Tectonic Wedging in the California Coast Ranges: Useful Model or an Unreasonable Hypothesis?

Maureen Berlin
GEOL 5690, Tectonics of the Western U.S., Fall 2005

Rocks in California (Figure 1) record a long history of subduction during the Mesozoic and early Cenozoic, followed by strike-slip offsets along the San Andreas fault system during the late Cenozoic. This Andean-type continental margin (Figure 2) produced an arc-trench system with corresponding arc volcanism, forearc deposition, and subduction and accretion of terranes.



Figure 1. California, showing place names, geologic provinces, and selected geologic units. From Fuis and Mooney 1990.

Figure 2 . Schematic diagram showing California as an Andean-type continental margin during late Mesozoic time (from Wallace 1990, modified from Dickinson 1981).

The history of Cenozoic deformation in the Franciscan subduction complex has been the subject of considerable debate. Various workers have assembled information on the surficial geology, structures at depth, and other constraints based on seismic, magnetic field, and gravity data. As a result, a variety of different tectonic models have been proposed for the California Coast Ranges and Great Valley forearc basin (Figure 3).

Figure 3 . Montage of tectonic models of the Great Valley forearc basin and Coast Range subduction complex. Rock unit abbreviations: CRF—Coast Range fault; CRO—Jurassic Coast Range ophiolite; EUMW—extended upper mantle wedge; FCX—Cretaceous Franciscan Complex; GVH—Great Valley monocline; GVO—Great Valley ophiolite; GVOM—Great Valley ophiolite mantle; GVS—Upper Jurassic -Cretaceous Great Valley Group; HF—Hayward fault; SA—Sierran arc; SAF—San Andreas fault; SB—Salinian block; SS/UP—stalled slab/underplate; UM—upper mantle [from Constenius et al. 2000 ("This paper" in G refers to Constenius et al. 2000), including figures modified from Ring and Brandon, 1994].

A key point of contention seems to be whether or not tectonic wedging can adequately inform geotectonic interpretations. Tectonic wedging has been a preferred model for interpretations of tectonic relations along the Great Valley flank of the California Coast Ranges. This model was derived from inferred structural relations at Coalinga near the geologic boundary between the California Coast Ranges and the Great Valley. Tectonic wedging has been succinctly described by Davis et al. (1983), and refers to tectonic behavior that is amenable to being modeled as a Coulomb or critical-taper wedge taper, which thins in the direction of motion:

"The overall mechanics of fold-and-thrust belts and accretionary wedges along compressive plate boundaries is considered to be analogous to that of a wedge soil or snow in front of a moving bulldozer. The material within the wedge deforms until a critical taper is attained, after which it slides stably, continuing to grow at constant taper as additional material is encountered at the toe. The critical taper is the shape for which the wedge is on the verge of failure under horizontal compression everywhere, including the basal decollement."

This website summarizes the regional geology, reviews general theories of tectonic wedging, and presents an overview of the history of Coast Range tectonic interpretation, including arguments for and against tectonic wedging in the California Coast Ranges. For a list of references, click here.


1. REGIONAL GEOLOGY

Figure 4 illustrates the position of important rock groups in California, in particular the Coast Range ophiolite, the Franciscan complex, and the Great Valley Group. The differences in these rocks (e.g. density, magnetic properties, and seismic velocities) are important in understanding the subsurface structure, which in turn is necessary for developing a model for past and current tectonic activity in the region.

Figure 4. Geologic map showing main Mesozoic tectonic elements of western California. Rocks of interest are the Coast Range ophiolite, the Franciscan complex, and the Great Valley Group. From Ring and Brandon (1994).

Before we get to the specific characteristics of these rocks, let's summarize the events that led to their location in western California. Much of this summary is presented in Wakabayashi and Unruh (1995).

170–150 Ma:
-Coast Range ophiolite emplaced. This is thought to be induced by mantle upwelling linked to lithospheric extension, but the origin of the ophiolite is controversial. Dickinson and others (1996) provide a summary of current theories on ophiolite emplacement based on various lithospheric spreading hypotheses.

165–130 Ma:
-East-dipping Franciscan subduction began at around 165 Ma, resulting in accretion and metamorphism of Franciscan complex terranes, including blueschist facies (Wakabayashi and Unruh 1995).
-Deposition of the Great Valley Group on oceanic crust of the Coast Range ophiolite also began during this time period, coincident with the early stages of Franciscan subduction (Dickinson and Seely 1979, as cited in Wakabayashi and Unruh 1995).
-A basal Great Valley sequence of late Jurassic age was deposited on the Coast Range ophiolite (Bailey, Blake, and Jones 1970, as cited in Wentworth et al. 1984.), although in most places the contact is faulted.

130–120 Ma:
-Contractional deformation associated with east-vergent thrust faulting in the Great Valley Group began (Unruh et al. 1995). This deformation coincides with the beginning of contraction in the Sevier belt in the Great Basin (Allmendinger 1992, as cited in Wakabayashi and Unruh 1995).

120–100 Ma:
-Contractional deformation ends, and forearc basin sediments are deposited on deformed lower Great Valley Group strata [Unruh et al. 1995, Moxon 1988 (as cited in Wakabayashi and Unruh 1995)].

100–70 Ma:
-Accretion of oceanic terranes in the Franciscan Complex (e.g., Marin Headlands, Geysers) (Murchey and Blake 1993, as cited in Wakabayashi and Unruh 1995). Metamorphism of blueschist facies continues, while some blueschists are uplifted along normal faults (Platt 1986, Jayko et al. 1987).

70 Ma–Present:
-Shortening in the Great Valley Group coincident with accretion of the coastal belt of the Franciscan complex (Moxon 1988, as cited in Wakabayashi and Unruh 1995).
-Onset of tectonic wedging that has continued episodically throughout the Cenozoic (Wakabayashi and Unruh 1995), although this point is debatable.

The properties of rock groups in our region of interest are important factors in characterizing the subsurface structure. Below is a brief summary of the different lithologies present in each group.

Coast Range ophiolite: middle to upper Jurassic oceanic crust: serpentine, gabbro, and basaltic volcanic rocks (Figure 5) (Hopson et al. 1981, as cited in Platt 1986). Ophiolitic breccias locally overlying the Coast Range ophiolite and resting beneath the Great Valley Group reflect extensional deformation at the sites of their construction (Robertson 1990, as cited in Dickinson et al. 1996).

Figure 5. Multiple basaltic sills of the sheeted dike and sill complex, Point Sal remnant of the Middle Jurassic Coast Range ophiolite. The ridge in background exposes sheeted sills and (to left of tree on the skyline) base of the overlying pillow lavas. From Dickinson et al (1996).

Franciscan complex: zonation in age, metamorphic grade (Figure 6), and structural style. Blake and Jones (1981) distinguish three major belts in the northern Californian Coast Ranges:

Figure 6. Metamorphic facies phase diagram.

Great Valley Group: consists primarily of mudstone intercalated with sandstone and conglomerate lenses. Made up of submarine-fan, basin-plain, slop, and submarine-canyon sediments deposited in a north-south-trending forearc basin (Robertson 1990, Ingersoll 1979, 1982, Ingersoll and Dickinson 1981, Brown and Rich 1967, Suchecki 1984, Williams et al. 1998; all as cited in Constenius et al. 2000)


2. TECTONIC WEDGING THEORY

Tectonic wedging has been succinctly described by Davis et al. (1983):

"The overall mechanics of fold-and-thrust belts and accretionary wedges along compressive plate boundaries is considered to be analogous to that of a wedge soil or snow in front of a moving bulldozer. The material within the wedge deforms until a critical taper is attained, after which it slides stably, continuing to grow at constant taper as additional material is encountered at the toe. The critical taper is the shape for which the wedge is on the verge of failure under horizontal compression everywhere, including the basal decollement."

"Tectonic wedging" refers to tectonic behavior that is amenable to being modeled as a Coulomb or critical-taper wedge taper, which thins in the direction of motion (Figure 7). The basic idea is that slight variations in surface slope (alpha) and basal slope (beta) from critical values (based on wedge strength and other factors such as pore pressure) will lead to a tectonic response that brings the system back into a stable state.

Figure 7. Schematic diagram of a critical-taper wedge, from Craig Jones, Fold and Thrust Belts and Orogenic Wedges handout.

The following derivation of a wedge taper expression is from Craig Jones' Fold and Thrust Belts and Orogenic Wedges handout. If we consider the force balance in the x direction on the shaded region of width dx in Figure 7, we can write the expression:

If we assume that the frictional resistance is about equal to the weight times the coefficient of friction, then:

We can approximate the normal stress parallel to x as sigma1 from the Coulomb failure criterion for a rectangular block, adjusting for the angle of z from the vertical:

After applying some simple geometric relationships and using small angle approximations, we can write an expression describing wedge taper, based upon both basal and intrinsic (e.g., pore pressure) parameters:

Platt (1986) conducted stability modeling to predict patterns of deformation in an accretionary wedge resulting from externally imposed changes in its geometry (Figures 8a-d). Frontal accretion refers to the accumulation of material at the tip of the wedge. Because this accreted material lengthens the wedge, alpha will be low in the frontal region, which will therefore be in compression. If the longitudinal stresses are large enough, internal shortening via out-of-sequence thrusting will occur. Underplating is the mechanism by which material is accreted to the underside of the wedge, after it has traveled down the basal decollement some distance.

Figure 8a. Early-stage model of an accretionary wedge, from Platt (1986). Frontal accretion is dominant, and alpha is too low in the frontal region, which therefore shortens internally.

Figure 8b. Alpha is high in the rear of the wedge, which extends by normal faulting and possibly by ductile flow at depth. The deeper parts of the wedge undergo high P/T-ratio metamorphism. From Platt (1986).

Figure 8c. Continued underplating and resultant extension have lifted high-pressure rocks toward the surface. Extension at the rear of the wedge causes lateral movement of material and late thrusting toward the prism front. From Platt (1986).

Figure 8d. In a mature wedge, underplating and extension have brought high-pressure rocks to within 15 km of the surface, accessible to future erosion. The wedge has lengthened and the uplifted rocks in the rear of the wedge will have several generations of accretionary structures overprinted by multistage normal faults. From Platt (1986).


3. PROPOSED TECTONIC MODELS

Below is a brief summary of various studies on tectonic wedging in the Coast Ranges. I have tried to highlight the methods of the research wherever possible; however, many of these papers are heavy on interpretation. Papers are presented in chronological order as an attempt to discern the history of thought on this subject.

Figure 10. Tectonic evolution of the Franciscan Complex, California, from Platt (1986). (a) Early Cretaceous: high pressure/low temperature metamorphism took place in sediment subducted beneath the Coast Range ophiolite and underlying mantle wedge. (b) Early Tertiary: underplating and resultant extension stretched the mantle wedge together with the ophiolite and Great Valley rocks. Note present-day erosion level reflects removal of a significant portion of the upper surface.

Figure 12. Illustration of metamorphic break due to superimposed (a) normal faults or (b) out-of-sequence thrust faults, from Ring and Brandon (1994).

a. b.

Figure 13. (a) Map of contours on top of crystalline basement in central California [from Jachens et al. (1995) after Wentworth et al. (1995)]. Note locations of seismic reflection and refraction profiles, as well as profiles along which magnetic models were constructed. (b) Summary map of basement surface contours, wells, and seismic lines, central California. Figure from Wentworth et al. (1995) to provide location context to (a).

Figure 14. Schematic crustal section across the Great Valley and Coast Ranges of California based on geophysical interpretation by Jachens et al. (1995). Arrows indicate direction of movement of Franciscan wedge relative to adjacent materials. Note that multiple stages of wedging are implied by rocks of the Franciscan Complex structurally both above and below the slab of Coast Range ophiolite located beneath the trace of the Coast Range Fault.

a. b.

Figure 16. (a) Index map showing relationship of Constenius et al. (2000) study area to elements of Great Valley and Coast Ranges. Axis of Great Valley magnetic and gravity high is shown as dashed line. (b) Geologic index map of the northern Great Valley showing seismic profiles and wells. From Constenius et al. (2000).

The authors suggest that imbricate thrust features, typically associated with tectonic wedging, are actually relict constructional patterns formed during deposition (Figure 17). The inactive nature of these features provides support for the lack of tectonic wedging in the Coast Range region.

Figure 17. Patterns of seismic reflectivity from the ophiolitic basement of the Sacramento basin compared with volcanic margins of the Atlantic Ocean basin. Constenius et al. proposed that the Coast Range ophiolite reflections, based on their west-dipping shingled geometry, are the products of interlayered volcanic flows, volcaniclastic rocks, and mafic intrusions similar to those beneath volcanic rifted margins. From Constenius et al. (2000).

Figure 18 is an example of a seismic profile used by Constenius et al. (2000) to assess the geometry of structures at depth. The authors interpret a complicated convergence of seismic reflections to indicate a fork-shaped pattern of reflectivity. This fork structure indicates a combination of syndepositional normal faulting and stratal onlap along the flank of the evolving Great Valley forearc basin, rather than postdepositional thrust wedging.

Figure 18. (e) Migrated time seismic profile TX 5 from the northwestern Sacramento basin that images the Paskenta and Elder Creek fault zones, Cretaceous discontinuity, and ophiolitic basement. (f) Depth-converted section with interpreted units. F indicates location of fork structure. From Constenius et al. (2000).

Figure 19 summarizes the authors' interpretation of subsurface structure, which was confirmed by their Bouguer gravity model. Bouguer gravity modeling was conducted to assess the subsurface continuity of the Coast Range ophiolite and to establish the geometry of the Franciscan Complex beneath the ophiolite. Working within the constraints determined by the seismic reflection profiles, borehole data, and regional geophysical results of Godfrey et al. (1997), the authors experimented with crustal structures and densities that would match the gravity data. Gravity highs occur over layered rocks of the Franciscan assemblage, particularly in areas containing large amounts of mafic volcanic rocks (generally part of an ophiolite belt) or high-pressure metamorphic-mineral facies (Wallace 1990). Most of the deepest lows are caused by thick accumulations of low-density Cenozoic sedimentary rocks (Wallace 1990).

Constenius et al. (2000) conclude that the "tectonic wedging" model is only a viable mechanism if it is restricted to crustal material beneath the Coast Range ophiolite, and that it cannot explain more shallow faulting.

Figure 19. Geologic cross section and Bouguer gravity models of the Great Valley monocline based on regional seismic refraction and gravity results of Godfrey et al. (1997), detailed Bouguer gravity models (Ruppel 1971), interpretation of seismic reflection profile TX 5, borehole data, and surface mapping by Maxwell (1974). F indicates location of fork structure. RC is the generalized reference column for densities of crust and upper mantle. From Constenius et al. (2000).

  • Dickinson (2002) interpreted seismic profiles across the Coalinga anticline (Figure 20) and correlated the character and orientation of seismic reflectors with surface geology. He concluded that a Franciscan thrust wedge was not required by any data, and was apparently incompatible with exposed rocks updip from the seismic reflectors inferred to represent Franciscan rocks.

Figure 20. Location of Coalinga (C) at eastern flank of California Coast Ranges in relation to major tectonic elements of central California. From Dickinson (2002).

Dickinson (2002) contrasts alternate interpretations of seismic data with and without a Franciscan thrust wedge (Figure 21). If the inferred upper surface of the Franciscan wedge (Figure 21a) is projected directly updip, its trace emerges at the surface though a continuous section of the Great Valley Group. To maintain the idea of a Franciscan thrust wedge beneath the core of the Coalinga anticline, it is necessary to postulate the presence of a gently dipping thrust surface at comparatively shallow depth beneath the full 15-km span of the exposed Great Valley Group. No analogous structure has been mapped, suggesting that the thrust wedge model is unreasonable.

Figure 21. Alternate interpretations of USGS seismic line SJ-19: (a) with Franciscan thrust wedge and (b) without Franciscan thrust wedge. Numbers within section indicate mean seismic interval velocities. From Dickinson (2002).

Unruh et al. (2004) disagree with the interpretation that east-dipping reflectors imaged in seismic profiles (Figure 23) represent ophiolitic fragments that were incorporated primarily by tectonic wedging, because thrust-related deformation thickens the crust, whereas both map and subsurface relationships indicate that blueschist-grade rocks of the Franciscan complex are in structural contact with relatively shallow upper crustal rocks, indicating net crustal attenuation (Jayko et al., 1987). Unruh et al. (2004) favor the alternative hypothesis that the magnetic bodies underlying the western outcrops of Great Valley Group strata are fragments of ophiolitic crust and mantle that were mixed with the Franciscan rocks by attenuational processes rather than thrust wedging (Jayko and Blake, 1986; Jayko et al., 1987).

CONCLUSIONS?

Recent studies (Constenius et al. 2000, Dickinson 2002, Ring and Richter 2004) have argued against a tectonic wedging model for late Cenozoic deformation of the Coast Ranges. That these studies use more modern geophysical tools than previous work suggests that the popularity of tectonic wedging may be faltering. It seems that more integration of the literature is needed in order to come to a consensus. Also, although some workers (Wentworth et al. 1984, Constenius et al. 2000, Jachens et al. 1995) present data and interpretation for large areas, a lot of studies are highly localized and therefore make extrapolation to the entire N-S extent of the Coast Ranges and Great Valley highly difficult.


4. REFERENCES

Allmendinger, R. W., 1992, Fold and thrust tectonics of the western United States exclusive of the accreted terranes, in Burchfiel, B. C., et al., eds., The Cordilleran orogen: Conterminous U.S.: Boulder, Colorado, Geological Society of America, Geology of North America, v. G-3, p. 583–607.Bailey, E. H., Blake, M. C. Jr., and Jones, D. L. 1970. On-land Mesozoic oceanic crust in Californian Coast Ranges. USGS Professional Paper 700-C, p. C70-C81.

Blake, M. C., Irwin, W. P., and Coleman, R. B. 1967. Upside-down metamorphic zonation, blueschist facies, along a regional thrust in California and Oregon. USGS Professional Paper 575C, p. 1-9.

Blake, M. C., and Jones, D. L. 1981. The Franciscan assemblage and related rocks in northern California: a reinterpretation. In Ernst, W. G., ed. The geotectonic development of California. Englewood Cliffs, New Jersey, Prentice-Hall, p. 307-328.

Brown, R.D., Jr., and Rich, E.I., 1967, Implications of two Cretaceous mass transport deposits, Sacramento Valley, California: Comment on a paper by Gary L. Peterson: Journal of Sedimentary Petrology, v. 37, p. 240–248.

Cloos, M. 1982. Flow melanges: Numerical modeling and geologic constraints on their origin in the Franciscan subduction complex, California. Geol. Soc. of Amer. Bull. v. 93, p. 330-345.

Constenius, K. N., Johnson, R. A., Dickinson, W. R., and T. A. Williams. 2000. Tectonic evolution of the Jurassic-Cretaceous Great Valley forearc, California: Implications for the Franciscan thrust-wedge hypothesis. GSA Bulletin v. 112, no. 11. p. 1703-1723. November.

Davis, D., Suppe, J., and F. A. Dahlen. 1983. Mechanics of fold-and-thrust belts and accretionary wedges. Journal of Geophysical Research v. 88, no. B2, p. 1153-1172.

Dickinson, W. R. 1981. Plate tectonic evolution of the southern Cordillera. In: Relations of tectonics to ore deposits in the southern Cordillera. Arizona Geological Society Digest, 14: 113-135.

Dickinson, W.R., 2002. Reappraisal of hypothetical thrust wedging at Coalinga: implications for tectonic relations along the Great Valley flank of the California Coast Ranges. Tectonics 21, 3-1– 3-14.

Dickinson, W. R., Hopson, C. A., and J. B. Saleeby. 1996. Alternate origins of the Coast Range ophiolite (California): introduction and implications. GSA Today v. 6, no.2. p. 1-10. February 1996.

Dickinson, W. R. and Seely, D. R. 1979. Structure and stratigraphy of forearc regions. American Association of Petroleum Geologists Bulletin, v. 63, p. 2-31.

Ernst, W. G. 1971. Metamorphic zonations on presumably subducted lithospheric slabs from Japan, California, and the Alps. Contributions to Mineralogy and Petrology, v. 34, p. 43-59.

Fuis, G. S. and W. D. Mooney. 1990. Lithospheric structure and tectonics from seismic-refraction and other data. In: Wallace, R. E. 1990. The San Andreas fault system, California. U.S. Geological Survey Professional Paper: 1515. Online at: http://education.usgs.gov/california/pp1515/copyright.html. Viewed December 2005.

Godfrey, N. J., B. C. Beaudoin, and S. L. Klemperer. 1997. Ophiolitic basement to the Great Valley forearc basin, California, from seismic and gravity data: Implications for crustal growth at the North American continental margin, Geol. Soc. Am. Bull., 109, 1536 – 1562.

Griscom, A. and R. C. Jachens. 1990. Tectonic implications of gravity and magnetic models along east-west seismic profiles along the Great Valley near Coalinga. In: The Coalinga, California Earthquake of May 2, 1983. Ed: M. J. Rymer and W. L. Ellsworth. USGS Professional Paper 1487, p. 69-78.

Hopson, C. A., Mattinson, J. M., and Pessagno, E. A. 1981. Coast Range ophiolite, western California. In Ernst, W. G., ed., The geotectonic development of California. Englewood Cliffs, New Jersey, Prentice-Hall, p. 418-510.

Ingersoll, R.V., 1979, Evolution of the Late Cretaceous forearc basin, northern and central California: Geological Society of America Bulletin, v. 90, p. 813–826.

Ingersoll, R.V., 1982, Initiation and evolution of the Great Valley forearc basin of northern and central California, in Leggett, J.K., ed., Trench-forearc geology: Sedimentation
and tectonics on modern and ancient active plate margins: Geological Society [London] Special Publication 10, p. 459–467.

Ingersoll, R.V., and Dickinson, W.R., 1981, Great Valley Group (sequence), Sacramento Valley, California, in Frizzell, V., ed., Upper Mesozoic Franciscan rocks and Great Valley sequence, central Coast Ranges, California: Annual Meeting Field Trips 1 and 4: Pacific Section, Society of Economic Paleontologists and Mineralogists, p. 1–33.

Jachens, R. C., Griscom, A., and C. W. Roberts. 1995. Regional extent of Great Valley basement west of the Great Valley, California: Implications for extensive tectonic wedging in the California Coast Ranges. Journal of Geophysical Research v. 100, no. B7, p. 12,769-12,790. July 10, 1995.

Jayko, A.S., Blake Jr., M.C. 1986. Significance of Klamath rocks between the Franciscan complex and Coast Range ophiolite, northern California. Tectonics 5, 1055– 1071.

Jayko, A.S., Blake, M.C., Jr., and Harms, T. 1987. Attenuation of the Coast Range Ophiolite by extensional faulting, and the nature of the Coast Range ‘‘thrust,’’ California: Tectonics, v. 6, p. 475–488.

Maxwell, J. C. 1974. Anatomy of an orogen. Geological Society of America Bulletin, v. 85, p. 1195-1204.

McLaughlin, R. J., Kling, S. A., Poore, R. Z., McDougall, K., and Beutner, E. C. 1982. Post-middle Miocene accretion of Franciscan rocks, northwest California. Geological Society of America Bulletin v. 93, p. 595-605.

Murchey, B. L., and Blake, M. C., Jr., 1993, Evidence for subduction of a major ocean plate along the California margin during the Middle to early Late Jurassic, in Dunne, G., and McDougall, K., eds., Mesozoic paleogeography of the western United States, II: Pacific Section, Society of Economic Paleontologists and Mineralogists, Book 71, p. 1–18.

Moxon, I. W., 1988, Sequence stratigraphy of the Great Valley basin in the context of convergent margin tectonics, in Graham, S. A., ed., Studies of the geology of the San Joaquin Basin: Pacific Section, Society of Economic Paleontologists and Mineralogists, Field Trip Guidebook 60, p. 3–28.

Platt, J. P. 1986. Dynamics of orogenic wedges and the uplift of high-pressure metamorphic rocks. Geol. Soc. Am. Bull., 97, 1037 – 1053.

Ring, U. and M. T. Brandon. 1994. Kinematic data for the Coast Range fault and implications for exhumation of the Franciscan subduction complex. Geology v. 22, p. 735-738.

Ring, U. and P. P. Richter. 2004. Normal faulting at convergent plate boundaries: Mylonitic extensional fabrics in the Franciscan subduction complex in Del Puerto Canyon, California, revisited. Tectonics, Vol. 23,

Robertson, A. H. F. 1990. Sedimentology and tectonic implications of ophiolite-derived clastics overlying the Jurassic Coast Range Ophiolite, northern California. American Journal of Science v. 290, p. 109-163.

Ruppel, B.D. 1971. Geophysics of the Great Valley–Franciscan junction between Paskenta and Stony Ford, California: Geological Society of America Bulletin, v. 82, p. 1717–1722.

Suchecki, R.K., 1984, Facies history of the Upper Jurassic– Lower Cretaceous Great Valley sequence: Response to structural development of an outer-arc basin: Journal of Sedimentary Petrology, v. 54, p. 170–191.

Suppe, J. 1973. Geology of the Leech Lake Mountain-Ball Mountain region, California: A cross-section of the northeastern Franciscan belt and its tectonic implications. University of California Publications in Geological Sciences, v. 107, 82 p.

Suppe, J. 1979. Structural interpretation of the southern part of the northern Coast Ranges and Sacramento Valley, California: Summary. Geological Society of American Bulletin, v. 90, p. 327-330.

Unruh, J., O'Connell, D., and L. V. Block. 2004. Crustal structure of the ancestral northwestern California forearc region from seismic reflection imaging: implications for convergent margin tectonics. Tectonophysics 392, p. 219-240.

Unruh, J. R., B. A. Loewen, and E. M. Moores. 1995. Progressive arcward contraction of a Mesozoic-Tertiary fore-arc basin, southwestern Sacramento Valley, California, Geol. Soc. Am. Bull., 107, 38 – 53.

Wakabayashi, J. and Unruh J. R. 1995. Tectonic wedging, blueschist metamorphism, and exposure of blueschists: Are they compatible? Geology 23 (1): 85-88 Jan 1995.

Wallace, R. E. 1990. The San Andreas fault system, California. U.S. Geological Survey Professional Paper: 1515. Online at: http://education.usgs.gov/california/pp1515/copyright.html. Viewed December 2005.

Walter, A. W. 1990. Upper-crustal velocity structure near Coalinga, as determined from seismic-refraction data. In: The Coalinga, California Earthquake of May 2, 1983. Ed: M. J. Rymer and W. L. Ellsworth. USGS Professional Paper 1487, p. 23-40.

Wentworth, C. M., Blake, M. C. Jr., Jones, D. L., Walter, A. W., and Zoback, M. D. 1984. Tectonic wedging associated with emplacement of the Franciscan assemblage, California Coast Ranges. In Blake, M.C., ed., Franciscan geology of northern California. Pacific Section, Society of Economic Paleontologists and Mineralogists, Field Trip Guidebook 43, p. 163-173.

Wentworth, C. M., Fisher, G. R., Levine, P. and R. C. Jachens. 1995. The surface of crystalline basement, Great Valley and Sierra Nevada, California: a digital map database. USGS Open File Report 95-96.

Wentworth, C. M., Walter, A. W., Zoback, M. D., and Blake, M. C., Jr. 1983. Possible obduction of Franciscan assemblage, southeasternmost Diablo Range, California (abs.). American Geophysical Union, Transactions, EOS v. 64, no. 45, p. 868.

Wentworth, C. M. and M. D. Zoback. 1990. Structure of the Coalinga region and thrust origin of the earthquake. In: The Coalinga, California Earthquake of May 2, 1983. Ed: M. J. Rymer and W. L. Ellsworth. USGS Professional Paper 1487, p. 41-68.

Williams, T.A., Graham, S.A., and Constenius, K.N. 1998. Recognition of a Santonian submarine canyon, Great Valley Group, Sacramento basin, California: Implications for petroleum exploration and sequence stratigraphy of deep-marine strata: American Association of Petroleum Geologists Bulletin, v. 82, p. 1575–1595.