<<

faults are exposed at the surface. In most Reactivation, trishear modeling, places, however, they remain blind, either covered by Cenozoic sediments, or ex- and folded basement in Laramide pressed at the surface as a monoclinal -propagation of Paleozoic and Mesozoic strata. Although the geometry uplifts: Implications for the origins of intra- and kinematics of some of these faults are relatively well understood, their origin (date of initial rupture) is often difficult to continental faults establish. Two end-member hypotheses have been proposed: (1) they formed Alexander P.Bump*, Department of Geosciences, University of Arizona, Tucson, during the Laramide (Hamilton, Arizona 85721, USA 1988); and (2) they formed at some previ- ous time(s) and were reactivated during ABSTRACT INTRODUCTION the Laramide (Walcott, 1890; Huntoon, Laramide uplifts are bounded by re- The processes of fault reactivation 1993; Marshak et al., 2000). verse faults of enigmatic origin. Two (Holdsworth et al., 2001) and tectonic in- Perhaps the more favored model has end-member hypotheses have been version (Coward, 1994) of old faults are been that many, if not most, Laramide proposed: (1) they formed during the among the most important issues in conti- faults are reactivated ancient weaknesses, Laramide orogeny as newly formed con- nental . Continental crust, unlike and in a few cases this can be proven di- tractional features; and (2) they formed oceanic crust, records the cumulative his- rectly. For example, Laramide reverse as normal faults at some previous time tory of multiple periods of tectonism, and faulting raised Precambrian strata and were reactivated during the Laramide. much of continental deformation is fo- along ancient faults in the Uinta This paper employs the trishear fault-fold cused into large-scale fault networks that Mountains of Utah and the Beltian em- model to test these ideas, based on the are repeatedly reactivated over long time bayment of Montana. Likewise, Walcott premise that, in (1), neoformed faults scales. The seismogenic upper crust may (1890), Huntoon (1993), and Timmons et should propagate from a regional detach- be the strongest layer (Jackson, 2002), al. (2001) documented exposures in the ment or crustal flaw within the crystalline and hence a guide for whole-litho- Grand Canyon with Precambrian syn- basement, whereas in (2), reactivated sphere deformation (Axen et al., 1998). extensional strata present in the hanging normal faults should begin their To understand a range of seismogenic wall of Laramide reverse faults but not in Laramide propagation from the base of processes, it is essential to better resolve the footwall. Similar evidence for reactiva- the Paleozoic cover (i.e., top of base- the interplay between old fault networks tion has also been reported for Laramide ment). Trishear folding takes place en- and new tectonic stress conditions faults exposed in the Salt River Canyon in tirely ahead of the propagating fault tip, (Huntoon, 1993; Marshak et al., 2000; central Arizona (Davis et al., 1981) and so trishear modeling of fold-fault geome- Timmons et al., 2001). This paper uses exhumed fault blocks in southwestern tries can be used to evaluate these alter- the unique geometries of the Laramide Montana (Schmidt and Garihan, 1983). nate possibilities. This paper documents uplifts in the southwestern United States Detailed studies of fault zone rocks have 23 uplifts that show a folded basement- to explore the importance of reactivation also yielded clear evidence of reactiva- cover contact demonstrating that fault versus new fault development in response tion, showing diachronous deformation at propagation began within the basement. to Laramide contractional deformation. different metamorphic grades (Mitra and Inverse and forward modeling suggest, The Late Cretaceous–early Tertiary Frost, 1981). however, that the faults began propagat- Laramide orogeny has been of particular These examples are exceptions, how- ing only a few kilometers below the interest to tectonicists because the hori- ever. While many Laramide uplifts offer basement-cover contact, too shallow for zontal shortening due to plate conver- excellent hanging wall exposure, the a regional detachment. It is suggested gence was expressed within the continen- footwalls and the uplift-bounding faults that these faults represent the formation tal foreland at distances up to twice as far are often buried beneath significant of footwall shortcuts (i.e., lower angle, from the plate margin as the more typical thicknesses of syn-orogenic and younger mechanically easier paths to the surface) and partly contemporaneous Sevier fold- sediments (Tweto, 1979; Love and to bypass the steep upper sections of re- thrust belt (Fig. 1). Style of contractional Christiansen, 1985). Many other activated listric faults. This idea unites the deformation in the Laramide province is Laramide faults never broke the surface two end-member hypotheses, allowing also different from that of the thrust belt, and are expressed only as monoclinal the large-scale map pattern of uplifts to involving a series of fault-bounded, base- fault-propagation folds (Tweto, 1979; be controlled by reactivated faults at ment-cored uplifts of varying size, orien- Hintze, 1980; Love and Christiansen, depth while exposing neoformed off- tation, and structural relief. In a few 1985). Seismic data often do not have shoots of those faults at the surface. places, such as Rattlesnake Mountain and the resolution necessary to document the Grand Canyon, the uplift-bounding stratigraphic evidence for , and ————— *Present address: BP North American Exploration, 501 Westlake Park Boulevard, Houston, Texas 77079, USA, [email protected].

4 MARCH 2003, GSA TODAY on either side of the province. Some authors (Hamilton, 1988; Yin, 1994) have opposed the idea of widespread reactivation, arguing instead that the major uplift-bounding Laramide faults are neoformed features created dur- ing the Laramide orogeny. Indeed, most faults show no direct evidence of a pre-Laramide history. The purpose of this paper is to exam- ine evidence for fault ancestry based on a relatively new criterion: the location of the initial fault tip (Allmendinger and Shaw, 2000; Allmendinger et al., 2003). The initial fault tip location is defined here as the point from which the fault tip began its upward propagation (growth) under Laramide compression. A preexist- ing fault reactivated in compression could be expected to begin propagating up- ward from the base of the postorogenic strata (Allmendinger et al., 2003). In the case of the Laramide, that would indicate an initial fault tip depth at the top of the Precambrian, i.e., at the basement-cover contact. Alternatively, a neoformed re- verse fault could be expected to do one of two things: either it might propagate upward from a horizontal detachment near the brittle-ductile transition where stresses are theoretically highest (Sibson, 1977; Ranalli, 2000), or it might begin with simultaneous up-dip and down-dip propagation from some initial earthquake focus at a flaw or stress con- centration in the crust (Allmendinger et al., 2003).

FAULT-PROPAGATION FOLD MODELS For Laramide faults that began growing in the Late Cretaceous or early Tertiary and are still buried, there is no way to observe the initial tip point directly. Figure 1. Map of the Laramide orogen (after King, 1969), showing locations of uplifts discussed in the text. Numbers 24 and 25 are the San Rafael Swell and the Circle Cliffs uplifts, However, its location can be inferred respectively. All others are given in Table 1. through geometric modeling of fold-fault relationships based on structural expo- boreholes seldom penetrate basement in strikes (Fig. 1) that are similar to those of sures at the surface and available subsur- the footwall, negating the possibility of known Precambrian in the southern face information. The key parameter is documenting differences in Precambrian Rocky Mountains (Erslev, 1993; Marshak the propagation to slip (p/s) ratio, which stratigraphy (if present) in the hanging et al., 2000; Timmons et al., 2001). describes the distance the fault tip moves wall and footwall. Thus, for most Second, in some cases, Laramide faults (propagates) per unit displacement of the Laramide faults, the argument for reactiva- parallel nearby Precambrian dikes or hanging wall. The p/s ratio is an inherent tion hinges chiefly on circumstantial evi- shear zones, suggesting that their geome- feature of all kinematic models of fault- dence. First, Laramide fault strikes show a try was governed by associated propagation folding, though most treat it scatter uncommon in neoformed fault Precambrian weaknesses (Schmidt and only implicitly (Suppe and Medwedeff, systems such as the extensional faults of Garihan, 1983). Third, Marshak et al. 1990; Narr and Suppe, 1994; Mitra and the Basin and Range. More specifically, (2000) pointed out that vergence varia- Mount, 1998). The power of the p/s ratio the large-scale map pattern of Laramide tions within the Laramide province resem- lies in its capacity to determine the loca- faulting is dominated by N-S and NW-SE ble those of rift provinces, with opposite tion of the fault tip prior to fault slip,

GSA TODAY, MARCH 2003 5 Figure 2. Schematic diagram of the trishear model (Erslev, 1991). Thick black line shows fault plane while red lines define the edges of the active trishear zone. Shading within the trishear zone reflects the intensity of deformation. Horizontal blue lines define stratigraphy. Note that strata below the initial fault tip remain unfolded after slip. Here deformation is distributed evenly across the trishear envelope (center concentration factor = 1). based only on geometric modeling of the within the trishear envelope and may ac- that the basement surface is planar, and post-slip fold shape (Allmendinger and cumulate very high strains. At higher p/s orientations show no sign of rotation Shaw, 2000). values, a given packet passes quickly at distances greater than 100 m from the To date, the only model to explicitly through the trishear zone and therefore fault (Stearns, 1978; Erslev and Rogers, consider the p/s ratio is the trishear model emerges relatively unstrained. Similarly, a 1993; Narr, 1993). Published seismic sur- (Erslev, 1991; Hardy and Ford, 1997; high ccf will concentrate deformation in veys reveal a number of other examples Allmendinger, 1998). This model accom- the center of the trishear zone, resulting (Gries and Dyer, 1985; Stone, 1993a). modates waning fault slip by folding in a local zone of highly strained rock. In other cases, surface exposures, seis- within a triangular zone, the apex of For the purposes of this paper, the mic lines, and well-controlled cross sec- which is pinned to the fault tip (Fig. 2A). most important point is that all trishear tions clearly show that the basement sur- Within this “trishear zone,” particle veloci- folding occurs ahead of the initial location face is folded. (Here the term fold is used ties diminish along tie lines perpendicular of the fault tip. The presence or absence only to denote a curvature of the base- to the fault, from a maximum in the hang- of folding adjacent to the fault at a given ment-cover contact, regardless of whether ing wall to zero in the footwall. The re- horizon can thus be used as a clue to that curvature is achieved by flexural slip sulting fold geometry depends on seven whether the fault began propagation from on subhorizontal , distributed variables, including the initial x and y lo- above or below that horizon. In the pre- faulting, or some other means.) This fold- cations of the fault tip, the fault dip (ramp sent case, it can be used to determine ing occurs over a half wavelength of 100 angle), the total fault slip, the trishear an- whether a given fault began propagation m to 5 km (Table 1), where half wave- gle, the p/s ratio, and the center concen- at or below the basement-cover contact. length is defined as the horizontal dis- tration factor (ccf), which determines how tance from flat-lying hanging wall to flat- folding is distributed within the trishear LARAMIDE BASEMENT GEOMETRY lying footwall, or fault if the footwall is not zone (Fig. 2B; Erslev, 1991; Erslev and In the cases of the largest uplifts, such imaged. Typical values are 500–1000 m. Rogers, 1993; Hardy and Ford, 1997; as the Colorado Front Range, erosion has Within the fold, the basement surface Allmendinger, 1998). Fault dip and the lo- removed the sedimentary cover and the may reach dips of 90° or even 75° over- cation of the fault tip can often be deter- uppermost basement, rendering it difficult turned. At the other extreme, a few uplifts mined from seismic data, and net slip or impossible to establish the geometry of display basement surfaces that never ex- may be determined from stratigraphic the basement surface near the bounding ceed 10° of dip. Most fall in the range of separation across the fault. Of the remain- faults. Outcrop exposures and seismic im- 30°–60° (Table 1). ing three variables, p/s and ccf exert the ages of smaller uplifts, however, reveal a It is important to point out that there most important control on fold shape range of forms. are several possible mechanisms for fold- (Allmendinger et al., 2003; E.A. Erslev, In some of these, the basement uncon- ing the basement surface, not all of them 2003, personal commun.). The p/s ratio formity is clearly an undeformed, planar requiring an initial fault tip below that sur- determines how long a given packet of surface. Careful surveys of extensive ex- face. First, damage incurred during fault rock remains within the trishear envelope posures on Rattlesnake Mountain in slip might result in local curvature of the and consequently, how severely that northwest Wyoming and the Big basement surface adjacent to the fault. In packet is deformed. At low p/s values, Thompson on the eastern edge the cases described above, however, the packets spend a relatively long time of the Colorado Front Range both indicate relatively large horizontal scale of the

6 MARCH 2003, GSA TODAY One of the best-fitting models I have produced to date is for the Waterpocket , which forms the steep eastern limb of the Circle Cliffs uplift in southern Utah (Fig. 1). Inverse modeling based on well logs and surficial exposures gives a reasonably good, though imperfect fit to the data (Fig. 3A). The principal problem is that the modeled fold wavelength is `too short, lacking the prolonged, gentle upper-limb dip observed in the field, and shallowing too quickly near the lower hinge. Forward modeling offers the possi- bility of improving the fit by allowing both spatial and temporal heterogeneity in trishear parameters. In the temporal case, the p/s ratio may be varied in accor- dance with the mechanical stratigraphy through which the fault propagates (Allmendinger, 1998). For the Waterpocket fold, initially rapid propaga- tion (p/s = 6.0) through stiff basement rocks followed by slower propagation (p/s = 2.1) through less competent sedi- mentary rocks can produce the observed broad-wavelength fold with long, gently dipping upper and lower limbs and sharply steeper middle limb. Similar re- Figure 3. Inverse (A) and forward (B) models of the Waterpocket fold, based on mapping of a surface transect and regional thicknesses of undeformed strata. Elevations are in meters above sults may also be achieved through spa- sea level and cross sections are drawn without vertical exaggeration. Color bar on each side of tial heterogeneity, that is, by allowing cross sections shows ages of strata and vertical contact locations based on well logs and folding to be concentrated toward the projection from the surface. C: Contour plots of errors in inverse modeling. Plots show two- middle of the trishear zone, similar to de- dimensional slices through the three-dimensional matrix of error values produced by grid- formation within a ductile searching for a best-fit over the specified ranges. Best fit is based on a chi-squared statistical (Erslev, 1991). The best-fit forward model analysis (Allmendinger, 1998). Note that all plots show well-defined regions of best fit. is an excellent match for the data (Fig. 3B) and requires an initial fault tip 2.3 km deformation argues against this. Second, (1993a) based on study of relations be- below the basement-cover contact. the basement surface may fold as a pas- tween fold shape and net fault slip. For other uplifts listed in Table 1, in- sive response to bends in the underlying verse modeling is limited by the complex- fault (fault bend folding). Nearly all of the TRISHEAR MODELING ity of observed fault geometries. Taking examples cited in Table 1 display some Accepting that folding of the basement only those examples where both hanging fault curvature, so this mechanism may surface may reflect Laramide fault propa- wall and footwall are imaged and where well be responsible for some of the fold- gation from a point below that surface, a >90% of the displacement is concentrated ing described above. However, the fault key question is how far below that sur- on a single fault leaves 15 possibilities for bends are generally quite small relative to face did fault propagation begin. An esti- inversion. Of these, one is a seismic time the magnitude of basement folding. mate may be obtained from accurate section and is rejected for having too little Finally, folding of the basement surface trishear modeling of the structures in true depth control. Three more have sig- could be produced by buckling as a result question. Current trishear modeling soft- nificantly curved faults and are also re- of a “room problem” set up by horizontal ware allows an inverse grid search jected. Of the remaining 11, only three shortening on a concave-upward fault wherein each trishear model parameter is have yielded reasonable inverse solutions. (Coward, 1994). In such cases, however, systematically and independently varied The eight failures are probably due to a folding is restricted entirely to the hanging over a user-specified range while the pro- combination of oblique slip and geomet- wall, whereas many of the examples cited gram searches for a best-fit model. ric complexities, including fault bends, in Table 1 show folding of the basement Although powerful, these programs are problems in pinpointing the current loca- surface in both hanging wall and foot- best applied to relatively simple structures tion of the fault tip, and accurately migrat- wall. For these reasons, it seems likely and are not capable of handling some of ing seismic data to show true depth. that fault-propagation folding created the complexities commonly observed in Results for the three reasonable inverse much of the observed basement curva- the field, such as multiple fault strands, models are quite varied. The bounding ture, a conclusion also reached by Stone fault bends, or oblique slip. fault on Casper Mountain appears to have

GSA TODAY, MARCH 2003 7 in the inversion space that might be con- faults (Fig. 4). Almost all normal faults fused with the global minimum (Fig. 3C). show some degree of concave-upward That said, experience with the trishear curvature (Coward, 1994). At the topo- modeling software shows that inverse graphic surface, failure is often tensile or models can almost always be refined and hybrid tensile-shear and the resulting fault improved through forward modeling dips are commonly near vertical. Dips de- (Allmendinger et al., 2003) and the pre- crease toward 60° at 3–4 km depth due to sent models are probably no different. the change in failure mechanism from hy- Furthermore, common sense suggests brid to shear fracturing (Walsh and that with seven independently variable Watterson, 1988). Proffet (1977) and parameters, there are probably other solu- Hamblin (1965) documented Basin and tions that satisfy the data. At present, the Range faults with dips that decreased at a models described above are best-fits, but rate of 0.5°–2° per 100 m of depth at future efforts and continued development near-surface levels. For crustal-scale faults, of trishear software may lead to further dips must decrease further toward 45° at refinements. ~10 km depth as the host rock rheology changes from brittle to plastic (Walsh and DISCUSSION AND CONCLUSIONS Watterson, 1988). Histograms showing The evidence described above suggests numbers of seismically active normal that there is a wide spectrum of fault ori- faults versus fault dip typically exhibit gins. Some uplifts, such as Rattlesnake strong peaks at 45° (Thatcher and Hill, Mountain and the Big Thompson anti- 1991; Collettini and Sibson, 2001). Finally, cline, show no evidence for deformation if the fault is detached at some deeper of the basement-cover contact and thus level, then it must eventually bend toward are probably bounded by ancient faults horizontal. Seismic activity has been doc- reactivated under Laramide compression. umented on normal faults with dips as Many others (Table 1) do show deforma- low as 30° (Collettini and Sibson, 2001). Figure 4. Schematic diagram of the creation tion of the basement-cover contact. These The structural level of the current base- of a listric normal fault (top) that is later appear to represent cases in which the ment surface with respect to Precambrian reactivated in compression (bottom), creating bounding faults began propagation from a footwall shortcut. Pre-extensional rocks are normal faults is uncertain. The Grand and deeper in the crust. Some of these, such shown in dark gray. Salt River Canyons expose Precambrian as the Island Park fault, are probably normal faults dipping 60°–85° at the base- nucleated only 400–600 m below the entirely neoformed faults that branch off ment-cover contact (Davis et al., 1981; basement surface, while the Island Park possibly older ones in the mid crust, Huntoon, 1993), which suggests a shal- fault bounding the Uinta uplift appears to forming as backthrusts in response to low structural level. In the Grand Canyon, have begun propagation from a point 11 Laramide contraction. These end mem- the hanging walls are composed of syn-rift km below the basement surface. This is bers are readily understood. sedimentary rocks (Timmons et al., 2001), very close to the 12 km deep forethrust- Several other faults, however, appear to also consistent with ideas suggesting that backthrust junction shown by Stone have begun propagation from a few kilo- the current basement surface may not (1993b) in a cross section of the Uintas. meters below the basement surface. The have been far below the syn-extensional The fault underlying the Willow Creek origin of these faults is less clear, as there topographic surface. On the other hand, anticline is in the middle, with an initial is no obvious reason for a fault to begin Ar-Ar evidence suggests that >6 km of tip ~6 km below the basement surface propagation there. It is possible that they rock was eroded between widespread ex- (Fig. 3). Allmendinger’s (1998) inversion represent out-of-the- thrusts tensional faulting at 800 and 1100 Ma and of the Rangely anticline in Colorado (Brown, 1993), formed during the creation Cambrian deposition (Heizler et al., shows an initial fault tip ~4 km below the of larger folds. However, this requires that 2000), which would indicate that current basement surface, and an initial fault tip the folds described here be located in the exposures represent deeper levels of the depth of 0.6 km below the basement sur- proximal footwall of a larger fold, which Precambrian normal fault systems face for the San Rafael monocline in cen- is not true in general. It is also possible (Marshak et al., 2000) that were reacti- tral Utah has also been found (personal that these faults nucleated around point vated in the Laramide. In either case, observations; G.H. Davis, 2002, personal weaknesses within the basement, perhaps however, the exposed faults are steep commun.). flaws or stress concentrators (Eisenstadt where they intersect the basement uncon- The question arises as to whether any and DePaor, 1987; Allmendinger et al., formity. of these solutions are unique. Contour 2003), though there is no clear reason to Under horizontal compression, it is eas- plots of the inverse modeling results (Fig. expect them at this crustal level. iest to reactivate faults dipping 25°–40° 3) suggest that the solution space is small Instead, I propose that they are foot- (Byerlee, 1978; Ranalli, 2000). At higher and in controlled experiments with wall shortcuts (McClay, 1989; Coward, dips, it is often easier to create a new, trishear inversions, Allmendinger et al. 1994), created by the compressional in- lower-angle fault than to reactivate the (2003) have found no local error minima version of upward-steepening normal old, steep one. Dip catalogs of seismically

8 MARCH 2003, GSA TODAY active reverse faults typically show a cut Laramide fault geometry to that of rift sys- ACKNOWLEDGMENTS off at 60° (Sibson and Xie, 1998; Collettini tems exposed in the southern Rockies, The manuscript has benefited from dis- and Sibson, 2001). Upward-steepening and Timmons et al. (2001) documented cussion with David Ferrill, Clem Chase, normal faults are thus susceptible to pure multiple episodes of Precambrian exten- Ofori Pearson, and George Davis, and re- reverse-sense reactivation only where sional faulting. At the same time, the fact view by Don Wise, Jon Spencer, Gautam they dip less than ~60°, i.e., in their lower remains that relatively few Laramide faults Mitra, Eric Erslev, John Suppe, and Karl extents, at depths >3–4 km beneath the show unequivocal evidence of reactiva- Karlstrom. Special thanks go to Eric Erslev contemporary topographic surface. The tion. This work suggests that ancient who pointed out ambiguities in the trishear upper, steep portion is difficult or impos- faults can be reactivated at depth but approach to Laramide structures. Trishear sible to reactivate in pure reverse motion form new paths to the surface. The large- software was generously provided by which often leads to the creation of one scale geometry of the orogen may thus be Rick Allmendinger and Nestor Cardozo. or more “footwall shortcuts,” i.e., splays controlled by ancient structures but the Funding was provided by National that branch off the old fault and form surficially exposed segments of the faults Science Foundation grant EAR-0001222 new, lower angle routes to the surface, need not be ancient themselves. (awarded to George Davis). My thanks bypassing the steep segment of the origi- In the broader picture, this work sug- to them all. nal fault (Fig. 4; McClay, 1989; Coward, gests that there is an entire spectrum of 1994). These may be unequivocally iden- faults ranging from 100% neoformed to REFERENCES CITED tified if the abandoned portion of the fault 100% reactivated. Many, perhaps even a Allmendinger, R.W., 1998, Inverse and forward numerical modeling of trishear fault-propagation folds: Tectonics, v. 17, retains normal-sense offset whereas the majority, of intracontinental faults lie be- p. 640–656. new splay shows reverse offset. The tween these end members. Under com- Allmendinger, R.W., and Shaw, J.H., 2000, Estimation of fault hanging wall may also include a half pressive stress, weak sections of existing propagation distance from fold shape: Implications for earth- quake hazard assessment: Geology, v. 28, p. 1099–1102. of older sedimentary rocks and/or faults may localize the initial failure. Allmendinger, R.W., Zapata, T., Manceda, R., and Dzelalija, an ancient shear zone of steeper dip than Bends or other strength heterogeneities in F., 2003, Trishear kinematic modeling of structures with ex- the uplift-bounding fault. those faults, however, may necessitate the amples from the Neuquen Basin, Argentina: American Association of Petroleum Geologists (in press). In the spirit of Davis (1926) and Wise growth of new, more favorably oriented Axen, G.J., Selverstone, J., Byrne, T., and Fletcher, J.M., 1998, (1963), this is both a hypothesis to be segments if slip continues. Perhaps the re- If the strong crust leads will the weak crust follow?: GSA tested and an attractive idea as it offers a sulting fault, such as might typify the Today, v. 8, no. 12, p. 1–8. means to unify the end-member interpre- Laramide or any other zone of intracra- Brown, W.G., 1993, Structural style of Laramide basement- cored uplifts and associated folds, in Snoke, A.W., et al., eds., tations. The scatter in Laramide fault tonic shortening, is thus a hybrid of Geology of Wyoming: Geological Survey of Wyoming strikes may well reflect tectonic inheri- linked ancient and neoformed segments Memoir No. 5, p. 312–371. Byerlee, J.D., 1978, Friction of rocks: Pure and Applied tance. Marshak et al. (2000) presented a and exhibits characteristics of both. Geophysics, v. 116, p. 615–626. compelling comparison of the large-scale

GSA TODAY, MARCH 2003 9 Chase, R.B., Schmidt, C.J., and Genovese, P.W., 1993, Huntoon, P.W., 1993, Influence of inherited Precambrian kinematics of twelve folds in the Rocky Mountain foreland, Influence of Precambrian rock compositions and fabrics on basement structure on the localization and form of Laramide in Schmidt, C.J., et al., eds., Laramide basement deformation the development of Rocky Mountain foreland folds, in , Grand Canyon, Arizona, in Schmidt, C.J., et al., in the Rocky Mountain foreland of the western United States: Schmidt, C.J., et al., eds., Laramide basement deformation in eds., Laramide basement deformation in the Rocky Mountain Boulder, Colorado, Geological Society of America Special the Rocky Mountain foreland of the western United States: foreland of the western United States: Boulder, Colorado, Paper 280, p. 1–44. Boulder, Colorado, Geological Society of America Special Geological Society of America Special Paper 280, Sibson, R.H., 1977, Fault rocks and fault mechanisms: Journal Paper 280, p. 45–72. p. 243–256. of the Geological Society of London, v. 133, p. 191–213. Collettini, C., and Sibson, R.H., 2001, Normal faults, normal Jackson, J., 2002, Strength of the continental lithosphere: Time Sibson, R.H., and Xie, G.Y., 1998, Dip range for intraconti- friction?: Geology, v. 29, p. 927–930. to abandon the jelly sandwich?: GSA Today, v. 12, no. 9, nental reverse fault ruptures: Truth not stranger than friction?: p. 4–10. Coward, M., 1994, Inversion tectonics, in Hancock, P., ed., Bulletin of the Seismological Society of America, v. 88, Continental deformation: New York, Pergamon Press, King, P.B., 1969, Tectonic map of North America: p. 1014–1022. p. 289–304. Washington, U.S. Geological Survey. Stearns, D.W., 1978, Faulting and forced folding in the Rocky Davis, G.H., Showalter, S.R., Benson, G.S., and McCalmont, Lange, J.K., and Wellborn, R.E., 1985, Seismic profile: North Mountain foreland, in Matthews, V.I., ed., Laramide folding L.S., 1981, Guide to the geology of the Salt River Canyon re- Park Basin, in Gries, R.R., and Dyer, R.C., eds., Seismic explo- associated with basement block faulting in the western United gion, Arizona: Arizona Geological Society Digest, v. 13, ration of the Rocky Mountain region: Denver, Rocky States: Boulder, Colorado, Geological Society of America p. 49–97. Mountain Association of Geologists and Denver Geological Memoir 151, p. 1–37. Society, p. 239–245. Davis, W.M., 1926, The value of the outrageous geological Stone, D.S., 1993a, Basement-involved thrust-generated folds hypothesis: Science, v. 63, p. 463–468. Love, J.D., and Christiansen, A.C., 1985, Geologic map of as seismically imaged in the subsurface of the central Rocky Wyoming: Reston, Virginia, U.S. Geological Survey. Mountain foreland, in Schmidt, C.J., et al., eds., Laramide Eisenstadt, G., and DePaor, D., 1987, Alternative model of basement deformation in the Rocky Mountain foreland of the thrust-fault propagation: Geology, v. 15, p. 630–633. Marshak, S., Karlstrom, K., and Timmons, J.M., 2000, western United States: Boulder, Colorado, Geological Society Inversion of Proterozoic extensional faults: An explanation for Erslev, E.A., 1991, Trishear fault-propagation folding: of America Special Paper 280, p. 271–312. the pattern of Laramide and Ancestral Rockies intracratonic Geology, v. 19, p. 617–620. deformation, United States: Geology, v. 28, p. 735–738. Stone, D.S., 1993b, Tectonic evolution of the Uinta moun- Erslev, E.A., 1993, Thrusts, back-thrusts, and detachment of tains: Palinspastic restoration of a structural cross-section McClay, K.R., 1989, Analogue models of inversion tectonics, Rocky Mountain foreland arches, in Schmidt, C.J., et al., eds., along longitude 109º15’, Utah: Salt Lake City, Utah in Cooper, M.A., and Williams, G.D., eds., Inversion Laramide basement deformation in the Rocky Mountain Geological Survey, 19 p. tectonics: London, Geological Society of London Special foreland of the western United States: Boulder, Colorado, Publication 44, p. 41–59. Suppe, J., and Medwedeff, D.A., 1990, Geometry and kine- Geological Society of America Special Paper 280: p. matics of fault-propagation folding: Ecolgae Geologicae 339–359. Mitra, G., and Frost, B.R., 1981, Mechanisms of deformation Helvetiae, v. 83, p. 409–454. within Laramide and Precambrian deformation zones in Erslev, E.A., and Rogers, J.L., 1993, Basement-cover geometry basement rocks of the Wind River Mountains: University of Thatcher, W., and Hill, D.P., 1991, Fault orientations in exten- of Laramide fault-propagation folds, in Schmidt, C.J., et al., Wyoming Contributions to Geology, v. 19, p. 161–173. sional and conjugate strike-slip environments and their impli- eds., Laramide basement deformation in the Rocky Mountain cations: Geology, v. 19, p. 1116–1120. foreland of the western United States: Boulder, Colorado, Mitra, S., and Mount, V.S., 1998, Foreland basement-involved Geological Society of America Special Paper 280, structures: American Association of Petroleum Geologists Timmons, J.M., Karlstrom, K.E., Dehler, C.M., Geissman, J.W., p. 125–146. Bulletin, v. 82, p. 70–109. and Heizler, M.T., 2001, Proterozoic multistage (ca. 1.1 and 0.8 Ga) extension recorded in the Grand Canyon Supergroup Gries, R., 1983, North-south compression of Rocky Mountain Narr, W., 1993, Deformation of basement in basement-in- and establishment of northwest- and north-trending grains in foreland structures, in Lowell, J.D., ed., Rocky Mountain fore- volved compressive structures, in Schmidt, C.J., et al., eds., the southwestern United States: Geological Society of land basins and uplifts: Denver, Rocky Mountain Association Laramide basement deformation in the Rocky Mountain America Bulletin, v. 113, p. 163–180. of Geologists, p. 9–32. foreland of the western United States: Boulder, Colorado, Geological Society of America Special Paper 280, Tweto, O., 1979, Geologic map of Colorado: Denver, Gries, R.R., and Dyer, R.C., 1985, Seismic exploration of the p. 107–124. Colorado, U.S. Geological Survey. Rocky Mountain region: Denver, Rocky Mountain Association of Geologists and Denver Geophysical Society, p. 300. Narr, W., and Suppe, J., 1994, Kinematics of basement-in- Walcott, C.D., 1890, Study of a line of displacement in the volved compressive structures: American Journal of Science, Grand Canyon of the Colorado, in northern Arizona: Hamblin, W.K., 1965, Origin of “reverse drag” on the down- v. 294, p. 802–860. Geological Society of America Bulletin, v. 1, p. 49–64. thrown side of normal faults: Geological Society of America Bulletin, v. 76, p. 1145–1164. Proffett, J.M., Jr., 1977, Cenozoic geology of the Yerington Walsh, J.J., and Watterson, J., 1988, Dips of normal faults District, Nevada, and implications for the nature and origin of in British Coal Measures and other sedimentary sequences: Hamilton, W.B., 1988, Laramide crustal shortening, in Basin and Range faulting: Geological Society of America Journal of the Geological Society of London, v. 145, Schmidt, C.J., and Perry, W.J., Jr., eds., Interaction of the Bulletin, v. 88, p. 247–266. p. 859–873. Rocky Mountain foreland and the Cordilleran thrust belt: Boulder, Colorado, Geological Society of America Memoir Ranalli, G., 2000, Rheology of the crust and its role in tectonic Wise, D.U., 1963, An outrageous hypothesis for the tectonic 171, p. 27–40. reactivation: Journal of , v. 30, p. 3–15. pattern of the North American Cordillera: Geological Society of America Bulletin, v. 74, p. 357–362. Hardy, S., and Ford, M., 1997, Numerical modeling of trishear Ray, R.R., and Keefer, W.R., 1985, Wind River Basin, central fault propagation folding: Tectonics, v. 16, p. 841–854. Wyoming, in Gries, R.R., and Dyer, R.C., eds., Seismic explo- Wise, D.U., and Obi, C.M., 1992, Laramide basement defor- ration of the Rocky Mountain Region: Denver, Rocky mation in an evolving stress field, Bighorn mountain front, Heizler, M.T., Timmons, J.M., and Karlstrom, K.E., 2000, Mountain Association of Geologists and Denver Geological Five Springs area, Wyoming: American Association of Ancestry of Ancestral Rocky Mountain faults: Geological Society, p. 201–212. Petroleum Geologists Bulletin, v. 76, p. 1586–1600. Society of America Abstracts with Programs, v. 32, no. 7, p. A-465. Schmidt, C.J., and Garihan, J.M., 1983, Laramide tectonic de- Yin, A., 1994, Mechanics of monoclinal systems in the velopment of the Rocky Mountain foreland of southwestern Colorado Plateau during the Laramide orogeny: Journal of Hintze, L.F., 1980, Geologic map of Utah: Salt Lake City, Montana, in Lowell, J.D., ed., Rocky Mountain foreland Geophysical Research, v. 99, p. 22,043–22,058. Utah Geological and Mineral Survey. basins and uplifts: Denver, Rocky Mountain Association of Holdsworth, J.A., Buick, R.E., and Hand, M., 2001, Geologists, p. 271–294. Manuscript submitted November 14, Continental reactivation and reworking: London, Geological Schmidt, C.J., Genovese, P.W., and Chase, R.B., 1993, Role of 2002; accepted January 15, 2003. ❖ Society of London Special Publication 184, 408 p. basement fabric and cover-rock lithology on the geometry and

10 MARCH 2003, GSA TODAY