the Sevier thrust belt of Idaho and Wyoming (Burtner and Nigrini, 1994). BACKGROUND Geologic History During the Late Jurassic, subduction along the western margin of North America consolidated into a continuous orogen and magmatic arc, with deformation propagating eastward to form the Sevier retroarc thrust belt and its associated foreland ­basin (DeCelles, 2004). In the United States section of the orogen, Late Cretaceous flattening of the subducting Farallon slab caused a cratonward sweep of magmatism coeval with propagation of deformation into the Sevier foreland basin (Coney and ­R eynolds, 1977; Dickinson and Snyder, 1978; Constenius, 1996; Saleeby, 2003). Although thin-skinned thrusting continued in the Sevier fold-thrust belt, deformation in the foreland, known as the Laramide Orogeny, had a thick-skinned, basement-involved style. These basement-involved thrusts dissected the foreland basin into a series of smaller basins and ranges bound by an anastamosing system of faults (Brown, 1988; Erslev, 1993), herein referred to as the Laramide Rocky Mountains (Figure 1). This thick-skinned deformation persisted until Eocene time (as did Sevier thrusting), when the subducting slab rolled back and magmatism migrated westward (Coney and Reynolds, 1977; Dickinson and Snyder, 1978; Constenius, 1996). Shortening in each Laramide range was on the order of several kilometers, and was accommodated on major rangebounding thrusts that dip ~30° beneath the ranges and extend at least into the midcrust (Smithson et al., 1978). ­Total shortening across the foreland was ~45 km 10711_ch03_ptg01_hr_037-070.indd 40 (~27.9 mi), or ~12–15% (Brown, 1988; Stone, 1993). In comparison, shortening across the ­U tah-Wyoming part of the Sevier thrust belt was at least 100 km (62.1 mi), or 50% (Royse et al., 1975; Coogan, 1992; Yonkee, 1992). Foreland basin ­sedimentary rocks were eroded from the uplifting basement blocks during the Laramide Orogeny and were deposited in adjacent basins (Dickinson et al., 1988). Basin-bounding, proximal conglomerates represent the erosion of more resistant Paleozoic rocks from the ranges (e.g., DeCelles et al., 1987; Perry, Jr. et al., 1988; DeCelles et al., 1991; Hoy and Ridgway, 1998). Dickinson et al. (1988) studied Laramide ­b asin fill and proposed nearly-synchronous onset of ­L aramide deformation (~75 to 65 Ma) across the foreland that terminated earlier in the northern foreland (~50 Ma) than the southern (~35 Ma) foreland. Gries (1983) suggested that north-south trending Laramide structures become younger from west (Campanian) to east (Paleocene), that northwest trending structures are Paleocene in age, and that east-west trending structures formed during Eocene time, and supported her hypothesis with ages from basin fill. DeCelles (2004) recognized an eastward sweep of deformation beginning in Campanian time and ending in Eocene time, based on proximal, syntectonic sedimentary rocks associated with several Laramide uplifts, and suggested the Laramide ­region may simply be the frontal part of the Cordilleran orogenic wedge. Post-Laramide burial is evidenced by remnants of Oligocene through Pliocene sedimentary rocks that are preserved at high elevations in some ranges. Rocks of early Miocene age that crop out on the crest of the Bighorn Range at ~3 km (~1.86 mi) elevation led McKenna and Love (1972) to reconstruct cross sections of Eocene, Oligocene, and Miocene basin fill across the Bighorn and Powder River basins. These cross sections indicated that the thickness of sedimentary rocks in the basins during Miocene time may have been more than 1 km (0.6 mi) greater than present. Love (1960) stated that by the end of Oligocene time only the upper ~0.3 to 1.3 km (~0.1 to 0.8 mi) of the highest peaks in the major Laramide ranges were exposed above the aggradational plain. These high-elevation outcrops of Oligocene and ­M iocene sedimentary rocks suggest that after the Laramide Orogeny, the Rocky Mountains were buried to high levels and have since been exhumed (Mackin, 1947; Love, 1960; McKenna and Love, 1972; Gregory and Chase, 1994; McMillan et al., 2006; Riihimaki et al., 2007). The cause, timing, and amount of Cenozoic burial and exhumation are still controversial. 6/5/13 7:59 AM An Overview of Low-temperature 41 Overview of Vitrinite Reflectance (%Ro) Vitrinite is an organic component of coal. Vitrinite reflectance (%Ro) is the percentage of incident light that is reflected off a polished particle of vitrinite, and is dependent upon the maximum paleotemperature of the sample and the time spent at that temperature (e.g., Senftle and Landis, 1991). %Ro tends to increase with increasing burial in a sedimentary basin and is commonly used to evaluate organic maturity in sedimentary rocks (e.g., Tissot and Welte, 1978). Burnham and Sweeney (1989) and Sweeney and Burnham (1990) presented a kinetic model for the evolution of %R o with temperature and time and showed that temperature has a larger influence on %Ro than time. Duddy et al. (1991) showed that the annealing kinetics of fission tracks in apatite (Laslett et al., 1987) are similar to the kinetics of %Ro (Burnham and Sweeney, 1989) and that a given degree of annealing in apatite will always be associated with the same value of %Ro. For example, a %Ro value of 0.7 is associated with total annealing of all fission tracks in Durango fluorapatite (Duddy et al., 1994). Uncertainties associated with measured %Ro values are typically about ±0.1% in shale, and ±0.05% in coal (B. Claxton, personal communication, 2012). AFT and %Ro techniques complement each other because %Ro values can be used to determine maximum paleotemperatures when temperatures were greater than the annealing temperature of fission tracks, while AFT data provide information on the timing and rate of cooling from the maximum paleotemperature that %Ro cannot. Both techniques provide information on the thermal history of a sample. Evaluation of Existing Thermochronologic Studies Many of the studies discussed in this chapter are more than 20 years old, over which time thermochronologic techniques have evolved significantly. Early fissiontrack researchers did not measure track lengths or recognize the influence of kinetic parameters such as apatite composition on annealing properties (e.g., Bryant and Naeser, 1980). For several years researchers used electron microprobes to analyze apatite composition for a few grains from a few samples, and applied these results to all samples in the study (e.g., Kelley and Blackwell, 1990). The diameter of a fissiontrack etch pit (Dpar) was recognized as an indicator of annealing kinetics by Donelick (1993). Some studies estimated Dpar qualitatively, rejecting grains with wide etch-pit diameters (e.g., Kelley, 2005), but now Dpar is measured routinely and used to determine annealing kinetics. Apatite which anneals rapidly (e.g., 10711_ch03_ptg01_hr_037-070.indd 41 fluorapatite) typically has Dpar values <~2 mm. Higher Dpar values are associated with annealing-resistant apatite (e.g., chlorapatite). Early AFT papers did not report details of the methodology used to etch the fission tracks. The concentration and type of acid, and the time allowed for etching, may be quite variable and influence the number of fission tracks visible for counting (Murrell et al., 2009). In recent years, etching has tended to follow a standard recipe (20 s with 5.5 M nitric acid at 21°C). Older studies did not correct track lengths for their orientation with respect to the crystallographic c-axis, making track-length measurements less reliable. Although AHe dating is a much newer technique, it has also experienced a rapid evolution in recent years. The recognition that grain size and radiation damage can affect AHe ages (Reiners and Farley, 2001; Shuster et al., 2006; Flowers et al., 2009) has led to different approaches to interpret AHe ages (e.g., Flowers et al., 2008). However, unless there were problems with the laboratory, the fundamental measurements of parent and daughter nuclides, and the resulting AHe age, are likely sound. The biggest issue with older studies is that very few grains or aliquots (typically two) were measured per sample, making it difficult to recognize the influence of radiation damage or grain size. In ­areas where radiation damage is suspected to be a problem, such as cratonic basement, at least several grains or aliquots should be dated per sample. If He implantation is suspected, the only recourse at present is to discard the aliquot or the entire sample. We anticipate that in the future this will be mitigated by routinely abrading grains (either mechanically or chemically) to remove the outer part of the crystal ­affected by He ­implantation before dating. In summary, because AFT dating techniques have developed over the last 30+ years, AFT ages from more recent studies can be considered to be more accurate and reliable than those from older studies. With AHe dating, measurement techniques have not changed significantly, but our understanding of how to interpret AHe ages has evolved. All of the studies discussed in this chapter, along with the methods each study used, have been summarized in Table 1 to allow the reader to evaluate the relevance of the ­results from each paper. We have also included a summary of each paper’s conclusions, and our opinion of the data quality (Table 1). We emphasize that the authors were using the techniques available to them at the time, and that our comments in no way reflect on the quality of their work. The improvement in the reliability and accuracy of AFT ages through time results from advancements in the ­understanding of the technique. 6/5/13 7:59 AM 42 Peyton and Carrapa Table 1. Summary of previous low-temperature thermochronologic studies in the Laramide Rocky Mountain region. Study Laramide Basins Beland (2002) Method Location Sample Type Lithologies AFT Track Lengths? AFT Track Lengths Corrected to c-axis? AFT, AHe Casper Arch thrust and Maverick Springs thrust, Green River Basin, Wyoming cuttings crystalline basement, sedimentary rocks Yes Measured parallel to c-axis Chen and Lubin (1997) ZFT Bighorn Basin, Wyoming surface sedimentary rocks NA NA Kelley (2002) AFT, ZFT Denver Basin, Colorado core sedimentary rocks Not documented Not documented Kelley and Blackwell (1990) AFT, ZFT Piceance Basin, Colorado core sandstone Yes No Naeser (1986, 1989), Naeser et al. (1989) AFT, ZFT Green River Basin, Wyoming core sandstone No - due to low track density and low apatite yield NA Naeser (1992) AFT Powder River Basin, Wyoming core sandstone Yes Measured parallel to c-axis 10711_ch03_ptg01_hr_037-070.indd 42 6/5/13 7:59 AM An Overview of Low-temperature 43 AFT Annealing Properties: Composition or Dpar? AFT Etching Technique (for apatite to be dated) AHe Max # Aliquots Forward or Inverse Modeling? Summary/Conclusions Comments electron microprobe, four samples Not documented 14 (single & multicrystal) Forward modeling Miocene AFT and AHe ages shallow, bimodal AFT and AHe ages deep (>2.5 km [>1.5 mi]). Burial by thrusting during the late Miocene? Results are probably reliable but are inconsistent with other studies and difficult to explain geologically. More work needed in these areas. NA NA NA No ZFT ages not reset by burial. Stratigraphic ages refined using ZFT ages. Insufficient number of grains dated per sample for a thorough provenance study (maximum of 14). No 25 s in 5 M nitric acid NA No Used Gaussian peak-fitting methods to distinguish age populations. ZFT and AFT ages document a predictable unroofing sequence for the Front Range. AFT and ZFT ages not documented in the paper. Don’t know if sufficient apatite grains were measured for a detrital study. Apatite annealing kinetics not discussed. electron microprobe, 4 samples 20 s in 5 M nitric acid NA inverse Rapid cooling during last 10 Ma due to downcutting of Colorado River. Background heat flow between 10 Ma and present was similar to today. Maximum paleotemperatures of samples < 240 °C. Counted a max of 20 grains/sample - modern studies would count more grains. Ages and conclusions probably reliable, but only measured composition on four samples. No Not documented NA No Latest phase of rapid cooling of 20°C or more was initiated ~4 to 2 Ma. No annealing of ZFT since deposition. Ages from this study should be used with caution. It is difficult to interpret ages with no track lengths or kinetic data. Conclusions are tested in this paper using forward modeling. Discarded apatite with anomalouslywide etch pits Not documented NA Yes. No details provided Samples cooled from above annealing temp. at ~12 Ma. AFT data indicate higher paleotemperatures than %Ro data. Discrepancy may be due to fluid flow. Ages probably reliable for fluorapatite. Present-day temperatures estimated from a geothermal gradient and probably not accurate. Cooling at 12 Ma determined from plotting ages from a large area (~90 3 40 km [~55.9 3 24.85 mi]) on the same age-elevation profile. (continues) 10711_ch03_ptg01_hr_037-070.indd 43 6/5/13 7:59 AM 44 Peyton and Carrapa (continued) Study Method Location Sample Type Lithologies AFT Track Lengths? AFT Track Lengths Corrected to c-axis? Laramide Ranges Bryant and Naeser (1982) AFT Front Range, Sawatch Range, Colorado surface crystalline basement No NA Cerveny (1990) AFT, ZFT Multiple ranges— Wyoming, Colorado, South Dakota and southern Montana surface, core crystalline basement Yes Measured parallel to c-axis Cerveny and Steidtmann (1993) AFT, ZFT Wind River Range, Wyoming surface, core crystalline basement Yes Measured parallel to c-axis Kelley and Chapin (1997) AFT Front Range, Colorado surface crystalline basement Yes Not documented Kelley and Chapin (2004) AFT Front Range, Colorado surface crystalline basement Yes Yes Kelley (2005) AFT Laramie, Park, Sierra Madre and Medicine Bow Ranges, Wyoming and Colorado surface crystalline basement Yes Yes Naeser et al. (2002) AFT, ZFT White River Uplift, Gore Range, Front Range, Colorado surface crystalline basement, sandstone Yes Not documented Omar et al. (1994), Giegengack et al. (1986) AFT Beartooth Range, Wyoming and Montana surface, cuttings crystalline basement, sedimentary rocks Yes No 10711_ch03_ptg01_hr_037-070.indd 44 6/5/13 7:59 AM An Overview of Low-temperature 45 AFT Annealing Properties: Composition or Dpar? AFT Etching Technique (for apatite to be dated) AHe Max # Aliquots Forward or Inverse Modeling? Summary/Conclusions Comments No Not documented NA No Using base of fossil PAZ, determined ~6.5 km (~4 mi) of uplift of Front Range between Cretaceous and present day. AFT ages from Sawatch Range suggest Eocene uplift during Laramide Orogeny, and Miocene uplift along eastern edge of range due to Rio Grande rifting. Ages should be used with caution. No track length or kinetic data. No 30–45 s in 7% nitric acid NA No Broad study covering many ranges. Wind River Range has complex cooling history. Cooling began in Wind River Range by ~85 Ma. Most rapid cooling in Wind River, Beartooth and Park Ranges between ~62 and 57 Ma. Cooling ~40 Ma at toe of Wind River thrust. Ages probably reliable, but no measure of annealing properties. No 30-45 s in 7% nitric acid NA See summary above for Cerveny (1990) Ages probably reliable, but no measure of annealing properties. Not documented Not documented NA No Mapped base of PAZ. More exhumation in center of Front Range than at edges. No table with ages and track length data. Many details of methodology not documented. Electron microprobe, 6 samples 25 s in 5 M nitric acid NA inverse AFT ages used to map structure of Precambrian basement and amount of denudation. Laramide cooling 75-45 Ma. Reliable ages. Qualitative Dpar only: unusually large etch pits not observed 25 s in 5 M nitric acid NA No AFT ages used to map structure of Precambrian basement and amount of denudation. Proterozoic structures reactivated by Laramide tectonics in some areas, but not others. Ages probably reliable, but no measure of annealing properties. Dpar and microprobe 40 s in 7% nitric acid at 24 °C NA Yes. Inverse? ZFT ages have not been reset since the Proterozoic. AFT ages from the central and eastern Front Range and White River Uplift record Laramide cooling. The Gore Range and western Front Range record ~30 Ma cooling related to the Rio Grande Rift. Ages probably reliable, but orientation of tracks to c-axis not documented. No Not documented NA inverse No annealing kinetic data. Rapid uplift of ~4 to 8 km (~2.4 to 4.9 mi) began ~61 Ma. Second Track lengths not corrected for orientation. episode of ~4 km (~2.4 mi) of uplift began between 15 and 5 Ma. (continues) 10711_ch03_ptg01_hr_037-070.indd 45 6/5/13 7:59 AM 46 Peyton and Carrapa (continued) Study Method Location Sample Type Lithologies AFT Track Lengths? AFT Track Lengths Corrected to c-axis? Pazzaglia and Kelley (1998) AFT Southern Rocky Mountains, Colorado and New Mexico surface crystalline basement — see comments Peyton et al. (2013) AHe, AFT Beartooth, Wind River, Laramie and Beartooth Ranges, Wyoming and Montana surface, cuttings crystalline basement Yes Yes Roberts and Burbank (1993) AFT Teton Range, Wyoming surface crystalline basement Yes No Shuster (1986) AFT Wind River Range, Green River Basin surface, core crystalline basement No NA Steidtmann et al. (1989) AFT Wind River Range, Wyoming surface crystalline basement — see comments Strecker (1996) AFT, ZFT Black Hills, South Dakota surface crystalline basement Yes No? Crowley et al. (2002) AHe Bighorn Range, Montana and Wyoming surface crystalline basement NA NA 10711_ch03_ptg01_hr_037-070.indd 46 6/5/13 7:59 AM An Overview of Low-temperature 47 AFT Annealing Properties: Composition or Dpar? AFT Etching Technique (for apatite to be dated) AHe Max # Aliquots Forward or Inverse Modeling? — — NA Dpar 20 s in 5.5 M nitric acid at 21°C No Summary/Conclusions Comments No Integrated geomorphology and AFT data to test different models for evolution of topography. Data support significant Laramide crustal thickening followed by low rates of exhumation and uplift. Used a compilation of previously-published AFT data. 10 (single & multicrystal) forward and inverse Radiation damage affecting AHe ages. Inverted AHe age effective U pairs to find best-fit thermal history. Exhumation likely started earlier in the Bighorn Range (before ~71 Ma) than in the Beartooth Range (before ~58 Ma). He implantation affecting some AHe ages. Important to date a sufficient number of apatite grains to recognize radiation damage and He implantation. Authors did not consider temperature change due to elevation or climate change. Not documented NA No Most AFT ages record Laramide cooling between 85 and 65 Ma. Second episode of exhumation during Miocene. Less exhumation in south relative to north. No annealing kinetic data. Tracks lengths not corrected for orientation. No Not documented NA No Uplift of Wind River Range began no later than 78 Ma and continued through late Eocene. Ages should be used with caution. No track length or kinetic data. Cerveny (1990) dated the same samples and measured track lengths. — — NA No Recognized differential exhumation of fault blocks from core of Wind River Range using AFT data. Used data from Shuster (1996) - see comments above. No. Recognized that granitic samples had lower AFT ages than metasedimentary samples, probably due to apatite chemistry. 15-18 s in 5% nitric acid at room temperature NA No Exhumation and cooling ~66-52 Ma. Increase in cooling rate ~58 Ma. AFT ages define an anticline, with older AFT ages at edge of range. No annealing kinetic data, but divided samples into suites based on rock type, which were interpreted separately. Track lengths not corrected for orientation - interpret lengths with caution. NA NA 2 (multicrystal) No Samples were from a fossil AHe PRZ. Either insufficient burial during Cretaceous to reset ages, low geothermal gradient, or diffusion properties of He in apatite different to Durango apatite. Insufficient aliquots measured per sample to recognize large scatter due to radiation damage. Peyton et al. (2012) later showed that radiation damage had increased the temperature of the PAZ, and that geothermal gradients may have been very low. 10711_ch03_ptg01_hr_037-070.indd 47 6/5/13 7:59 AM 48 Peyton and Carrapa LARAMIDE BASINS Subsurface well samples have been used to study the thermal evolution of Laramide basins. Unlike the ranges, which reached their maximum burial immediately before the onset of the Laramide Orogeny, the basins reached maximum burial during the Neogene, and have subsequently experienced some amount of erosional exhumation (Love, 1960; Dickinson, 1986; Nuccio, 1994b; Roberts et al., 2008). Most basinal thermochronologic studies have documented the timing of post-Laramide cooling and maximum burial of the basins using AFT dating (Naeser, 1986, 1989; Kelley and Blackwell, 1990; Naeser, 1992). All detrital ZFT ages reported from Laramide basins are coincident with, or older than, their stratigraphic age, implying that there was insufficient burial and heating during the Cenozoic to reset the ZFT ages, that is, temperatures were less than ~240°C (Naeser, 1986; Chen and Lubin, 1997). Green River Basin Naeser (1986; 1989) and Naeser et al. (1989b) presented results from AFT and ZFT analysis of samples from the Wagon Wheel #1 well (Section 5, Township 30 N, Range 108 W, Sublette County, Wyoming) on the Pinedale Anticline in the northern Green River Basin (approximate location on Figure 1). AFT length measurements were not possible due to low track density and low apatite yield. Annealing kinetics such as Dpar or apatite composition were not measured. The shallowest sample dated was from a depth of 2237 m (7340 ft). Interpretation of the AFT ages led the authors to suggest that the most recent phase of cooling of at least 20°C initiated ~4 to 2 Ma during Pliocene time (Naeser, 1986, 1989; Naeser et al., 1989b). This interpretation was based on AFT ages between ~4 and 2 Ma that were found between temperatures of ~95°C and 120°C in the wellbore. Naeser (1989) noted that this cooling is consistent with other evidence for widespread erosion and uplift in the Green River Basin starting in the late Pliocene, although she did not estimate an erosion amount from this cooling. All ZFT ages suggested that well samples did not reach temperatures high enough to anneal fission tracks in zircon. Dickinson (1986) presented results of burial history modeling using %Ro data and the Lopatin model of Waples (1980) for the northern Green River Basin, including the Wagon Wheel #1 well. Dickinson (1986) required his burial models to have ~20–30°C of cooling between 3 Ma and present to be consistent with Naeser’s (1986) conclusions. He noted that 10711_ch03_ptg01_hr_037-070.indd 48 a component of this cooling was from a surface temperature decrease of ~10°C between the Pliocene and present caused by surface uplift of the region (Dorr et al., 1977; Dickinson, 1986). However, Dickinson (1986) used uncorrected present-day borehole temperatures from well logs, which may be several degrees cooler than actual temperatures, and thus his model may include more cooling than necessary. By projecting the %Ro data to a value of 0.2% (the normal assumed Ro value for newly deposited sediments), Dickinson (1986) suggested that 1520 m (4,986 ft) of section has been eroded from the surface at the Wagon Wheel #1 well location, although stratigraphic data indicate much less erosion (~470 m [~1541 ft]). He proposed that a 2°C/km increase in the geothermal gradient between Eocene and Pliocene time could result in this disparity, ­although he acknowledged that there is no direct evidence for this increase. Using modern modeling tools (e.g., HeFTy, Ketcham, 2005; Ketcham et al., 2007) it is possible to reevaluate the interpretation of these young AFT ages. Because these samples reside within the present-day PAZ the ages should be interpreted with caution. Inverse modeling of AFT data is not possible without track-length data, so forward modeling was used to test the validity of Naeser ’s (1986) proposed interpretation. Data relevant to the thermal history of the Wagon Wheel #1 well, such as %Ro, depositional age (Dickinson, 1986) and AFT data (Naeser, 1986, 1989) were compiled (Table 2). Measured values of %Ro (Dickinson, 1986) suggest that burial paleotemperatures were high enough to anneal fluorapatite samples (%Ro > 0.7 for all samples, Table 2) (Duddy et al., 1994) unless they were resistant to annealing. The ranges of measured AFT ages and individual grain ages within each sample (Table 2) suggest that either not all detrital grains were totally reset by burial, and therefore many grains must be resistant to annealing, or that some %Ro values were less than 0.7 due to the large uncertainty (up to +/– 0.1%) in Ro values. The latter case is unlikely because all but the shallowest two samples have %Ro > 0.8. Naeser (1986) dated between six and 11 apatite grains per sample, which is an insufficient number to recognize multiple detrital populations (Galbraith, 2005). Frequency histograms of individual-grain ZFT ages indicated that most detrital zircon grains (and hence we assume detrital apatite grains) cooled (1) between ~100 Ma and time of deposition (i.e., during the Laramide and Sevier orogenies), (2) ~300 Ma (during the Ancestral Rockies Orogeny), or (3) ~600 Ma (during Proterozoic rifting) (Naeser, 1986, 1989). The present-day corrected temperature within the borehole (Naeser, 1989) and approximate 6/5/13 7:59 AM 10711_ch03_ptg01_hr_037-070.indd 49 73 81 2465 2731 3098 3361 3370 4022 4557 4904 WWH-9 NN1 WWH-13 WWH-16 WWH-24 WWH-26 WWH-21 NN2 128 121 110 96 96 90 81 74 68 Temperature Corrected °C Rock Springs Rock Springs Ericson Lance Lance Lance Lance Lance Tertiary Formation 82.5 79 74 71 71 70.5 70 69 63 Depositional Age (Ma) AFT data from Naeser (1986, 1989). Depositional ages and %Ro data from Dickinson (1986). 100 m (328 ft). 122 114 102 84 84 67 61 2237 WWH-5 Temperature Uncorrected °C Depth (m) Sample Name Table 2. Compilation of data from the Wagon Wheel #1 well in the Green River Basin. 1.63 1.49 1.13 1.03 0.96 0.95 0.85 0.78 0.72 %Ro 0 2.6 2.4 3.8 8 14.3 36.5 36.3 20.9 AFT Age (Ma) 0 2.3 1.7 2.7 7.3 5.6 8.9 8.4 4.6 AFT Error (Ma) 6 9 11 6 6 9 8 10 8 # Grains NA 0-10 0-36 0-17 0-19 0-35 NA 6-81 9-67 Grain Age Range (Ma) An Overview of Low-temperature 49 6/5/13 7:59 AM 50 Peyton and Carrapa depositional age (Dickinson, 1986) are known for each sample ­(Table 2). Assumptions for the forward model included (1) an average surface temperature of 15°C during the Cretaceous (Amiot et al., 2004), (2) heating of rocks due to burial between the time of their deposition and 50 Ma, (3) no heating or cooling between 50 and 4 Ma (an unrealistic but simplifying assumption), (4) paleogeothermal gradients from 50 Ma to present the same as the present-day geothermal gradient, and (5) detrital grains remained at surface temperatures after initial cooling until deposition in these rock layers. We constructed a time-temperature model in HeFTy for sample WWH-9 of Naeser (1986), and adjusted the maximum temperature of the model between 50 and 4 Ma to provide a match to the %R o data ­(Figure 2A). HeFTy uses the kinetics of Sweeney and Burnham (1990) to calculate a %R o from a model time-temperature path. This model time-temperature path was then extrapolated to other samples using the present-day geothermal gradient between 50 and 0 Ma, and time of deposition adjusted for each sample. %Ro values were calculated for each time-temperature path and compared to measured values (Figure 2B). For all samples but one, calculated %R o values are within 0.06% of measured %R o ­v alues, leading us to conclude that the calculated values provide a good match to the measured values ­( Figure 2B). With no AFT kinetic data available, we calculated AFT ages for Dpar values between 1.65 mm (default value of HeFTy) and 5 mm, and calculated AFT ages for apatite that cooled immediately before deposition (i.e., volcanic origin, [Figure 3A]), at ­115–100 Ma (Figure 3B), and at 315–300 Ma (Figure 3C) to represent different source terranes. Model results (Figures 2 and 3) show that Naeser’s (1986, 1989) conclusion of at least 20°C of cooling between 4 Ma and present in the Wagon Wheel #1 well is reasonable if the apatites dated were resistant to annealing, as suggested by the %Ro data. Model results also show that AFT ages of detrital grains that have experienced the same thermal history since deposition can vary greatly depending upon the annealing properties and provenance of the apatite, and illustrate the importance of measuring kinetic data such as composition or Dpar, track lengths, and a sufficient number of grains to recognize detrital populations. Inclusion of %Ro data is critical in this case for estimation of maximum burial temperature and recognition of the presence of annealing-resistant apatite. We have not tested other possible time-temperature paths here, but recognize that there are almost certainly multiple forward models that will fit the data. It is possible, for example, that temperatures between 50 and 4 Ma could be adjusted in the forward model (e.g., to represent less burial/lower temperatures in the Eocene, and more burial/higher temperatures between the Oligocene and Pliocene), and still produce the same results. The key is to find a timetemperature path that gives both a good mathematical fit to the thermochronologic data and that corresponds to a plausible geological model. Figure 2. (A) Time-temperature paths for three possible Powder River Basin apatite grains from sample WWH-9, a volcanic grain which cooled as it was deposited (solid line), a detrital grain which cooled 115–100 Ma (long dashed line), and a detrital grain which cooled 315–300 Ma (short dashed line). (B) Actual measured %Ro data (open circles) (Dickinson, 1986) versus calculated %Ro from forward ­thermal modeling (black triangles). 1 km (0.6 mi). 10711_ch03_ptg01_hr_037-070.indd 50 Naeser (1992) calculated AFT ages for 13 core samples from 11 wells in the southwestern Powder River Basin (Figure 1). These Cretaceous-age samples were from sandstones in the Lewis Shale, the Parkman Sandstone Member of the Mesaverde 6/5/13 7:59 AM An Overview of Low-temperature 51 2000 Dpar = 1.65 mm = 3 mm r D pa 3000 r 4 = mm D pa 5 Grain age = Formation age pa r = 3500 D Depth (m) 2500 mm A 4000 4500 AFT age (Naeser, 1989) 5000 0 10 20 30 40 50 60 70 80 90 100 AFT age (Ma) 2000 Dpar = 1.65 mm B 2500 D par =3 mm Depth (m) 3000 D par =4 mm 3500 D pa r = 5 mm 100 Ma detrital grains 4000 4500 AFT age (Naeser, 1989) 5000 0 10 20 30 40 50 60 70 80 90 100 AFT age (Ma) 2000 Dpar = 1.65 mm C 2500 D par = 3 mm Depth (m) 3000 D par = 3500 4 mm 300 Ma detrital grains 4000 4500 AFT age (Naeser, 1989) 5000 0 10 20 30 40 50 60 70 80 90 100 AFT age (Ma) Figure 3. Results of forward modeling for apatite grains which (A) cooled as they were deposited (i.e., volcanic grains), (B) cooled 115–100 Ma, and C) cooled 315–300 Ma. Solid and dashed lines show ages predicted for different Dpar values. Black circles are AFT ages of Naeser (1986, 1989), error bars are 2s. Horizontal gray bars are the range of individual grain AFT ages for each sample (Naeser, 1986). 1000 m (3280 ft). 10711_ch03_ptg01_hr_037-070.indd 51 Formation, the Shannon Sandstone bed of the Steele Member of the Cody Shale, and the Frontier Formation. Lacking downhole temperature data, Naeser (1992) calculated present-day sample temperatures using a mean geothermal gradient from geothermal gradient maps. Kinetic data were not measured directly, but instead were estimated qualitatively from the width of etched tracks; grains with anomalously wide tracks (and hence high annealing temperatures) were discarded. Between four and thirteen grains were counted per sample. Track lengths were measured for 6 of the 13 samples, but other samples had insufficient tracks for counting. AFT ages were younger than depositional ages at two standard deviations, indicating that all sandstones had experienced significant annealing. For samples from similar temperatures and depths, Naeser (1992) created composite track-length distributions and concluded that much of the complexity in the six track-length distributions was due to the small number of tracks measured. Naeser (1992) noted that all samples between present-day temperatures of ~70°C and 110°C have AFT ages of ~12 Ma, and concluded that all samples were heated to high enough temperatures to almost totally anneal before cooling began ~12 Ma. Naeser (1992) observed that the time and rate of cooling could not be determined precisely because (1) it is possible that not all grains were totally annealed before cooling, (2) analytical uncertainties in the AFT ages were large, (3) some Cl-rich grains may have been included, and (4) the deeper samples were from the present-day PAZ. In contrast to the northern Green River Basin, the southwestern Powder River Basin has %R o values that are anomalously low (~0.6% near the basin axis) compared to other organic geochemical data (Merewether and Claypool, 1980) and the maximum temperatures suggested from AFT ages (Nuccio, 1990; Naeser, 1992). AFT data suggest cooling of at least 35°C starting at ~12 Ma (Naeser, 1992), but geological data imply that no more than ~630 m (~2066 ft) of post-mid-Miocene rock has been eroded (Nuccio, 1990). Naeser (1992) suggested that either geothermal gradients were higher in the past, or that hightemperature fluids moved through the sandstones and influenced apatite annealing. These two different possibilities have very different implications for source-rock maturity and petroleum generation in the Powder River Basin. The low %Ro values suggest that it is unlikely that geothermal gradients were higher in the past, because %R o would also have been affected. Thus, hot fluids may have annealed apatite grains in porous and permeable sandstones 6/5/13 7:59 AM 52 Peyton and Carrapa but left impermeable vitrinite-bearing layers unaffected. Another possible explanation is that %R o values may have been underestimated (Price and Barker, 1985). Age (Ma) 0 100 200 300 400 500 600 700 0 Casper Arch AFT 1000 Bighorn Basin Wind River Basin Beland (2002) studied two boreholes that drilled through Precambrian basement carried on thrusts at the northwestern (American Quasar 9-24 Tribal well, Section 9, Township 42 N, Range 105 W, Fremont County, Wyoming) and eastern (W.A. Moncrief #16-1 Tepee Flats well, Section 16, Township 37 N, Range 86 W, Natrona County, Wyoming) boundaries of the Wind River Basin (these wells penetrated the Maverick Springs thrust and Casper Arch thrust, respectively). AFT ages and track lengths were measured in both wells, and AHe ages in the Maverick Springs well only. Apatite composition was measured for four samples using an electron microprobe. Beland (2002) only measured the lengths of confined fission tracks that were parallel to the crystal c-axis. Interestingly, both wells do not follow the typical trend of thermochronometer age decreasing with depth. Results from the Maverick Springs well show young (Miocene) AFT and AHe ages at depths less than ~2.5 km, and bimodal ages (AFT <15 and >300 Ma; AHe <10 and >175 Ma) below ~2.5 km (~1.5 mi) (Figure 4). AFT ages from the Casper Arch well are 10711_ch03_ptg01_hr_037-070.indd 52 Maverick Springs AHe 2000 Depth (m) Chen and Lubin (1997) used the ZFT technique to date sedimentary rock samples from the Upper Jurassic Morrison and Lower Cretaceous Cloverly formations from surface outcrop in the Bighorn Basin. They ­divided their zircon samples into those from volcanic ash beds (bentonites), and those from detrital samples. With one exception, ZFT ages of volcanic origin were consistent with their corresponding stratigraphic ages, indicating that samples had not reached sufficiently high temperatures since deposition to partially anneal the fission tracks in the zircons. Using detrital ZFT age spectra, Chen and Lubin (1997) concluded that the amount of ash input into the Bighorn Basin intensified ~175 Ma and that this increase lasted at least until 115 Ma, through Morrison and Cloverly deposition. Using the youngest detrital age in the population, they refined stratigraphic ages of the detrital samples. ZFT age spectra indicated that the Morrison and Cloverly formations received sediment from multiple source ­areas of different ages. Maverick Springs AFT 3000 4000 5000 6000 Figure 4. Age-elevation plot showing AFT (green) and AHe (blue) ages for a well from the Maverick Springs Dome (squares) at the northern boundary of the Wind River Basin, and a well from the Casper Arch (triangles) at the eastern boundary of the Wind River Basin. From Beland (2002). Error bars are 2s. 1000 m (3280 ft). similar, with young ages (<7 Ma) above ~3.5 km (~2.1 mi) depth, and older ages (>300 Ma) below ~4 km (~2.4 mi) (Figure 4). Using forward modeling to match AFT ages and track length distributions, Beland (2002) suggested the most plausible explanation for these results was two episodes of thrusting, the first during the Laramide Orogeny which was followed by basin filling, and the second during ­Miocene time (<15 Ma), followed by erosion. The second episode of thrusting likely reactivated existing Laramide thrusts and created new thrust imbricates. The older AFT and AHe ages at depth indicate that these samples were not buried deeply enough by Phanerozoic sedimentary rocks for ages to be completely reset. All young, shallow AFT ages in Beland’s (2002) study are from apatite grains that have Fl-rich (i.e., ­Cl-poor) compositions, and thus anneal at lower temperatures than the Cl-rich apatite with Paleozoic ages from deeper in the wells (Green et al., 1986). Bimodal AFT ages (<15 Ma and >300 Ma) from the lower three samples in the Maverick Springs well could be due to contamination from shallow in the well, since the young grains are also Fl-rich (Beland, 2002). However, 6/5/13 7:59 AM An Overview of Low-temperature 53 as Beland (2002) noted, the deeper samples may contain different apatite compositions that have different AFT annealing temperatures. If the young AFT ages in the deep samples were due to different annealing properties, then it might not be necessary to invoke Miocene thrusting to explain the young ages at shallow depths. However, a similar bimodal pattern is seen in the AHe ages from the Maverick Springs well, even though the closure temperature of the AHe system is unaffected by apatite composition (Warnock et al., 1997). It is possible that old AHe ages were a result of He implantation, which has recently been suggested as a cause of anomalously old AHe ages in Precambrian basement rocks in the Laramide foreland (Reiners et al., 2008; Peyton et al., 2012). Beland (2002) argued that the younger, deep AFT ages are unlikely to be simply an effect of annealing temperature, because the Cl-rich apatites with Paleozoic ages were at temperatures where total annealing should have occurred. Thus, he concluded that the deep, Paleozoicage samples had not been exposed to their present-day high temperatures for very long, consistent with burial by thrusting during the late Miocene. Similarly, even if the shallower, young ages were caused by hot fluid flow, it is still difficult to explain why the deep samples were not annealed, given their present-day temperatures. Beland (2002) cited several examples from the Rocky Mountain foreland where Neogene shortening has been documented, and suggested that extensive erosion across the region may have resulted in few ­examples being preserved or identified. Such late thrusting at the edges of the Laramide basins would have significantly impacted the thermal maturity and migration history of any potential subthrust ­petroleum play, although it is unlikely it would have affected the basin centers. Other low-temperature thermochronology studies that have analyzed samples from wells which penetrate Precambrian basement on basinbounding thrusts have not found inverted cooling age profiles with young AFT or AHe ages at shallower depths than older ages (Cerveny and ­S teidtmann, 1993; Omar et al., 1994; Peyton et al., 2012). ­Further work at the boundaries of the Wind River ­B asin is warranted. Colorado) (Figure 1). Seventeen samples were from sandstone core in the Upper Cretaceous Mesaverde Group, while two were from sandstone cuttings from the Wasatch Formation. Seven of the core samples contained sufficient apatite to measure track lengths. Kelley and Blackwell (1990) used an electron microprobe to analyze the compositions of apatite from four samples but found no strong correlations between Cl content and AFT age. Track lengths were not corrected for their orientation with respect to the apatite c-axis. AFT ages ranged from 42.8 ± 8.2 Ma for the shallowest sample, to 1.0 ± 0.2 Ma for the deepest sample (Figure 5). All AFT ages were younger than the stratigraphic ages of the formations, indicating that all samples resided within a PAZ at some time. The uppermost three samples represent a fossil PAZ ­(Figure 5) with a marked change in slope at a presentday estimated temperature of ~66°C (1390 m depth [4560 ft]) representing the base of the fossil PAZ. ­I nversion modeling of AFT ages and track lengths showed that cooling of 10–15°C/Ma occurred in the last 5 Ma (Kelley and Blackwell, 1990). ZFT ages were not reset, indicating that temperatures never Piceance Basin Kelley and Blackwell (1990) determined AFT and ZFT ages of 19 samples from three wells <55 m (180.4 ft) apart at the Multi-Well Experiment site in the eastcentral Piceance Basin, northwestern Colorado (Section 34, Township 6 S, Range 94 W, Garfield County, 10711_ch03_ptg01_hr_037-070.indd 53 Figure 5. AFT results from the Piceance Basin (Kelley and Blackwell, 1990). 1000 m (3280 ft). 6/5/13 7:59 AM 54 Peyton and Carrapa reached 240 ± 25°C for any length of time. Kelley and Blackwell (1990) compared their fission-track results to results of burial history modeling. %Ro and burial data indicate that maximum paleotemperature at the bottom of the wellbore was ~150°C to 200°C at ~10 Ma, and that 1500 m (4,921 ft) of sedimentary rocks were eroded since 10 Ma by downcutting of the Colorado River (Bostic, 1983; Bostic and Freeman, 1984), consistent with the AFT and ZFT results. Denver Basin Kelley (2002) measured AFT and ZFT ages of detrital grains from synorogenic sediments from two wells in the Denver Basin (Section 17, Township 8S, Range 63W, Elbert County, Colorado, and Section 9, Township 7S, Range 67W, Douglas County, Colorado) ­(Figure 1). These shallow (<1 km depth [<0.6 mi]) samples were not buried deeply enough for AFT ages to have been reset during the Neogene. Age populations represent a combination of cooling ages of the source terrane, in this case the Front Range, and a volcanic component derived from ash blown from the west. Detrital material varied from recycled Mesozoic and Paleozoic sedimentary grains to grains derived from the Precambrian basement. In both wells, the percentage of metamict zircon grains increased up section, indicating an ­increase through time in the contribution of grains from the Proterozoic basement as it was unroofed. ­Kelley (2002) also used the base of the pre-Laramide fossil PAZ as a stratigraphic marker within basement rocks. Grains eroded from basement above the fossil PAZ typically have an AFT age >100 Ma, whereas grains from below the PAZ have AFT ages between 70 and 50 Ma. ­Kelley (2002) documented a predictable AFT age sequence representing unroofing of the southern ­Colorado Front Range during the Laramide Orogeny. LARAMIDE RANGES Surface samples of Precambrian basement from Laramide ranges record either AFT and AHe exhumation ages of ~80 to 40 Ma related to the Laramide Orogeny (Dickinson and Snyder, 1978), or older ages of a preLaramide fossil PAZ or PRZ. Unlike ­low-temperature thermochronologic results from the Laramide ­basins, which can be used to calibrate models of basin ­e volution and petroleum maturation, results from the ranges generally provide information on the structure and evolution of the crystalline basement uplifts, and possibly on the thermal history of sedimentary rocks beneath the range-bounding thrusts. 10711_ch03_ptg01_hr_037-070.indd 54 The spatial distribution of the base of the fossil PAZ has been used to document faulting, folding, and differential exhumation of basement. AFT ages are often younger in the cores of Laramide ranges, and become older toward the edges, indicating higher exhumation in the core and a domal structure of the range (e.g., Roberts and Burbank, 1993; Strecker, 1996; Kelley and Chapin, 1997; Kelley, 2005). Neogene AFT ages from the mountains of central Colorado and northern New Mexico are related to extension of the Rio Grande rift (Bryant and Naeser, 1980; Lindsey et al., 1986; Pazzaglia and Kelley, 1998; Naeser et al., 2002). ZFT ages reported from Laramide ranges are all Proterozoic, indicating that basement rocks have not experienced temperatures >~240°C since Proterozoic time (Cerveny, 1990; Cerveny and Steidtmann, 1993; Naeser et al., 2002). AHe ages from ranges within the Rocky Mountain foreland often demonstrate significant scatter (tens to hundreds of Ma) and are sometimes older than corresponding AFT ages (Crowley et al., 2002; Peyton et al., 2012). As discussed in Chapter 2 of this volume, radiation damage of apatite, along with He implantation from sources external to the apatite, are likely causing many of these issues. While the ­effects of radiation damage on AHe ages are predictable and provide constraints on the thermal history (Flowers et al., 2007; Flowers, 2009; Flowers et al., 2009; Peyton et al., 2012), He implantation is unpredictable and renders AHe ages geologically meaningless (Reiners et al., 2008; Orme, 2011; Peyton et al., 2012). To document AHe age scatter and allow for the potential identification of radiation damage, several apatite grains should be dated for each crystalline bedrock sample. Recognizing AHe age scatter due to ­radiation damage or He implantation may be especially difficult in sedimentary rocks because there may also be variation in AHe grain ages from different source terranes. Grain abrasion to remove the outer 20 mm of the apatite crystal (Kohn et al., 2008) is one possible way to address He implantation, but is not yet a routine and tractable part of AHe dating. Careful analysis of plots of AHe age versus effective U concentration (eU = [U] + 0.235 [U]), and AHe age versus grain size, along with comparison with AFT ages, may allow the identification of some samples that exhibit AHe age scatter due to He implantation rather than radiation damage (Flowers and Kelley, 2011; Peyton et al., 2012). Bighorn Range All published results from the Bighorn Range are presented in Figure 6. Crowley et al. (2002) studied the Bighorn Range using AHe dating of surface samples. 6/5/13 7:59 AM An Overview of Low-temperature 55 A 4000 Surface samples 3000 10 Subsurface samples 1000 20 30 0 40 -1000 BIGHORN RANGE AHe Peyton et al. (2012) AFT Peyton et al. (2012) AHe Crowley et al. (2002) AFT Cerveny (1990) -2000 -3000 50 0 100 200 300 Temperature °C Elevation (m) 2000 60 70 400 Age (Ma) B -250 250 -250 250 1250 750 Depth below pC unconformity (m) Depth below pC unconformity (m) 750 1750 2250 2750 3250 3750 1250 1750 2250 2750 3250 3750 4250 4250 4750 5250 AHe Peyton et al. (2012) AFT Peyton et al. (2012) AHe Crowley et al. (2002) AFT Cerveny (1990) 4750 5250 0 20 40 60 80 Age (Ma) 0 100 200 Age (Ma) 300 400 Profiles of AHe age versus elevation from these surface samples have low slopes, which Crowley et al. (2002) suggested represent a fossil pre-Laramide PRZ. ­Crowley et al. (2002) acknowledged that these ages are problematic because burial depth during the Cretaceous should have been great enough to reset AHe ages. They suggested that either burial was not as great 10711_ch03_ptg01_hr_037-070.indd 55 100 120 140 Figure 6. Age-elevation plot of thermochronology results from the Bighorn Range (Peyton et al., 2012). Reprinted by permission of the American Journal of Science. 50 m (164 ft). as originally thought, that geothermal gradients were unusually low, or that closure temperature of the AHe system was higher than expected based on He diffusion kinetic data available at the time. The effect of radiation damage on AHe ages and closure temperatures was not recognized until more recently (Shuster et al., 2006; Flowers et al., 2009). Later work by Peyton et al. 6/5/13 7:59 AM 56 Peyton and Carrapa (2012) using subsurface samples from the Gulf Granite Ridge #1-9-2D well (Section 9, Township 53N, Range 84W, Sheridan County, Wyoming), along with surface samples from Cloud Peak, confirmed that radiation damage of apatite had effectively increased the closure temperature of the AHe system, resulting in preservation of a fossil PRZ in Bighorn surface samples. Peyton et al. (2012) also found a low present-day geothermal gradient of 14°C/km in the borehole and suggested that gradients could have been this low in the past. AFT results from the Bighorn Range indicate that almost all surface sample ages belong to a fossil PAZ, and that insufficient exhumation has occurred to expose rocks that record Laramide AFT cooling ages (Cerveny, 1990; Peyton et al., 2012). Cerveny (1990) used track-length data to suggest that exhumation and cooling of the Bighorn Range began ~75 Ma. Peyton et al. (2012) interpreted the change in slope on an AFT age-elevation profile, which represents the onset of rapid exhumation from Cloud Peak, to be between ~99 and 57 Ma. They could not refine the onset of cooling and/or exhumation further using these AFT ages due to lack of track-length data, large errors, and few data points. Inverse modeling of AHe ages using a radiation-damage diffusion model showed that exhumation and/or cooling of the Bighorn Range started before (and possibly significantly before) ~71 Ma (Peyton et al., 2012). Peyton et al. (2012) also suggested from AHe inversion results that the Bighorn Range cooled by ~80°C during the Laramide Orogeny, corresponding to as much as ~5.7 km (~3.5 mi) of exhumation if the paleogeothermal gradient was similar to the present-day gradient of 14°C/km. This amount compared favorably with the ~5.5 km (~3.4 mi) of exhumation estimated from sequential restorations of cross sections (Hoy and Ridgway, 1997). Peyton et al. (2012) did not consider temperature changes due to climate or elevation change in their calculations. Using oxygen isotopic data from vertebrates, Amiot et al. (2004) showed that mean annual temperatures during the Late Cretaceous were similar to present-day for low latitudes (<30°N), were between ~1°C and 8°C warmer between 30°N and 60°N, and were significantly warmer (up to ~25°C) at high latitudes (>60°N). For northern Wyoming at a present-day latitude of 45°N, cooling of the mean annual temperature at sea level due to climate change from the Late Cretaceous to present was likely offset by warming due to the southward movement of the North American plate by at least 5° of latitude (Besse and Courtillot, 2002; Amiot et al., 2004). The elevation of Wyoming has changed substantially from the Late Cretaceous, when the entire state was near sea level. Present-day elevations range from 10711_ch03_ptg01_hr_037-070.indd 56 ~1 km (~0.6 mi) in the northeast corner of the state to ~4 km (~2.4 mi) at the highest peaks. Basin floor ­elevations are generally ~1.2 to 2.3 km (0.7 to 1.4 mi). Using a temperature lapse rate of 6 ± 2°C/km, a surface sample from 2 km (1.2 mi) elevation in the Bighorn Range probably cooled ~12°C due to elevation change alone. Peyton et al. (2012) estimated 80°C of cooling by inverting AHe data from a wellbore sample at 1.3 km (0.8 mi) elevation, 0.76 km (0.4 mi) below the surface. Although the surface temperature at the well location cooled ~12°C between the Late Cretaceous and present, the effects of changing surface temperature decrease with depth and as a result the cooling estimates of Peyton et al. (2012) are probably overestimated by <10°C. Reducing the amount of cooling by 10°C would reduce the estimate of exhumation by ~0.7 km (~0.4 mi), or 12%, to 5 km (3.1 mi) (using the present-day geothermal gradient of 14°C/km from Peyton et al., 2012). Beartooth Range Omar et al. (1994) calculated AFT ages of samples from the Amoco Beartooth #1 well (Section 19, Township 8S, Range 20E, Carbon County, Montana) in the northeast corner of the Beartooth Range (Figure 1), along with surface samples collected along the Beartooth Highway (U.S. Highway 212). Cerveny (1990) also dated surface samples from along the same section of highway using the AFT technique. Peyton et al. (2012) determined AHe ages from the Amoco Beartooth #1 well, but selected surface samples closer to the well ­location. All published results are shown in Figure 7. The AFT results of both Cerveny (1990) and Omar et al. (1994) show that the highest samples in the Beartooth Range (>3.1 km [>1.9 mi] elevation) are in a fossil PAZ, with AFT ages ranging from ~350 to 100 Ma. Below ~3.1 km (~1.9 mi) elevation, AFT ages range from ~68 to 48 Ma and have narrow, unimodal track length distributions with mean track lengths between 14.8 and 11.8 mm, resulting from rapid exhumation during the Laramide Orogeny (Cerveny, 1990; Omar et al., 1994). Peyton et al. (2012) inverted AHe age-eU pairs for a sample from the Amoco Beartooth #1 well using a radiation damage diffusion model and extrapolated the best-fit solution to other sample elevations using the present-day geothermal gradient. They found that the AHe age-eU distribution predicted by the best-fit thermal history matched the age-eU distributions at other sample locations that showed an age-eU correlation, confirming the influence of radiation damage on AHe ages. AHe ages from an intrusive sample were older than the zircon U/Pb crystallization age, 6/5/13 7:59 AM An Overview of Low-temperature 57 Figure 7. Age-elevation plot of thermochronology results from the Beartooth Range (Peyton et al., 2012). Inset shows same data but with an expanded time scale for more detail. All error bars are 2s. Reprinted with ­permission of the American Journal of ­Science. 1000 m (3280 ft). s­ uggesting that He implantation was affecting some samples from the Beartooth Range (Peyton et al., 2012). Using AFT ages, Omar et al. (1994) proposed that uplift of the northeast corner of the Beartooth Range began at ~61 Ma. Cerveny (1990) concluded from AFT ages that a major uplift event began in the Beartooth Range at ~68 Ma and increased in rate at ~57 Ma. Peyton et al. (2012) concluded that Laramide exhumation and cooling in the Beartooth Range started before ~58 Ma, and that the range cooled by ~98°C, corresponding to ~5 km (~3.1 mi) of exhumation (assuming a paleogeothermal gradient of 20°C/km), although again they did not consider decreases in temperature 10711_ch03_ptg01_hr_037-070.indd 57 due to climate change or increasing elevation. The surface sample used by Peyton et al. (2012) to estimate ~98°C of cooling was from an elevation of 2.1 km (1.3 mi). Therefore, ~12°C of this cooling was likely due to elevation change, reducing estimates of exhumation by ~0.5 km (~0.3 mi). Wind River Range The Wind River Range (Figure 1) has been studied using both AFT dating (Shuster, 1986; Steidtmann et al., 1989; Cerveny, 1990; Cerveny and Steidtmann, 6/5/13 7:59 AM 58 Peyton and Carrapa 1993; Peyton et al., 2012) and AHe dating (Peyton et al., 2012). Cerveny and Steidtmann (1993) dated samples from several traverses throughout the range, and interpreted a complex cooling history, with the earliest cooling initiating between ~85 and 75 Ma at the northeast and southwest flanks of the range, and rapid cooling in the core of the range between 62 and 57 Ma. AFT ages from Peyton et al. (2012) show similar cooling ages to those of Cerveny and Steidtmann (1993) in the core of the range (Gannett and Fremont Peaks). Inverse modeling of the AFT ages and track lengths from Gannett Peak showed rapid cooling and exhumation between ~60 Ma and 50 Ma (M. Fan, personal communication, 2012). AHe ages from Gannett Peak and Fremont Peak are consistently older than corresponding AFT ages by an average of 23 Ma and 14 Ma respectively, and do not increase significantly with increasing elevation as would be expected in a fossil PRZ. Peyton et al. (2012) interpreted these anomalously old AHe ages to be the result of He implantation rather than radiation damage. Steidtmann et al. (1989) used AFT dating to recognize differential exhumation of fault blocks in the core of the range. They suggested that these blocks moved relative to one another after Laramide cooling. Using sedimentological evidence they concluded that the core of the range was reactivated during late Oligocene time. Shuster (1986), Cerveny and Steidtmann (1993), and Peyton et al. (2012) dated subsurface samples from the Air Force well (Section 2, Township 32N, Range 107W, Sublette County, Wyoming) on the west side of the range (Figure 8). AFT ages from Cerveny and Steidtmann (1993) preserve a fossil PAZ above ~1100 m (~3608 ft) ­elevation, and have a steep slope below ~1100 m (~3608 ft) elevation, indicating rapid cooling at ~42 Ma. Individual grain AHe ages for the well range from 8 Ma at an elevation of –623 m (–2043 ft), to 82 Ma at an elevation of 1626 m (5334 ft), with the exception of one aliquot with an age of 163 Ma (Figure 8). Samples with multiple single-grain aliquots from the Air Force well show AHe age scatter of between ~20 and 40 Ma. At least one sample from the Air Force well showed a good correlation between AHe age and eU, indicating that radiation damage affected the AHe ages. Inverse modeling of data from this sample (elevation 1626 m [5,334 ft]) suggested that the Wind River Range cooled at least 69°C during the Laramide Orogeny, corresponding to ~3 km to 4 km (1.8-~2.4 mi) of exhumation using the present-day geothermal gradient (19°C/km from Cerveny, 1990), and that exhumation started before ~66 Ma (Peyton et al., 2012). Similar to Bighorn and Beartooth Range, Peyton et al. (2012) probably overestimate the amount of cooling due to exhumation by not considering surface-temperature effects. Ranges of Colorado, Southern Wyoming, and Northern New Mexico Figure 8. Age-elevation plot of thermochronology results from the Air Force well in the Wind River Range (Peyton et al., 2012). Reprinted by permission of the American Journal of Science. 10711_ch03_ptg01_hr_037-070.indd 58 Kelley and Chapin (1997; 2004), and Pazzaglia and Kelley (1998) measured AFT ages from the southern Front Range, Wet Mountains, the Taos Range in the Sangre De Cristo Mountains, and the Sierra Nacimiento (Figure 1). In the Front Range they recognized the base of a pre-Laramide, fossil PAZ that separates samples with AFT ages > 90 Ma and short mean track lengths (10–12 mm), from those with AFT ages of 40 to 75 Ma and longer mean track lengths (12–14 mm). Farther south, the northern parts of both the Sierra Nacimiento and the Wet Mountains preserve Laramide exhumation ages, whereas their southern parts preserve Rio Grande Rift-related Neogene exhumation ages. Pazzaglia and Kelley (1998) suggested that these age changes coincide with, and were perhaps influenced by, boundaries between Proterozoic basement provinces. AFT ages from the Taos Range 6/5/13 7:59 AM An Overview of Low-temperature 59 vary between 18 Ma and 34 Ma and reflect cooling during early Rio Grande rift deformation. Exhumation and cooling due to Rio Grande rifting was also documented along the flanks of the Blue River half graben between the Front Range and the Gore Range of central Colorado (Figure 1) (Naeser et al., 2002). AFT ages on the eastern side of the Gore Range vary from ~5 Ma at the base of the range to ~27 Ma at Mount Powell on the crest of the range. Modeling of the AFT data showed that samples cooled from temperatures > 130°C at ~30 Ma to surface temperatures (0°C to 5°C) by 5 Ma (Naeser et al., 2002). AFT ages from the western Front Range are similar to those from the western Gore Range, and modeling showed rapid cooling beginning ~25 Ma to 15 Ma (Naeser et al., 2002). Naeser et al. (2002) also dated samples from the White River Uplift (Figure 1), which gave AFT ages between 64 Ma and 34 Ma, with long mean track lengths between 13.5 mm and 14.8 mm. They interpreted these ages to result from exhumation and cooling during the Laramide Orogeny. AFT ages from the central Front Range below ~3500 m (11,482 ft) vary between ~75 Ma and 50 Ma, with long track lengths indicative of rapid cooling during the Laramide Orogeny (Bryant and Naeser, 1980; Kelley and Chapin, 1997; Naeser et al., 2002). Ages are youngest towards the center of the range. At elevations above ~3500 m (~11,482 ft), AFT ages vary between ~458 Ma and 90 Ma, and the samples, which have shortened track lengths, are interpreted to represent a fossil PAZ (Kelley and Chapin, 1997; Naeser et al., 2002). The elevation of the base of this PAZ indicates that the Front Range has a broad domal structure that has been dissected by reverse faults (Kelley and Chapin, 1997). ­Estimates of rock uplift for the central Front Range vary from 6 km (3.7 mi) at the eastern edge of the range to 7.5 km (4.6 mi) at the summit of Mount Evans; estimates for the amount of rock eroded vary from 3.5–5.2 km (2.1 to 3.2 mi) (Bryant and Naeser, 1980; Kelley and Chapin, 1997). Kelley (2005) mapped the base of the PAZ in the Laramie, Medicine Bow, and Park Ranges (Figure 1), and determined if Proterozoic basement structures were reactivated during the Laramide Orogeny. Similar cooling ages and histories on each side of the Cheyenne belt in the Laramie Range (Figure 1) indicate that this was not a zone of weakness during the Laramide Orogeny. Structural elevation of the AFT PAZ in the Laramie Range indicates a broad basement domal structure, similar to that observed in the Front Range, Black Hills, and Bighorn Range. In the Medicine Bow Mountains, however, Kelley (2005) incorporated data from Cerveny (1990) and showed that the Cheyenne belt was likely reactivated during the Laramide 10711_ch03_ptg01_hr_037-070.indd 59 Orogeny. AFT ages north of the Cheyenne belt shear zone in the Medicine Bow Mountains represent rapid Laramide cooling, whereas ages south of the shear zone are from a PAZ, indicating different amounts of exhumation on either side. In the Park Range, the Fish Creek–Soda Creek shear zone was not strongly reactivated during the Laramide Orogeny (Kelley, 2005). Cerveny (1990) found that samples from the east side of the Park Range had long (>14 mm) mean track lengths and AFT ages between 61 Ma and 60 Ma, leading him to conclude that a cooling event occurred between 61 Ma and 60 Ma. Three AFT ages and corresponding track length data from Laramie Peak in the Laramie Range led Cerveny (1990) to conclude that cooling occurred ~65 Ma. AHe ages from Laramie Peak showed scatter of hundreds of Ma (Peyton et al., 2012), with all AHe ages older than the three AFT ages for the range. The large dispersion of these anomalously old AHe ages was interpreted to be due to radiation damage of apatite (Peyton et al., 2012). Peyton et al. (2012) also dated samples from a wellbore at the northern end of the Laramie Range (Texaco Government Rocky Mountain #1 well, Section 12, Township 32N, Range 76W, Converse County, Wyoming). Using forward and inverse modeling of these AHe ages, Peyton et al. (2012) concluded that a Laramide-age thrust was present in the wellbore, and consequently that samples above and below the thrust had experienced different thermal histories. Teton Range The present-day Teton Range formed by normal faulting that began during the late Miocene (~9 Ma). Offset of Pliocene tuffs across the Teton fault shows that the hanging wall east of the Teton normal fault subsided ~5 km (~3.1 mi), and the footwall was uplifted to ~2 km (~1.2 mi) above the present surface (Roberts and Burbank, 1993, and references therein). The majority of AFT ages from three profiles across the Teton Range vary between 85 Ma and 65 Ma, indicating that the range was also exhumed during the Laramide Orogeny (Roberts and Burbank, 1993). Unlike other nearby ranges, where the majority of exhumation was due to the Laramide Orogeny, the Tetons experienced major exhumation twice, first during the Laramide Orogeny with the formation of the ancestral Teton-Gros Ventre Uplift, and later during Miocene time (Love et al., 1972; Dorr et al., 1977). AFT track lengths record this complex exhumation history, with lengths decreasing and distributions broadening with decreasing present-day elevation. Most samples currently at the 6/5/13 7:59 AM 60 Peyton and Carrapa surface resided in the PAZ after Laramide exhumation, and were exhumed to the surface during the Miocene (Roberts and Burbank, 1993). None of the ages record Miocene exhumation, indicating that all of the samples were within or above the PAZ at the end of Laramide exhumation. No ages represent the pre-Laramide fossil PAZ, which has been eroded. Samples from the northern end of the range record younger AFT ages (67–26 Ma) than those from the south (83–67 Ma), leading Roberts and Burbank (1993) to conclude that the north end of the range resided in the PAZ for a longer time than the southern end, and cooled more recently to temperatures below the PAZ. They also concluded that there was less exhumation in the southern part of the range relative to the north, resulting in a tilting of the Precambrian-Cambrian unconformity ~3° to the south. Black Hills The Blacks Hills of South Dakota are the easternmost basement-cored range formed during the Laramide Orogeny (Figure 1). Unlike other ranges discussed here, no range-bounding reverse fault is mapped at the Black Hills. Paleozoic and Mesozoic rocks dip concentrically away from the Precambrian core of the range. Both Cerveny (1990) and Strecker (1996) determined AFT ages of surface samples from the Black Hills, but to date, no AHe studies have been reported. Strecker (1996) dated 19 samples from Precambrian granite and metasedimentary rocks across the range. AFT ages ranged from 53.5 Ma ± 2.4 Ma to 113.5 Ma ± 11.4 Ma. Multiple populations of track lengths indicated that none of the samples were totally annealed before cooling. Strecker (1996) did not measure apatite composition but did recognize that all granitic samples had lower AFT ages than metasedimentary samples, and suggested that this was due to compositional differences resulting in different annealing temperatures. Spatially, the AFT ages define a fold, with older ages at the edges of the uplift. An age-elevation plot of the widely spaced samples showed no indication of AFT age increasing with present-day elevation, with older ages at lower elevations near the edge of the range. However, by flattening the dip of the fold limb about the hinge of the anticline and projecting the ages to the new corrected elevations, the age-elevation plot shows a much clearer trend of increasing age with increasing sample elevation (Strecker, 1996). Strecker ’s (1996) method for correcting elevations is similar to that used by Crowley et al. (2002) in the Bighorn Range, who calculated elevation relative to the Precambrian-Cambrian unconformity. Interpretation 10711_ch03_ptg01_hr_037-070.indd 60 of this age versus corrected elevation plot led Strecker (1996) to conclude that exhumation and cooling began in the Black Hills in earliest Paleocene time (~66 Ma)and continued until ~52 Ma, with an increase in cooling rate at ~58 Ma. Estimates of exhumation ranged from 3.8 km (2.3 mi) ± 1.2 km (0.7 mi) to 5.0 km (3.1 mi) ± 1.5 km (0.9 mi), with denudation rates most likely ~0.2 km/Ma to 0.4 km/Ma. Cerveny (1990) dated three samples from Precambrian granite in the Black Hills. The oldest age of 262 Ma ± 28 Ma was found on the outer edge of the batholith. Ages of 79.4 Ma ± 9 Ma and 71.4 Ma ± 5.7 Ma were found structurally lower in the batholith. Similar to Strecker’s (1996) data, the three ages of Cerveny (1990) increase with structural elevation but not with actual elevation. A bimodal track length distribution and short mean track length (12.1 mm ± 2.4 mm) indicate that the oldest age was from a fossil PAZ. Mean track length distributions close to 14 mm for the younger two samples were interpreted to indicate that exhumation and uplift occurred at ~75 Ma (Cerveny, 1990). It is interesting that Cerveny (1990) concludes that uplift was occurring at ~75 Ma, whereas Strecker (1996) concludes that cooling began ~66 Ma in the same area. With only three data points it is possible that Cerveny (1990) misinterprets his two youngest ages as indicating rapid cooling, whereas with more data it might have been apparent that they actually belong to a fossil PAZ, although the long mean track lengths suggest otherwise. Uinta Mountains One sample collected from the Uinta Mountain Group in the eastern Uinta Mountains provided an AFT age of 52.28 Ma ± 4.97 Ma, but had insufficient tracks to provide a reliable mean track length (C. Painter, personal communication, 2012). Unfortunately, due to the quartzitic nature of the Uinta Mountain Group, additional surface samples did not yield any apatite grains. Two samples collected from the McMoran-Freeport 43-2 Middle Mountain well (Section 2, Township 2N, Range 25E, Daggett County, Utah) at depths of 1372 m (4500 ft) and 2286 m (7500 ft) yielded AFT ages of 38.9 Ma ± 2.6 Ma and 30.3 Ma ± 2.2 Ma respectively (K. Constenius and S. Kelley, personal communication, 2012). Each of these samples was a composite of well cuttings over a depth range of ~150 m (500 ft) from metamorphic rocks of the Archean Owiyukuts Complex. Sample depths represent the central depth of the composite depth range. Track lengths were not available. Two grains with large etch pits were noted 6/5/13 7:59 AM An Overview of Low-temperature 61 for the sample from 1372 m (4501 ft) (K. Constenius and S. Kelley, personal communication, 2012). Bottomhole temperatures from well logs were corrected using the methods of Waples and Ramly (2001), Waples et al. (2004a), and Waples et al. (2004b) and extrapolated to sample depths. Results indicate that present-day temperatures of these samples are ~45°C ± 10°C at 1372 m depth, and ~60°C ± 10°C at 2286 m (7500 ft). These temperatures suggest that the shallowest sample was cooler than the present-day AFT PAZ (~60–120°C for fluorapatite, Green et al., 1989), and may represent cooling during the late Eocene (~39 Ma); however, without track-length and kinetic data it is difficult to evaluate if fission tracks have been partially annealed in the deeper sample. Summary of Laramide Cooling Ages To summarize the results reviewed previously, we have mapped Laramide cooling and exhumation ages from ranges across the entire region (Figure 9). When available, the Laramide cooling ages determined by inverse modeling of either AFT or AHe data are shown. Figure 9. Map showing 112 108 CMB 104 57 48 Ma (AFT) SOUTH DAKOTA 80 40 Ma (AHe model) 57 54 Ma (AFT surface) (Omar et al., 1994; 85 Ma (AHe model) 100 Peyton et al., 2012) (Peyton et al., 2012) BT MONTANA MR WYOMING LP 77 60 Ma 83 67* Ma 62 Ma TR (Roberts & Burbank, 1993) (Cerveny, 1990; Strecker, 1996) BH (Cerveny, 1990) (Peyton et al., 2012) (Cerveny, 1990) WR UTAH thru st 39 30* Ma LR 79 60 Ma NEBRASKA 71 Ma (Cerveny, 1990) MB 76 46 Ma COLORADO (Kelley, 2005 Cerveny, 1990)PR 52* Ma 70 45 Ma (AFT model) (Kelley & Chapin, 2004) (Painter, p.c. 2012) AHe surface samples (bold symbol outlines indicate results from inverse modeling) AHe subsurface samples DB WRU GR PB 40 UB (Kelley, 2005) (Kelley, 2005 Cerveny, 1990) (Constenius & Kelley, p.c. 2012) UM 40 HU 82 65 Ma GRB Sevier ? GM front IDAHO 70 60 Ma (AHe forward model) 71 Ma CB (Peyton et al., 2012 Cerveny & Steidtmann, 1993 Fan, p.c. 2012) WRB BL PRB OC 60 50 Ma (AFT surface model) 39 44 Ma (AFT subsurface) 70 60 Ma (AHe model) 44 44 BB FR 75 50 Ma (Kelley & Chapin, 2004 Kelley & Chapin, 1997) 63 34 Ma (Naeser et al., 2002) SR AFT ages from surface samples (track lengths >12 mm) * no or variable 112track lengths AFT subsurface samples UU WM Precambrian crystalline basement Youngest Laramide cooling ages: 80 Ma + 60 50 Ma 80 70 Ma 50 40 Ma 70 60 Ma < 40 Ma SC TA Younger ages related to Rio Grande rifting COLORADO PLATEAU AHe ages related to He implantation or PRZ 0 112 10711_ch03_ptg01_hr_037-070.indd 61 100 N SN 52 33 Ma 200 km ARIZONA NEW MEXICO (Pazzaglia & Kelley, 2004) 108 104 AFT and AHe cooling ages from multiple studies of Laramide ranges. Colors represent youngest cooling ages. Thick dashed line is outline of the Colorado Plateau. Thin dashed line represents the Cheyenne Belt. See text for description of how dates were selected. Abbreviations as follows: BB, Bighorn Basin; BL, Black Hills; BH, Bighorn Range; BT, Beartooth Range; CB, Cheyenne Belt; CMB, Crazy Mountains Basin; DB, Denver Basin; FR, Front Range; GM, Granite Mountains; GR, Gore Range; GRB, Green River Basin; HU, Hartville Uplift; LP, Lima Peaks; LR, Laramie Range; MB, Medicine Bow Mountains; MR, Madison Range; OC, Owl Creek Mountains; PB, Piceance Basin; PR, Park Range; PRB, Powder River Basin; SC, Sangre de Cristo Range; SN, Sierra Nacimiento; SR, Sawatch Range; TA, Taos Range; TR, Teton Range; UB, Uinta Basin; UM, Uinta Mountains; UU, Uncompahgre Uplift; WM, Wet Mountains; WR, Wind River Range; WRB, Wind River Basin; WRU, White River Uplift; p.c., personal communication. 100 km (62 mi). 6/5/13 7:59 AM 62 Peyton and Carrapa If no inversion results are available, we include cooling ages interpreted from age-elevation profiles. If they are available and significant, we present results for multiple age-elevation profiles in a range. For example, age-elevation profiles in the Wind River Range are available from high-elevation samples (e.g., Gannett Peak and Fremont Peak) and subsurface samples (e.g., Air Force well). Figure 9 shows both the older cooling ages from higher in the range, and younger cooling ages from the subsurface. We exclude AFT ages that have mean track lengths <12 mm because they are not indicative of rapid cooling. In areas with no age-elevation profiles, we use cooling ages interpreted from individual AFT ages and track length distributions (e.g., for the Owl Creek and Granite mountains we used only single AFT ages with track lengths >12 mm). We have also included a new, as yet unpublished AFT surface age from the Uinta Mountains (C. Painter, personal communication, 2012), as well as subsurface AFT ages from the McMoran-Freeport 43-2A Middle Mountain well at the eastern end of the Uinta Mountains (K. Constenius and S. Kelley, personal communication, 2012). We do not include cooling ages from the Laramide basins, or ages that are interpreted to be due to cooling from events unrelated to the Laramide Orogeny (e.g., Miocene uplift due to the Rio Grande Rift or basin-and-range extension at the Teton Range). Symbol colors in Figure 9 represent the youngest recorded cooling age. Examination of Figure 9 shows an apparent younging of both oldest and youngest cooling ages to the south and southwest, with the youngest ages from the Uinta Mountains, White River Uplift, and Sierra Nacimiento, and the oldest ages from the Bighorn, Beartooth, and Teton Ranges. It is interesting to note that the youngest ages (<40 Ma) are from samples that are near the edge of the Colorado Plateau (Figure 9). These youngest ages generally agree with a study of basin-fill sediments by Dickinson et al. (1988), who suggested that Laramide deformation terminated at ~50 Ma in the northern foreland and at ~35 Ma in the southern foreland. Care should be taken when interpreting such a diverse data set. The absence of old ages may indicate either that an area has been exhumed and older ages eroded, or that cooling started at a later time than in areas with older ages. The presence of a fossil PAZ or PRZ in an age-elevation profile can help with this distinction. If a fossil PAZ or PRZ is absent, then erosion of older ages is a possibility. Similarly, the depth and present-day temperature of a subsurface sample should also be considered: Young ages from a wellbore sample may be due to the sample residing within the present-day PRZ or PAZ (i.e., the age is partially reset), and not due to its cooling history. 10711_ch03_ptg01_hr_037-070.indd 62 APPLICATION TO MATURATION STUDIES Combining basin modeling (burial-history or geohistory modeling) with low-temperature thermochronology and %Ro maturity data can provide information on if and when source rocks reached high enough temperatures for a sufficient amount of time to generate oil and gas. Low-temperature thermochronometers may also reveal the timing of tectonically driven exhumation and fluid flow by recording associated cooling. The success of exploration for conventional petroleum plays is critically dependent upon the timing of both petroleum generation and trap formation. In unconventional reservoirs, the shale may be the source, seal, and reservoir, but fracturing, and hence timing of tectonics, is often critical for successful production. Workflows and examples of reconstructing the thermal history of a sedimentary basin by integrating basin modeling, low-temperature thermochronology, %R o, and sometimes fluid inclusion data have been discussed by Duddy et al. (1994), Duddy et al. (1998), Crowhurst et al. (2002), Green et al. (2004), and Parnell et al. (2005). Burial-history models are typically created for samples from a wellbore, using stratigraphic thicknesses and depositional ages, porosity-depth relationships, lithology and mineralogy (to determine thermal conductivity), the present-day geothermal gradient calculated from present-day corrected borehole temperatures, and estimates of paleosurface temperatures and heat-flow history. In many studies that integrate basin modeling, lowtemperature thermochronology, and %Ro data, an initial estimate of the burial history is first determined for a sample by assuming no cooling or exhumation, a constant geothermal gradient, and constant ­basal heat flow. This initial estimate is often referred to as the “default thermal history” (e.g., Duddy et al., 1994; Green et al., 2004; Parnell et al., 2005). %R o values can then be calculated for this default thermal history and compared to actual %Ro values for the sample. If predicted and calculated %R o values match, then the sample is currently at its maximum temperature and has not experienced any cooling. However, if reliable measured %Ro values are higher than those predicted by the default thermal model, then the sample experienced higher paleotemperatures in the past than its present-day temperature. Similarly, low-temperature-thermochronology ages can be calculated from the default thermal history and compared to actual values. If measured ages are younger than predicted ages, then the sample was likely at higher temperatures in the past than predicted by the default model. 6/5/13 7:59 AM An Overview of Low-temperature 63 Estimated erosion Present-day surface Depth l uria al b ition Add Aquifer with hot fluids d se ea cr tion he at w flo flow flow fluid fluid tion ura nt ura In gd Lon rt d adie Sho y gr t-da sen 10711_ch03_ptg01_hr_037-070.indd 63 Paleo-surface temp. Pre The maximum paleotemperature experienced by a sample can be estimated from %Ro data using the ­kinetics of Burnham and Sweeney (1989) and Sweeney and Burnham (1990). A range of possible maximum paleotemperatures can also be estimated from AFT age- and track-length data using an inversion program such as HeFTy, which calculates a best-fit thermal history and associated confidence limits (Ketcham, 2005; Ketcham et al., 2007). If individual grain AHe data for a sample show a correlation between grain age and grain size (Reiners and Farley, 2001), or grain age and eU content (i.e., radiation damage, Flowers et al., 2009), then AHe data can also be inverted to find the best-fit thermal history (Flowers et al., 2007; Flowers, 2009). Plotting the maximum paleotemperature against depth in a wellbore gives an estimate of paleogeothermal gradient (Bray et al., 1992). If the paleogeothermal gradient is parallel to the present-day geothermal gradient, but offset to higher temperatures, then samples were likely buried deeper in the past but geothermal gradients and heat flow were similar to present day (Figure 10). On the other hand, if the paleogeothermal gradient is higher than present day, then heat flow was higher in the past (Bray et al., 1992). Extrapolating these paleogeothermal gradients to an assumed paleosurface temperature provides an estimate of the amount of material eroded from the basin (Figure 10). Nonlinear paleogeothermal gradients can result from large changes in thermal conductivity (e.g., from a low-conductivity shaly section to a higher-conductivity carbonate-evaporite sequence), but in many cases indicate the passage of hot fluids. Although the movement of high-temperature fluids necessary to create a noticeable change in geothermal gradient will be fundamentally vertical, the fluids may also have a lateral component to their flow, and thus may not affect all parts of a sedimentary section equally. Duddy et al. (1994) suggested that hot fluid flow in a formation would result in a characteristic shape of the paleotemperature-depth profile, with higher paleogeothermal gradients above the formation that experienced the hot fluid flow than below it. Using the maximum paleotemperatures and paleogeothermal gradient, the default thermal history is adjusted to create a basin thermal history that more closely matches the %R o and thermochronometer data. Similarly, the time-temperature histories from inversion of thermochronometer data can be used iteratively to refine the basin model. Using AFT and %Ro data, Green et al. (2004) followed the above methodology to determine a thermal history with two episodes of cooling from peak paleotemperatures for a well from the Otway Basin in Paleotemperature (oC) Figure 10. Paleotemperature plotted versus depth. Dotted black line is present-day geothermal gradient. Solid line is paleogeothermal gradient if sample was buried deeper in the past but the paleogeothermal gradient and heat flow were similar to present day. Gray solid line is the projection of this paleogeothermal gradient to the assumed paleosurface temperature, which gives an estimate of the amount of material eroded. Long dashed line is paleogeothermal gradient if heat flow was higher in the past. Solid red line is paleogeothermal gradient after a long duration of fluid flow. Dashed red line is paleogeothermal gradient after a short duration of fluid flow. Modified from Duddy et al. (1994) and Duddy et al. (1998). 6/5/13 7:59 AM 64 Peyton and Carrapa southeastern Australia. With insufficient data to identify a variation of AHe age with grain size, they used forward modeling of AHe data to refine the timing of the youngest cooling episode. Parnell et al. (2005) integrated burial history modeling, AFT, %R o and fluid ­inclusion data to recognize a hot fluid event in Cenozoic sandstones on the UK Atlantic margin that was confined to fractures and did not reset AFT ages. ­Crowhurst et al. (2002) refined basin thermal-history models for the Taranaki Basin of New Zealand using AHe and %R o data using a paleotemperature ­approach, and then forward modeled AHe ages for four possible Neogene cooling scenarios. A model with two discrete episodes of cooling between 8.5 and 2 Ma provided the best fit to the measured AHe ages. In that case, AHe dating provided additional control on the nature of the cooling history due to its sensitivity to lower temperatures than AFT dating. The relative ­timing of maximum burial, and hence ­hydrocarbon generation, and the formation of trapping structures are critical to predicting which structures may contain petroleum in the Taranaki Basin (Armstrong et al., 1996; Crowhurst et al., 2002). Osadetz et al. (2002) used burial history modeling and AFT data to interpret the thermal history of the Canadian Williston Basin. They concluded that although maximum burial of basement was attained during late Cretaceous–early Tertiary time, maximum paleotemperatures were attained during the late Paleozoic. Burial depths of <2 km (<1.2 mi) in the late Paleozoic imply that heat flow was significantly elevated, and that lower Paleozoic source rocks may have generated oil at that time. Armstrong (2005) reviewed and summarized this paper and its results, so it will not be discussed further here. As mentioned earlier, there have been no published studies from the Laramide foreland (Figure 1) that used low-temperature thermochronology to determine source-rock maturation and basin history. Dickinson (1986) incorporated the conclusions of Naeser (1986) into his basin model, but did not iteratively adjust his model to match AFT ages. Two studies have been published from the adjacent Sevier foldthrust belt, one from the Idaho-Wyoming part of the thrust belt (Burtner and Nigrini, 1994), the other from the Canadian fold-thrust belt in southern Alberta (Osadetz et al., 2004). Burtner and Nigrini (1994) determined AFT ages for surface and subsurface samples from the IdahoWyoming part of the Sevier fold-thrust belt. Almost all AFT ages were significantly younger than depositional ages and were interpreted as cooling ages. They used forward and inverse modeling of AFT ages and tracks to calculate time-temperature 10711_ch03_ptg01_hr_037-070.indd 64 histories for two samples. After computing a burial history using BasinMod (Platte River Associates, http://www.platte.com), basal heat flow was varied until the time-temperature histories matched those calculated from the AFT data. After varying the amount of Cretaceous strata eroded and the heat flow model, Burtner and Nigrini (1994) could not match deep AFT and %Ro data. They concluded that movement on the Crawford thrust ~80–90 Ma was coeval with a major decrease in the geothermal gradient, and suggested that this decrease was due to gravity-driven fluid flow. Increased heat flow ~110– 90 Ma in Lower Cretaceous Gannett Group rocks was attributed to movement of hot fluids prior to movement on the Crawford thrust. Osadetz et al. (2004) sampled rocks for AFT dating from both the hanging wall and footwall of the Late Cretaceous–Paleocene Lewis thrust, and the hanging wall and footwall of the Eocene–Oligocene Flathead normal fault in southern Alberta. Using only samples with a sufficient number of measured tracks, they modeled thermal histories using MonteTrax (Gallagher, 1995). Major cooling events from AFT modeling could be related to times of major displacement on the Lewis and Flathead faults. Thermal-history models from samples at high levels of the Lewis thrust showed one cooling event that began ~75 Ma, coincident with displacement on the Lewis thrust. Samples from the footwall of the Flathead normal fault (both hanging wall and footwall of the Lewis thrust) also recorded a cooling event that initiated in the Eocene due to extension on the Flathead fault. AFT ages from Oligocene sediments on the hanging wall of the Flathead fault represent cooling ages of the sediment source terrains, and were not reset by burial. By integrating AFT modeling results with %Ro data, Osadetz et al. (2004) concluded that the geothermal gradient during Lewis thrusting (~8.6°C/km to 12°C/km) was lower than present day (~17°C/km), which they attributed to advective cooling by topographically driven groundwater flow. They also concluded that petroleum generation below the Lewis thrust was either due to tectonic burial by the overriding Lewis thrust, or deep burial by rapidly deposited sedimentary rocks that are no longer preserved. CONCLUSION Results from AFT and AHe dating of samples from the Rocky Mountain foreland show that to study the timing of the Laramide Orogeny, the best places to collect samples are either from the ranges, where AFT and 6/5/13 7:59 AM An Overview of Low-temperature 65 AHe ages reflect either Laramide cooling or an older PRZ or PAZ, or the upper kilometer of sedimentary rock from wellbores in the basins, which represent a detrital unroofing sequence if thermochronometers have not been reset due to burial. Detrital thermochronology from preserved surface sedimentary rocks may also provide information on older exhumation events that are no longer preserved in the ranges today (e.g., Pennsylvanian sedimentary rocks in the Laramide Rocky Mountains may record cooling ages associated with the Ancestral Rockies Orogeny, even though the present-day ranges record Laramide cooling). This application has not been exploited yet in the western United States. The best way to investigate Neogene exhumation of the Rocky Mountains is using samples from deep boreholes in the basins where ages were reset due to burial during and after the Laramide Orogeny. Reviewing the results of more than 30 years of low-temperature thermochronology in the Rocky Mountains, a regional temporal evolution of the Laramide Orogeny is elusive, but new AFT data (C. Painter, K. Constenius and S. Kelley, personal communication, 2012) combined with existing ages seem to indicate a younging of cooling ages to the south and southwest. Where multiple elevation transects have been dated from a single range, it is apparent that cooling and exhumation histories within ranges are complex (Cerveny, 1990; Cerveny and Steidtmann, 1993; Roberts and Burbank, 1993; Kelley and Chapin, 2004; Kelley, 2005; Peyton et al., 2012). For example, it is perhaps unrealistic to conclude that the Wind River Range experienced rapid cooling at 42 Ma (Cerveny, 1990; Cerveny and Steidtmann, 1993), and better to conclude (as Cerveny and Steidtmann, 1993, did) that the west side of the range in the vicinity of the Air Force well experienced rapid cooling at that time, and that this scenario may not apply to the entire range. Studies from the Beartooth Range (Cerveny, 1990; Omar et al., 1994; Peyton et al., 2012), Wind River Range (Cerveny, 1990; Cerveny and Steidtmann, 1993; Peyton et al., 2012), Park Range (Cerveny, 1990) and Black Hills (Strecker, 1996) all concluded that (possibly rapid) cooling occurred ~62 Ma to 57 Ma, suggesting a regionwide cause. The earliest cooling documented to date in the Laramide foreland is ~85 to 75 Ma in the Wind River Range (Cerveny and Steidtmann, 1993), Bighorn Range (Cerveny, 1990; Peyton et al., 2012), and Teton Range (Roberts and Burbank, 1993). Significant scatter of AHe ages, likely due to radiation damage and He implantation (Peyton et al., 2012), along with often-large error bars on many AFT ages (± ~5– 10 Ma 1s) and variation in AFT ages between different 10711_ch03_ptg01_hr_037-070.indd 65 workers (e.g., Cerveny, 1990; Strecker, 1996), makes resolving detailed timing of exhumation of one range versus another difficult, and leaves room for more work in the region. Our understanding of the diffusion of He in apatite is still developing, particularly in rocks that have spent a long time (relative to their age) in the PAZ (e.g., the Precambrian crystalline basement exposed in the Laramide ranges). Although well established in other areas, AHe dating is still somewhat of a nascent technique in the Rocky Mountain region. Recognition of the influence of radiation damage on AHe ages from cratonic regions such as the Rocky Mountains has advanced the understanding and utility of these ages significantly in recent years (Flowers et al., 2007; Flowers, 2009; Flowers and Kelley, 2011; Peyton et al., 2012). The development of a tractable way to remove the effects of He implantation will expand their usefulness further. Low-temperature thermochronometers such as AFT and AHe dating have many potential uses in understanding the structural and thermal evolution of petroleum-producing regions. Age-elevation profiles from ranges adjacent to sedimentary basins can provide the timing and rate of exhumation of the source area of the sedimentary rocks. If the cooling and exhumation are directly related to tectonics, then they also constrain the timing of deformation, which may have implications for hydrocarbon migration and trapping. Age-elevation profiles from wells within a sedimentary basin provide information on the timing and amount of burial and exhumation of the basin. Dating a suite of samples from as large a depth range as possible in a single wellbore reduces overall uncertainty, since all samples must conform to a single basin thermal-history model. If shallow sedimentary rocks have not been thermally reset, their thermochronometer ages are detrital cooling ages and may help constrain the provenance and maximum possible depositional age of the rock, as well as the unroofing history of the source areas. Thermochronometer ages can provide specific information on timing and amount of cooling, and can thus be used to constrain basin thermal-history models. Integrating a proposed burial-history model derived from stratigraphic and sedimentary data with (1) thermochronometer data, (2) constraints on total heating from %Ro, and (3) constraints on temperatures at specific times in the past from fluid inclusions, allows one to arrive, iteratively, at a highly constrained thermalhistory model. This new model can in turn improve our understanding of the level and timing of sourcerock maturation, and thus improve our exploration success. 6/5/13 7:59 AM 66 Peyton and Carrapa Low-temperature thermochronology is an underutilized petroleum exploration tool in the Rocky Mountains. We have illustrated that it has been used successfully throughout the world to constrain sourcerock maturation history and burial history of sedimentary basins, as well as tectonic history and sediment provenance. With the current high interest in unconventional shale reservoirs throughout the Rocky Mountain region, we recommend that low-temperature thermochronology be added to the explorationist’s tool box. ACKNOWLEDGeMENTS We thank Shari Kelley, Jim Steidtmann, and Doug Waples for their constructive reviews. Connie Knight and Rich Bottjer provided many helpful suggestions for improving the manuscript. Majie Fan lent her expertise in paleoaltimetry and paleotemperatures. Clay Painter generously shared his AFT age from the Uinta Mountains. We thank Kurt Constenius and Shari Kelley for sharing their data from the McMoranFreeport 43-2 Middle Mountain well. Doug Waples also provided assistance with well-log temperature corrections. We are grateful to Cirque Resources for providing a quiet place to write. REFERENCES CITED Amiot, R., C. Lécuyer, E. Buffetaut, F. Fluteau, S. Legendre, and F. Martineau, 2004, Latitudinal temperature gradient during the Cretaceous Upper Campanian–Middle Maastrichtian: d18O record of continental vertebrates: Earth and Planetary Science Letters, v. 226, p. 255–272. Armstrong, P. A., D. S. Chapman, R. H. Funnell, R. G. Allis, and P. J. J. Kamp, 1996, Thermal modeling and hydrocarbon generation in an active-margin basin: Taranaki Basin, New Zealand: AAPG Bulletin, v. 80, p. 1216–1241. Armstrong, P. A., 2005, Thermochronometers in sedimentary basins: Reviews in Mineralogy and Geochemistry, v. 58, p. 499–525. Beland, P. E., 2002, Apatite fission track and (U-Th)/He thermochronology and vitrinite reflectance of the Casper Arch, Maverick Springs Dome, and the Wind River Basin: Implications for Late Cenozoic deformation and cooling in the Wyoming foreland, M.S. thesis, University of Wyoming, Laramie, p. 97. Besse, J., and V. Courtillot, 2002, Apparent and true polar wander and the geometry of the geomagnetic field over the last 200 Myr: Journal of Geophysical Research, v. 107, p. 2300–2331. Bostic, N. H., 1983, Vitrinite reflectance and temperature ­gradient models applied at a site in the Piceance Basin, Colorado: AAPG Bulletin, v. 67, p. 427–428. 10711_ch03_ptg01_hr_037-070.indd 66 Bostic, N. H., and V. L. Freeman, 1984, Tests of vitrinite reflectance and paleotemperature models at the Multi-well experiment site, Piceance Creek Basin, Colorado: USGS Open-File Report OF 84-757, p. 110–120. Bray, R. J., P. F. Green, and I. R. Duddy, 1992, Thermal history reconstruction using apatite fission track analysis and vitrinite reflectance: A case study from the UK East Midlands and southern North Sea, in R. F. P. Hardman, ed., Exploration Britain: Geological insights for the next decade. GSL Special Publication, v. 67, p. 3–25. Brown, W. G., 1988, Deformational style of Laramide uplifts in the Wyoming foreland: GSA Memoir 171, p. 1–25. Bryant, B., and C. W. Naeser, 1980, The significance of ­fission-track ages of apatite in relation to the tectonic history of the Front and Sawatch Ranges, Colorado: GSA Bulletin, v. 91, p. 156–164. Burnham, A. K., and J. J. Sweeney, 1989, A chemical kinetic model of vitrinite maturation and reflectance: ­Geochimica et Cosmochimica Acta, v. 53, p. 2649–2657. Burtner, R. L., and A. Nigrini, 1994, Thermochronology of the Idaho–Wyoming thrust belt during the Sevier Orogeny: A new, calibrated, multiprocess model: AAPG Bulletin, v. 78, p. 1586–1612. Cecil, M. R., M. N. Ducea, P. W. Reiners, and C. G. Chase, 2006, Cenozoic exhumation of the northern Sierra Nevada, California, from (U-Th)/He thermochronology: GSA Bulletin, v. 118, p. 1481–1488. Cerveny, P. F., 1990, Fission-track thermochronology of the Wind River Range and other basement cored uplifts in the Rocky Mountain foreland, Ph.D. thesis, University of Wyoming, Laramie, p. 180. Cerveny, P. F., and J. R. Steidtmann, 1993, Fission track thermochronology of the Wind River Range, Wyoming: Evidence for timing and magnitude of Laramide exhumation: Tectonics, v. 12, p. 77–91. Chen, Z., and S. Lubin, 1997, A fission-track study of terrigenous sedimentary sequences of the Morrison and Cloverly formations in the northeastern Bighorn Basin, Wyoming: The Mountain Geologist, v. 34, p. 51–62. Coney, P. J., and S. J. Reynolds, 1977, Cordilleran Benioff zones: Nature, v. 270, p. 403–406. Constenius, K. N., 1996, Late Paleogene extensional collapse of the Cordilleran foreland fold and thrust belt: GSA ­Bulletin, v. 108, p. 20–39. Coogan, J. C., 1992, Thrust systems and displacement ­transfer in the Wyoming–Idaho–Utah thrust belt: Ph.D. dissertation thesis, University of Wyoming, Laramie, 496 p. Crowhurst, P. V., P. F. Green, and P. J. J. Kamp, 2002, Appraisal of (U-Th)/He apatite thermochronology as a thermal history tool for hydrocarbon exploration: An example from the Taranaki Basin, New Zealand: AAPG Bulletin, v. 86, p. 1801–1819. Crowley, P. D., P. W. Reiners, J. M. Reuter, and G. D. Kaye, 2002, Laramide exhumation of the Bighorn Mountains, Wyoming: An apatite (U-Th)/He thermochronology study: Geology, v. 30, p. 27–30. 6/5/13 7:59 AM An Overview of Low-temperature 67 DeCelles, P. G., et al., 1987, Laramide thrust-generated alluvial-fan sedimentation, Sphinx Conglomerate, southwestern Montana: AAPG Bulletin, v. 71, p. 135–155. DeCelles, P. G., M. B. Gray, K. D. Ridgway, R. B. Cole, P. Srivastava, N. Pequera, and D. A. Pivnik, 1991, Kinematic history of a foreland uplift from Paleocene synorogenic conglomerate, Beartooth Range, Wyoming and Montana: GSA Bulletin, v. 103, p. 1458–1475. DeCelles, P. G., 2004, Late Jurassic to Eocene evolution of the Cordilleran thrust belt and foreland basin system, western U.S.A.: American Journal of Science, v. 304, p. 105–168. Dickinson, W. R., and W. S. Snyder, 1978, Plate tectonics of the Laramide Orogeny, in V. Matthews, ed., Laramide folding associated with basement block faulting in the western United States: GSA Memoir 151, p. 355–366. Dickinson, W. R., M. A. Klute, M. J. Hayes, S. U. Janecke, E. R. Lundin, M. A. McKittrick, and M. D. Olivares, 1988, Paleogeographic and paleotectonic setting of Laramide sedimentary basins in the central Rocky Mountain ­region: GSA Bulletin, v. 100, p. 1023–1039. Dickinson, W. W., 1986, Analysis of vitrinite maturation and Tertiary burial history, northern Green River Basin, Wyoming: USGS Bulletin 1886, p. F1–F13. Donelick, R. A., 1993, A method of fission track analysis utilizing bulk chemical etching of apatite, Patent 5267274, U.S.A. Dorr, J. A., Jr., D. R. Spearing, and J. R. Steidtmann, 1977, Deformation and deposition between a foreland uplift and an impinging thrust belt: Hoback Basin, Wyoming: GSA Special Paper 177, 82 p. Duddy, I. R., P. F. Green, K. A. Hegarty, and R. J. Bray, 1991, Reconstruction of thermal history in basin modelling using apatite fission track analysis: What is really possible? Offshore Australia Conference Proceedings, v. 1, p. III-49–III-61. Duddy, I. R., P. F. Green, R. J. Bray, and K. A. Hegarty, 1994, Recognition of the thermal effects of fluid flow in sedimentary basins, in J. Parnell, ed., Geofluids: Origin, migration and evolution of fluids in sedimentary basins: GS Special Publication No. 78, p. 325–345. Duddy, I. R., P. F. Green, K. A. Hegarty, R. J. Bray, and G. W. O’Brien, 1998, Dating and duration of hot fluid flow events determined using AFTA and vitrinite reflectancebased thermal history reconstruction, in J. Parnell, ed., Dating and duration of fluid flow and fluid rock interaction: GS Special Publication 144, v. 41–51. Erslev, E. A., 1993, Thrusts, back-thrusts and detachment of Laramide foreland arches, in C. J. Schmidt, R. B. Chase, and E. A. Erslev, eds., Laramide basement deformation in the Rocky Mountain foreland of the western United States: GSA Special Paper 280, p. 125–146. Flowers, R. M., D. L. Shuster, B. P. Wernicke, and K. A. Farley, 2007, Radiation damage control on apatite (U-Th)/ He dates from the Grand Canyon region, Colorado ­Plateau: Geology, v. 35, p. 447–450. Flowers, R. M., B. P. Wernicke, and K. A. Farley, 2008, Unroofing, incision, and uplift history of the southwestern 10711_ch03_ptg01_hr_037-070.indd 67 Colorado Plateau from apatite (U-Th)/He thermochronometry: GSA Bulletin, v. 120, p. 571–587. Flowers, R. M., 2009, Exploiting radiation damage control on apatite (U-Th)/He dates in cratonic regions: Earth and Planetary Science Letters, v. 277, p. 148–155. Flowers, R. M., R. A. Ketcham, D. L. Shuster, and K. A. ­F arley, 2009, Apatite (U-Th)/He thermochronometry using a ­r adiation damage accumulation and annealing model: Geochimica et Cosmochimica Acta, v. 73, p. 2347–2365. Flowers, R. M., and S. A. Kelley, 2011, Interpreting data dispersion and “inverted” dates in apatite (U-Th)/He and fission-track datasets: An example from the U.S. midcontinent: Geochimica et Cosmochimica Acta, v. 75, p. 5169–5186. Galbraith, R. F., 2005, Statistics in fission track analysis: Interdisciplinary Statistics: Boca Raton, Florida, Chapman and Hall/CRC, 219 p. Gallagher, K., 1995, Evolving temperature histories from apatite fission-track data: Earth and Planetary Science Letters, v. 136, p. 421–435. Giegengack, R., G. I. Omar, and K. R. Johnson, 1986, A reconnaissance fission track uplift chronology for the northwest margin of the Bighorn Basin: Geology of the Beartooth Uplift and adjacent basins, Yellowstone Bighorn Research Association, 179–184 p. Green, P. F., I. R. Duddy, A. J. W. Gleadow, P. R. Tingate, and G. M. Laslett, 1986, Thermal annealing of fission tracks in apatite: 1. A qualitative description: Chemical Geology (Isotope Geoscience Section), v. 59, p. 237–253. Green, P. F., I. R. Duddy, G. M. Laslett, K. A. Hegarty, A. J. W. Gleadow, and J. F. Lovering, 1989, Thermal annealing of fission tracks in apatite: 4. Quantitative modelling techniques and extension to geological timescales: Chemical Geology, v. 79, p. 155–182. Green, P. F., P. V. Crowhurst, and I. R. Duddy, 2004, Integration of AFTA and (U-Th)/He thermochronology to enhance the resolution and precision of thermal history reconstruction in the Anglesea-1 well, Otway Basin, SE Australia: PESA Eastern Australasian Basins Symposium II Proceedings, Petroleum Exploration Society of Australia, Adelaide, p. 117–131. Gregory, K. M., and C. G. Chase, 1994, Tectonic and climatic significance of a late Eocene low-relief high-level geomorphic surface, Colorado: Journal of Geophysical Research, v. 99, no. B10, p. 20,141–20,160. Gries, R. R., 1983, North-south compression of the Rocky Mountain foreland structures, in J. D. Lowell and R. R. Gries, eds., Rocky Mountain foreland basins and uplifts: Denver, Rocky Mountain Association of Geologists, p. 9–32. Hagen, S. E., and R. C. Surdam, 1984, Maturation history and thermal evolution of Cretaceous source rocks of the Bighorn Basin, Wyoming and Montana, in J. Woodward, F. F. Meissner, and J. L. Clayton, eds., Hydrocarbon source rocks of the Greater Rocky Mountain region: Denver, Rocky Mountain Association of Geologists, p. 321–338. 6/5/13 7:59 AM 68 Peyton and Carrapa House, M. A., B. P. Wernicke, K. A. Farley, and T. A. Dumitru, 1997, Cenozoic thermal evolution of the central Sierra Nevada, California, from (U-Th)/He thermochronometry: Earth and Planetary Science Letters, v. 151, p. 167–179. Hoy, R. G., and K. D. Ridgway, 1997, Structural and ­s edimentological development of footwall growth ­s ynclines along an intraforeland uplift, east-central Bighorn Mountains, Wyoming: GSA Bulletin, v. 109, p. 915–935. Hoy, R. G., and K. D. Ridgway, 1998, Deformation of Eocene synorogenic conglomerate in the footwall of the Clear Creek thrust fault, Bighorn Mountains, Wyoming: The Mountain Geologist, v. 35, p. 81–90. Kelley, S. A., and D. D. Blackwell, 1990, Thermal history of the multi-well experiment (MWX) site, Piceance Creek Basin, northwestern Colorado, derived from fission-track analysis: Nuclear Tracks and Radiation Measurements, v. 17, p. 331–337. Kelley, S. A., and C. E. Chapin, 1997, Internal structure of the southern Front Range, Colorado, from an apatite fissiontrack thermochronology perspective, in D. W. Bolyard and S. A. Sonnenberg, eds., Geologic history of the Colorado Front Range field trip guidebook: Denver, Rocky Mountain Association of Geologists, p. 19–30. Kelley, S. A., C. E. Chapin, and K. E. Karlstrom, 2001, Laramide cooling histories of Grand Canyon, Arizona, and the Front Range, Colorado, determined from apatite fissiontrack thermochronology, in R. A. Young and E. E. Spamer, eds., Colorado River origin and evolution: Grand Canyon Association, p. 37–42. Kelley, S. A., 2002, Unroofing of the southern Front Range, Colorado: A view from the Denver Basin: Rocky Mountain Geology, v. 37, p. 189–200. Kelley, S. A., and C. E. Chapin, 2004, Denudation history and internal structure of the Front Range and Wet Mountains, Colorado, based on apatite-fission-track thermochronology: New Mexico Bureau of Geology and Mineral Resources, Bulletin, v. 160, p. 41–68. Kelley, S. A., 2005, Low-temperature cooling histories of the Cheyenne Belt and Laramie Peak shear zone, Wyoming, and the Soda Creek-Fish Creek shear zone, Colorado, in K. E. Karlstrom and G. R. Keller, eds., The Rocky Mountain Region: An evolving lithosphere, American Geophysical Union, Geophysical Monograph Series 154, p. 55–70. Ketcham, R. A., 2005, Forward and inverse modeling of lowtemperature thermochronometry data: Reviews in Mineralogy and Geochemistry, v. 58, p. 275–314. Ketcham, R. A., A. Carter, R. A. Donelick, J. Barbarand, and A. J. Hurford, 2007, Improved modeling of fission-track annealing in apatite: American Mineralogist, v. 92, p. 799–810. Kohn, B., C. Spiegel, D. Phillips, and A. Gleadow, 2008, Rubbing out apatite helium age-spread in fast-cooled rocks, in J. I. Garver and M. J. Montario, eds., Proceedings from the 11th International Conference on Thermochronometry, Anchorage, Alaska, p. 144. 10711_ch03_ptg01_hr_037-070.indd 68 Laslett, G. M., P. F. Green, I. R. Duddy, and A. J. W. Gleadow, 1987, Thermal annealing of fission tracks in apatite. 2. A quantitative analysis: Chemical Geology, v. 65, p. 1–13. Lindsey, D. A., P. A. M. Andriessen, and B. R. Wardlaw, 1986, Heating, cooling, and uplift during Tertiary time, northern Sangre de Cristo Range, Colorado: GSA Bulletin, v. 97, p. 1133–1143. Love, J. D., 1960, Cenozoic sedimentation and crustal movement in Wyoming: American Journal of Science, v. 258-A, p. 204–214. Love, J. D., J. C. Reed, R. L. Christiansen, and J. R. Stacy, 1972, Geologic block diagram and tectonic history of the Teton region, Wyoming-Idaho: United States Geological Survey, Miscellaneous Geologic Investigations Map I-730, scale 1:145320. Mackin, J. H., 1947, Altitude and local relief of the Bighorn area during the Cenozoic: Wyoming Geological Association Field Conference Guidebook, v. 2, p. 103–120. McKenna, M. C., and J. D. Love, 1972, High-level strata containing early Miocene mammals on the Bighorn Mountains, Wyoming: American Museum of Natural History Novitates, no. 2490, p. 1–31. McMillan, M. E., P. L. Heller, and S. L. Wing, 2006, History and causes of post-Laramide relief in the Rocky Mountain orogenic plateau: GSA Bulletin, v. 118, p. 393–405. Merewether, E. A., and G. E. Claypool, 1980, Organic composition of some upper Cretaceous shale, Powder River Basin, Wyoming: AAPG Bulletin, v. 64, p. 488–500. Murrell, G. R., E. R. Sobel, B. Carrapa, and P. Andriessen, 2009, Calibration and comparison of etching and thermal modeling techniques for apatite fission-track thermochronology: GSL Special Publications 2009, v. 324, p. 73–85. Naeser, C. W., I. R. Duddy, D. P. Elston, T. A. Dumitru, and P. F. Green, 1989a, Fission-track dating: Ages for Cambrian strata and Laramide and post-middle Eocene cooling events from the Grand Canyon, Arizona, in D. P. Elston, G. H. Billingsley, and R. A. Young, eds., Geology of the Grand Canyon, northern Arizona: American Geophysical Union, p. 139–144. Naeser, C. W., I. R. Duddy, D. P. Elston, T. A. Dumitru, and P. F. Green, 2001, Fission-track analysis of apatite and zircon from Grand Canyon, Arizona, in R. A. Young and E. E. Spamer, eds., Colorado River origin and evolution: Grand Canyon Association, p. 31–35. Naeser, C. W., B. Bryant, M. J. Kunk, K. S. Kellogg, R. A. Donelick, and W. J. Perry Jr., 2002, Tertiary cooling and tectonic history of the White River uplift, Gore Range, and western Front Range, central Colorado: Evidence from fission-track and 39Ar/40Ar ages, in R. M. Kirkham, R. B. Scott, and T. W. Judkins, eds., Late Cenozoic evaporite tectonism and volcanism in west-central Colorado: GSA Special Paper 366, p. 31–53. Naeser, N. D., 1986, Neogene thermal history of the northern Green River Basin, Wyoming—evidence from fissiontrack dating, in D. L. Gautier, ed., Roles of organic matter in sediment diagenesis: Society of Economic Paleontologists and Mineralogists Special Publication 38, p. 65–72. 6/5/13 7:59 AM An Overview of Low-temperature 69 Naeser, N. D., 1989, Thermal history and provenance of rocks in the Wagon Wheel No. 1 well, Pinedale Anticline, northern Green River Basin—evidence from fission-track dating: USGS Bulletin 1886, p. E1–E13. Naeser, N. D., C. W. Naeser, and T. H. McCulloh, 1989b, The application of fission-track dating to the depositional and thermal history of rocks in sedimentary basins, in N. D. Naeser and T. H. McCulloh, eds., Thermal history of sedimentary basins; methods and case histories: London, Springer-Verlag, p. 157–180. Naeser, N. D., 1992, Miocene cooling in the southwestern Powder River Basin, Wyoming; preliminary ­e vidence from apatite fission-track analysis: USGS Bulletin 1917-O, 17 p. Nuccio, V. F., 1990, Burial, thermal, and petroleum generation history of the Upper Cretaceous Steele Member of the Cody Shale (Shannon Sandstone Bed horizon), Powder River Basin, Wyoming: USGS Bulletin 1917-A, 17 p. Nuccio, V. F., 1994a, Vitrinite reflectance data for the Paleocene Fort Union and Eocene Wind River formations, and burial history of a drill hole located in central Wind River Basin, Wyoming: USGS Open-File Report 94-220, p. 1–43. Nuccio, V. F., 1994b, Vitrinite reflectance data for the Paleocene Fort Union and Eocene Wind River formations, and burial history of a drill hole located in central Wind River Basin, Wyoming: USGS Open-File Report OF 94-0220, p. 1–43. Omar, G. I., T. M. Lutz, and R. Giegengack, 1994, Apatite fission-track evidence for Laramide and post-Laramide uplift and anomalous thermal regime at the Beartooth overthrust, Montana-Wyoming: GSA Bulletin, v. 106, p. 74–85. Orme, D. A., 2011, Effects of external parent nuclides on apatite helium dates: M.S. thesis, University of Arizona, Tucson. Osadetz, K. G., B. P. Kohn, P. B. Feinstein, and P. B. O’Sullivan, 2002, Thermal history of the Canadian Williston Basin from apatite fission-track thermochronology: implications for petroleum systems and geodynamic history: Tectonophysics, v. 349, p. 221–249. Osadetz, K. G., B. P. Kohn, S. Feinstein, and R. A. Price, 2004, Foreland belt thermal history using apatite fission-track thermochronology: Implications for Lewis thrust and Flathead fault in the southern Canadian Cordilleran petroleum province., in R. Swennen, F. Roure, and J. W. Granath, eds., Deformation, fluid flow, and reservoir appraisal in foreland fold and thrust belts: AAPG Hedberg Series, no. 1, p. 21–48. Parnell, J., P. F. Green, G. Watt, and D. Middleton, 2005, Thermal history and oil charge on the UK Atlantic margin: Petroleum Geoscience, v. 11, p. 99–112. Pazzaglia, F. J., and S. A. Kelley, 1998, Large-scale geomorphology and fission-track thermochronology in topographic and exhumation reconstructions of the Southern Rocky Mountains: Rocky Mountain Geology, v. 33, p. 229–257. Perry, Jr., W. J., J. C. Haley, D. J. Nichols, P. M. Hammons, and J. D. Ponton, 1988, Interactions of Rocky Mountain 10711_ch03_ptg01_hr_037-070.indd 69 foreland and Cordilleran thrust belt in Lima region, southwest Montana, in C. J. Schmidt, and W. J. Perry Jr., eds., Interaction of the Rocky Mountain foreland and the Cordilleran thrust belt: GSA Memoir 171, p. 267–290. Peyton, S. L., P. W. Reiners, B. Carrapa, and P. G. DeCelles, 2012, Low-temperature thermochronology of the northern Rocky Mountains, western U.S.A.: American Journal of Science, v. 312, p. 145–212. Peyton, S. L., and B. Carrapa, 2013, An introduction to low-temperature thermochronologic techniques, methodology and applications, in C. Knight, J. Cuzella, and L. Cress, eds., The application of structural methods to Rocky Mountain hydrocarbon exploration and development, AAPG Special Publication. Price, L. C., and C. E. Barker, 1985, Suppression of vitrinite reflectance in amorphous rich kerogen—a major unrecognized problem: Journal of Petroleum Geology, v. 8, p. 59–84. Reiners, P. W., and K. A. Farley, 2001, Influence of crystal size on apatite (U-Th)/He thermochronology: an example from the Bighorn Mountains, Wyoming: Earth and Planetary Science Letters, v. 188, p. 413–420. Reiners, P. W., S. N. Thompson, B. J. Tipple, S. L. Peyton, J. M. Rahl, and A. Mulch, 2008, Secondary weathering phases and apatite (U-Th)/He ages: Geochimica et Cosmochimica Acta, v. 72, p. A784. Riihimaki, C. A., R. S. Anderson, and E. B. Safran, 2007, ­I mpact of rock uplift on rates of late Cenozoic Rocky Mountain river incision: Journal of Geophysical Research, v. 112, F03S02. Roberts, L. N. R., T. M. Finn, M. D. Lewan, and M. A. Kirschbaum, 2008, Burial history, thermal maturity, and oil and gas generation history of source rocks in the Bighorn Basin, Wyoming and Montana: USGS Scientific Investigations Report 2008-5037, p. 1–27. Roberts, S. V., and D. W. Burbank, 1993, Uplift and thermal history of the Teton Range (northwestern Wyoming) defined by apatite fission-track dating: Earth and Planetary Science Letters, v. 118, p. 295–309. Royse, F., Jr, M. A. Warner, and D. L. Reese, 1975, Thrust belt structural geometry and related stratigraphic problems Wyoming–Idaho–northern Utah: Rocky Mountain Association of Geologists Symposium, Denver, p. 41–54. Saleeby, J., 2003, Segmentation of the Laramide slab—­ Evidence from the southern Sierra Nevada region: GSA Bulletin, v. 115, p. 655–668. Senftle, J. T., and C. R. Landis, 1991, Vitrinite reflectance as a tool to assess thermal maturity, in R. K. Merrill, ed., Source and migration processes and evaluation techniques: AAPG Treatise of Petroleum Geology, Handbook of Petroleum Geology, p. 119–125. Shuster, D. L., R. M. Flowers, and K. A. Farley, 2006, The influence of natural radiation damage on helium diffusion kinetics in apatite: Earth and Planetary Science Letters, v. 249, p. 148–161. Shuster, M. A., 1986, The origin and sedimentary evolution of the northern Green River Basin, western Wyoming, University of Wyoming, Laramie, 323 p. 6/5/13 7:59 AM 70 Peyton and Carrapa Smithson, S. B., J. Brewer, S. Kaufman, J. Oliver, and C. Hurich, 1978, Nature of the Wind River thrust, Wyoming, from COCORP deep-reflection data and from gravity data: Geology, v. 6, p. 648–652. Steidtmann, J. R., L. T. Middleton, and M. W. Schuster, 1989, Post-Laramide (Oligocene) uplift in the Wind River Range, Wyoming: Geology, v. 17, p. 38–41. Stockli, D. F., K. A. Farley, and T. A. Dumitru, 2000, Calibration of the apatite (U-Th)/He thermochronometer on an exhumed fault block, White Mountains, California: ­Geology, v. 28, p. 983–986. Stone, D. S., 1993, Basement-involved thrust-generated folds as seismically imaged in the subsurface of the central Rocky Mountain foreland, in C. J. Schmidt, C. G. Chase, and E. A. Erslev, eds., Laramide basement deformation in the Rocky Mountain foreland of the western United States, GSA Special Paper, 280, p. 271–318. Strecker, U., 1996, Studies of Cenozoic continental tectonics: Part I, Apatite fission-track thermochronology of the Black Hills uplift, South Dakota; Part II. Seismic sequence stratigraphy of the Goshute Valley halfgraben, Nevada, Ph.D. thesis, University of Wyoming, Laramie, University of Wyoming, p. 344. Sweeney, J. J., and A. K. Burnham, 1990, Evaluation of a ­simple model of vitrinite reflectance based on chemical kinetics: AAPG Bulletin, v. 74, p. 1559–1570. 10711_ch03_ptg01_hr_037-070.indd 70 Tissot, B. P., and D. H. Welte, 1978, Petroleum formation and occurrence: A new approach to oil and gas exploration: Berlin, Springer-Verlag, 521 p. Waples, D. W., 1980, Time and temperature in petroleum formation; application of Lopatin’s method to petroleum exploration: AAPG Bulletin, v. 64, p. 916–926. Waples, D. W., and M. Ramly, 2001, A statistical method for correcting log-derived temperatures: Petroleum Geoscience, v. 7, p. 231–240. Waples, D. W., J. Pacheco, and A. Vera, 2004a, A method for correcting log-derived temperatures in deep wells, calibrated in the Gulf of Mexico: Petroleum Geoscience, v. 10, p. 239–245. Waples, D. W., M. R. Pedersen, and P. Kuijpers, 2004b, Correction of single log-derived temperatures in the Danish Central Graben, North Sea: Natural Resources Research, v. 13, p. 229–239. Warnock, A. C., P. K. Zeitler, R. A. Wolf, and S. C. Bergman, 1997, An evaluation of low-temperature apatite U-Th/ He thermochronometry: Geochimica et Cosmochimica Acta, v. 61, p. 5371–5377. Yonkee, W. A., 1992, Basement-cover relations, Sevier orogenic belt, northern Utah: GSA Bulletin, v. 104, p. 280–302. 6/5/13 7:59 AM