You are on page 1of 12

Click Here

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112, F03014, doi:10.1029/2005JF000433, 2007

Full Article

for

Tectonics, fracturing of rock, and erosion


Peter Molnar,1,2 Robert S. Anderson,1,3 and Suzanne Prestrud Anderson3,4
Received 2 November 2005; revised 23 March 2007; accepted 30 April 2007; published 9 August 2007.

[1] We argue that by fracturing rock, not by raising it relative to base level, tectonics plays

its most important role in causing rapid incision of valleys and rapid erosion of hillslopes. Tectonic deformation riddles the upper crust with fractures, which not only provide avenues for water flow and thus promote weathering and further disintegration of rock but also fragment bedrock into debris that is readily extracted and transported by surface processes. Bends in active faults require straining of adjacent rock masses. Aftershocks that occur subsequent to slip on primary faults reflect penetrative brittle deformation of the upper crust. At least some aftershocks must nucleate or lengthen cracks, which contribute to the comminution of these rock masses. Scaling rules suggest that dimensions of ruptures for very small (M < 2) earthquakes can be meters or less. The Gutenberg-Richter recurrence relationship implies that such earthquakes are common, as high-magnification seismographs in low-noise environments confirm. Moreover, large differences among fault plane solutions for aftershocks show that the small faults on which they occur are not parallel to one another; some faults must intersect. Thus the upper crust in tectonically active regions should be fragmented into blocks down to the scale of boulders or smaller. Dismembered rock arrives at the Earths surface already prepared to be transported away. As a corollary, both deeply exhumed lower crust and posttectonic igneous rock, never deformed under brittle conditions and not deformed recently, should be less susceptible to detachment and subsequent transport than fractured rock.
Citation: Molnar, P., R. S. Anderson, and S. P. Anderson (2007), Tectonics, fracturing of rock, and erosion, J. Geophys. Res., 112, F03014, doi:10.1029/2005JF000433.
All indurated rocks and most earths are bound together by a force of cohesion which must be overcome before they can be divided and removed. The natural processes by which the division and removal are accomplished make up erosion. They are called disintegration and transportation. [Gilbert, 1877, pp. 93 94] Erosion is the result of two complex process. The first group comprises those which accomplish the disintegration of the rocks, reducing them to fragments, pebbles, sand, and clay. The second comprises those process which remove the debris and carry it away to another part of the world. [Dutton, 1882, p. 64]

1. Introduction
[2] In the past 15 years, following the synthesis by Dahlen and Suppe [1988] and then accelerating with work of Koons [1989, 1990], Beaumont et al. [1992], and others, geomorphologists and structural geologists have focused
1 Department of Geological Sciences, University of Colorado, Boulder, Colorado, USA. 2 Cooperative Institute for Research in Environmental Science, University of Colorado, Boulder, Colorado, USA. 3 Institute of Arctic and Alpine Research, University of Colorado, Boulder, Colorado, USA. 4 Department of Geography, University of Colorado, Boulder, Colorado, USA.

Copyright 2007 by the American Geophysical Union. 0148-0227/07/2005JF000433$09.00

attention on the interaction between tectonics and geomorphology, instead of treating each in isolation of the other. Removal of material from the Earths surface by erosion should reduce magnitudes of vertical compressive stress, and as a result horizontal deviatoric stresses should increase in regions undergoing horizontal shortening, even with no change in magnitudes of horizontal compressive stresses across the belts. (To remind readers, deviatoric stress, t ij, is stress, sij, minus pressure, p, if we treat compressive stress as positive: t ij = sij pd ii. Stress does physical work; deviatoric stress causes deformation.) Thus erosion facilitates further crustal shortening [e.g., Dahlen and Suppe, 1988]. Moreover, because of isostasy, as erosion removes rock from the surface, exhumed rock rises to a mean elevation nearly the same as that of the rock eroded, a fact recognized by Wager [1937] and probably others, and included in elementary textbooks [e.g., Holmes, 1944, pp. 189 190, 1965, pp. 575 576]. Thus a belt of high terrain, a mountain belt, can be sustained simply by erosion and isostasy for a much longer period than the division of present-day elevations by current erosion rates might suggest [e.g., Sleep, 1971]. [3] Many studies of the interaction between geomorphology and tectonics continue to assign tectonics a causal role, with tectonics seen solely as the source of relief, and erosion its consequence. Our purpose here is to argue that tectonics may play a more important role in accelerating erosion rates by fracturing rock than by creating large-scale relief. Frac1 of 12

F03014

F03014

MOLNAR ET AL.: TECTONICS, FRACTURING, AND EROSION

F03014

Figure 1. Drilling rates as a function of the spacing of fractures in limestone (Muschelkalk), modified from Figure 17 of Thuro [1997]. The % scale on the right shows the enhancement in drilling rate with decreased spacing of fractures. Thuro did not report the dimensions of the drill bit used to carry out the tests, but we suppose it to be a few centimeters in diameter. The light blue boxes presumably show the scatter in data, the horizontal black lines the mean values, and dashed red line an empirical fit. Where the spacing of the fractures is large compared to the dimensions of the drill bit, the presence of fractures does not weaken the rock attacked by the drill bit. turing can both reduce rock strength, which will lead to greater erosion rates (all else equal), and broaden the range of geomorphic processes that can operate (e.g., mass loss through pervasive weathering of fractured rock, glacial or fluvial quarrying, or deep-seated landslides along fracture planes). Perhaps more important than reducing the strength of the rock, however, may be the disaggregation of bedrock into blocks that are easily extracted and transported by geomorphic processes. Although we claim nothing original in suggesting either that tectonics fractures rock or that fracturing of rock facilitates its removal, these concepts do seem poorly appreciated. [4] The fracturing of rock can affect geomorphic processes in two obvious ways, but most discussions seem to focus on only one. First, because jointed rock with cracks penetrating through it is weaker than intact rock [e.g., Pettifer and Fookes, 1994; Segall, 1984; Selby, 1980], rivers, debris flows, glaciers, or simply gravity acting directly on hillslopes should be able to dislodge fragments of it more easily than when these erosive agents are confronted with intact bedrock. The preponderance of slopes near the angle of repose suggests that only in special circumstances is rock exposed at the Earths surface strong enough to produce tall, near vertical walls of rock like those of El Capitan in Yosemite Valley, the North Face of the Eiger in the Swiss Alps, or the south face Dhaulagiri in Nepal [e.g., Burbank et al., 1996; Schmidt and Montgomery, 1995; Selby, 1980]. [5] The vast majority of sediment carried by rivers, and a substantial part of that carried by glaciers, is derived from adjacent hillslopes. Yet, if rivers and glaciers are to incise valleys, they must remove bedrock, and the degree to which the bedrock has been fragmented into transportable debris (sediment) before rushing water or sliding ice come in contact with it, the more rapidly will rivers and glaciers incise their valley floors. Accordingly, the second role of

rock fracture is to provide debris that can be plucked or quarried by rivers and glaciers from valley floors without requiring that rivers, debris flows, or glaciers perform the work of fracturing the rock. [6] Weak rock will certainly erode more quickly than strong rock, as shown qualitatively by river profiles [e.g., Hack, 1957, 1973] and quantitatively by laboratory experiments of abrasion using pebbles and cobbles in rapidly flowing water [e.g., Attal and Lave, 2006; Sklar and Dietrich, 2001, 2004], by applications of these experiments to field settings [Stock et al., 2005], and by field experiments in which samples are glued to bedrock beneath glaciers [Boulton and Vivian, 1973]. In addition, most would assume that fractured rock is more easily eroded than intact rock and that fracture density plays a role in setting erodibility. In his strength classification, Selby [1980] found fracture spacing to be the feature most important in limiting hillslopes, more important than intrinsic strength of intact rock. Yet, insofar as tests made with drilling [e.g., Thuro, 1997] or with dredging of rock from the seafloor [e.g., Vervoort and De Wit, 1997] are representative, where the spacing of fractures exceeds the dimensions of tools attacking the rock, the existence of fractures appears to play no more than a minor role in weakening the rock. For instance, Thuro [1997, p. 433] showed that where the spacing of joints or fractures is large compared with the dimensions of a drill bit, these discontinuities do not affect the drilling rate (Figure 1). Similarly, the excavation of rock with large fracture spacing remains difficult, either until fracture spacing is small enough or where the intrinsic strength of the material, measured by the point load index, is low [Pettifer and Fookes, 1994]. Because plucking or quarrying of blocks from valley floors, by both rivers [e.g., Hancock et al., 1998] and glaciers [e.g., Briner and Swanson, 1998], is thought to be more effective than abrasion, this second role of fracturing as a geomorphic agent, by transforming intact bedrock into separate fragments of debris that can be efficiently transported by rivers, debris flows, or glaciers, deserves more attention than it has received. [7] It seems to us that these two effects of fracturing are often conflated. Some modern treatments of fluvial erosion [e.g., Sklar and Dietrich, 2004], erosion by debris flows [e.g., Stock and Dietrich, 2006; Stock et al., 2005], and glacial erosion [e.g., Lliboutry, 1994] treat the disintegration and initial transportation as occurring simultaneously, as of course they often, but not always, do. Others seem to treat fractures as unchanging and uniform conditions, tied either to the rocks or to the surface, and whose formation remains unaddressed. We see tectonics as playing a key role in controlling the condition of the rock upon arrival at the surface by dictating the strain history to which a parcel of rock has been subjected in the brittle zone. Here, we follow the view expressed by Dutton [1882] and Gilbert [1877], and quoted above: that the two processes of disintegration and transportation can occur separately.

2. Tectonics, Mountain Building, and Fracturing of Rock


[8] Tectonically active belts like the Himalaya, Taiwan, or the Southern Alps of New Zealand and some recently active belts like the Western Alps of Europe seem to be

2 of 12

F03014

MOLNAR ET AL.: TECTONICS, FRACTURING, AND EROSION

F03014

Figure 2. Cartoon illustrating strain within a layer of crust that bends where a fault is not planar. One of the roles of tectonics is the generation of faults (cracks) within rock that is ultimately delivered to the surface of the Earth. In the example shown here, rock in the hanging wall moves laterally to the left, then through a region where the layer is bent, and finally upward on the dipping segment of the fault. Cracks, shown as red lines, form to accommodate strain associated with bending where the fault is curved; this bending serves as a tectonic, crustal-scale rock crusher. Crack lengths and densities increase with time spent in the bend. The resulting fracture field, with inactive cracks shown in blue, is then translated to the surface, where the fractures influence surface processes. eroding more quickly than tectonically quiescent regions. At the same time, some elevated regions with steep slopes and deep canyons, like the Guyana Shield [Brown et al., 1992; Stallard, 1988], the Western Ghats along the west coast of India [Gunnell et al., 2003], or Sri Lanka [Vanacker et al., 2007; von Blanckenburg et al., 2004], do not erode rapidly, despite high rainfall and high relief in these regions. Few would doubt that tectonic deformation generates rock fractures, and some have explicitly noted that this process should affect erosion rates [e.g., Koons and Craw, 2002]. For instance, in contrasting rapid erosion of the Andes and slow erosion of comparably high terrain of the Guyana Shield, Stallard [1988, p. 225] wrote: In tectonically active environments, most of the material exposed to weathering has undergone rapid tectonic uplift involving brittle deformation, where surely brittle deformation means fracturing of rock (and we ignore his comparable emphasis on rapid tectonic uplift). Weissel and Seidl [1997] showed how not only the presence of joints, but also their orientations affected river incision and hillslope evolution in escarpment retreat. Kellogg [2001] mapped both fresh and early Cenozoic fractures in the Front Range of Colorado and recognized that these fractures have facilitated the current, relatively rapid erosion. Finally, Loveless et al. [2005] demonstrated pervasive tectonic cracking at the surface in northern Chile, where erosion rates are so low and soil is virtually absent that the cracks can be seen on the surface. Nevertheless, despite several studies that specifically mention tectonics and fracturing, we suspect that many do not appreciate the degree to which such fracturing must occur, or the variety of arguments that show it to be noteworthy. Although most readers recognize that the exposed Earths surface in many regions is fractured, in many regions, such

as in the Sierra Nevada, Wind River Range, Wyoming, and Western Ghats of India, fractures are sparse, tens of meters apart. Accordingly, we develop the argument that tectonic deformation, commonly associated with slip on one or just a few major faults, facilitates erosion by fracturing rock adjacent to major faults. [9] Below we argue first that the shapes of major faults, which are rarely planar, require additional deformation of the adjacent rock masses for slip to occur on them. Second, seismicity, at least within continental regions, also suggests that tectonic movements require the fracturing of rock, not just the transport of large blocks with respect to one another. (We do not argue that each earthquake reflects the creation or extension of a new fault; few earthquakes do this.) Finally, treatments not only of earthquakes, but also of faulting in general, suggest scale invariance, and hence justify extrapolations to small dimensions of faulting and of the blocks that result from such faulting. 2.1. Strain Within Deforming Belts Caused by Slip on Nonplanar Faults [10] Deformation in all major thrust belts should include large strain, simply because dip-slip faults that approach the surface cannot be planar [e.g., Suppe, 1983, 1985]. At least one of the hanging wall and footwall must deform adjacent to faults that are not planar, and where such strain occurs in the upper crust, brittle deformation should prevail. [11] The straining of blocks of crust that slide past one another on faults that are not planar has been widely noted and studied in hanging walls of thrust faults. Suppes [1983, 1985] method for balancing cross sections, for instance, approximates fault surfaces as connected planar segments. This approach predicts that strain will be localized near regions where two planar segments intersect (Figure 2) and can be manifest as minor faults that rupture the surface [e.g., Ishiyama et al., 2004]. In a similar situation but above a concave upward major fault zone, pervasive reverse faulting can slice rock, as for example in Ventura Basin [Shaw and Suppe, 1994] and on the eastern flank of the Southern Alps in New Zealand [e.g., Little et al., 2002]. Moreover, detailed mapping in the Southern Alps reveals highly fractured rock [Cox et al., 1997], as required by the subsurface faulting on segments that are not parallel to one another, proposed by Little et al. [2002]. [12] As a specific example, consider the Himalaya, where a major thrust fault dips gently beneath the Lesser Himalaya (with dips of 3 6) and more steeply beneath at least part of the Greater Himalaya. Using receiver functions determined with an array of broadband seismographs in the Himalaya and southern Tibet, Schulte-Pelkum et al. [2005] reported a dip of 12 15 beneath the Greater Himalaya. The large flexural rigidity of the Indian plate implies that it cannot be bent sharply and that it slides beneath the Himalaya as a nearly rigid block. Thus the $6 9 difference in the dips of the segments of the thrust fault that underlie the Lesser and Greater Himalaya suggests that the hanging wall must deform (Figure 3), and as India converges with the Himalaya, the vertical components of velocity above these segments of fault must also differ. The rock in the Greater Himalaya must rise more rapidly than rock in the Lesser Himalaya by an amount approximately equal to the rate of underthrusting (v) times the sine of the

3 of 12

F03014

MOLNAR ET AL.: TECTONICS, FRACTURING, AND EROSION

F03014

Figure 3. (bottom) Idealized cross section across the Himalaya showing a sharply curved thrust fault and (top) sketch of the expected vertical component of movement of the rock above it. The Indian lithosphere flexes down beneath the Ganga Basin, where sediment accumulates, and underthrusts the Lesser Himalaya. Beneath the Greater Himalaya, the fault dips more steeply than beneath the lesser Himalaya. Thus rock in the hanging wall over the steeper segment is ramped upward more rapidly than above the gently dipping segment.

difference in dips: v sin 6 = 0.1 v. For underthrusting at $20 mm/yr, this implies a difference in vertical displacement rate of 2 mm/yr. Judging by the width of the zone where earthquakes occur [e.g., Pandey et al., 1995], which is comparable with the width of the transition from low elevations of the Lesser Himalaya and high elevations of the Greater Himalaya, this difference in vertical movement occurs over a zone no more than a few tens of kilometers wide, say 20 km. For underthrusting at 20 mm/yr, shearing of the hanging wall on vertical planes would be accomplished in 1 Myr, and the total accumulated strain would be 0.1(=2 km/20 km). As most rock cannot sustain strains of 0.001 without breaking, strains of 0.1 require that the hanging wall be fractured. Moreover, if this shear occurs on planes that dip not vertically but less steeply, shear on those planes must be more rapid [e.g., Molnar, 1987]. [13] Wobus et al. [2005] showed that in this part of the Himalaya in Nepal, faulting occurs with a vertical component of slip rate of 0.6 0.2 mm/yr, less than half and perhaps only 20% of the total rate of $ 2 mm/yr suggested above. Thus slip on the fault recognized by Wobus et al. [2005] accounts for only part of the strain that must accumulate in this region. Moreover, Burbank [2005] pointed out that the fault discovered by Wobus et al. [2005] does not manifest itself clearly in the relief in the Annapurna region a few tens of kilometers to the west. Apparently deformation, which must accommodate $2 mm/yr of relative movement, is yet more diffuse in the region to the west. The occurrence of brittle strain of $0.1 does not require that fractures be closely spaced, and hence does not guarantee that fractures will slice and dice the rock into fragments easily transported by rivers. The diffuse zone of accurately located microearthquakes, however, does imply pervasive straining of the hanging wall near the ramp [Pandey et al., 1995].

2.2. Seismicity and Fracturing [14] Two aspects of seismicity are relevant to our argument. First, many of the earthquakes that occur in the immediate aftermath of major earthquakes (aftershocks) occur off the main fault or faults that rupture in the main shock [e.g., King et al., 1985; Yielding et al., 1981]. Second, the size distribution of small earthquakes suggests a distribution of dimensions of faults in the affected rock. [15] 1. Locations of aftershocks indicate diffuse brittle deformation in areas adjacent to major faults. Although aftershocks of great earthquakes commonly are used to define the rupture areas of such earthquakes [e.g., Das and Henry, 2003; Fedotov, 1965; Kelleher et al., 1973; Sykes, 1971], many, and in some cases most, aftershocks of intracontinental earthquakes do not occur on the fault that ruptured during the main shock. Accordingly, these aftershocks do not register continued slip on major faults. Instead they reflect deformation within the blocks of crust that moved past one another during the main shocks. This is the norm for earthquakes in Greece, where aftershocks commonly occur in both the hanging and footwalls of the causative normal faults [e.g., Hatzfeld et al., 1986, 1997, 2000; King et al., 1985; Soufleris et al., 1982]. For instance, aftershocks of the Kozani-Grevena earthquake in northern Greece on 13 May 1995 (Figure 4) require slip on conjugate faults, as well as activity not on these faults [Hatzfeld et al., 1997]. Recent relocations, with reported uncertainties of only 150 m in all three coordinates and coupled with surface deformation based on radar interferometry, demonstrate that slip occurred on at least three intersecting faults [Resor et al., 2005]. Moreover, the scatter in these locations requires internal deformation of both the footwall and hanging wall (Figure 4). [16] Similarly for thrust regimes, like the 1981 El-Asnam earthquake in Algeria [e.g., Ouyed et al., 1983; Yielding et al., 1981] (Figure 5), the 1994 Northridge earthquake in southern California [Shearer et al., 2003], the 2000 Bhuj earthquake, Gujarat, India [Bodin and Horton, 2004; Negishi et al., 2002], or the 2003 San Simeon earthquake in central California [Hardebeck et al., 2004], aftershocks occurred throughout the hanging walls and in parts of the footwalls of the thrust faults. By correlating waveforms,

Figure 4. Northwest-southeast cross section of aftershocks of the Kozani-Grevena (Greece) earthquake of 13 May 1995, whose locations were determined by Hatzfeld et al. [1997]. Note that the distribution of earthquakes does not permit them to lie on a single listric, let alone planar, fault.

4 of 12

F03014

MOLNAR ET AL.: TECTONICS, FRACTURING, AND EROSION

F03014

al., 1970] occur on the fault segments that ruptured in the main shocks, but those following moderate earthquakes, which rupture short faults slipping at low slip rates, commonly do not occur on the faults that ruptured in the main shocks. For instance, Stein and Lisowski [1983] showed that for the relatively small Homestead Valley earthquakes in 1979, which ruptured a vertical strike-slip fault $4 6 km long that extended to a depth of 5 km in southern California, aftershocks occurred many kilometers east and west of the north-south trending rupture, over an area 15 km 25 km in dimensions and hence also east and west of the surface

Figure 5. (a) Map of aftershocks relocated by Ouyed et al. [1983] and (b) example of a cross section through the seismicity whose location, A-B, is shown by the red line in Figure 5a. In both plots, solid symbols denote earthquakes with the most accurate locations: those for which rootmean-square residuals were less than 0.15 s, locations were assigned horizontal and vertical uncertainties of less than 2 km, and at least one station lay within an epicentral distance less than twice the computed focal depth. Open symbols denote less precisely located events. Symbol sizes scale with magnitudes. Again, note that the distribution of earthquakes requires than many faults be active during the aftershock sequence. Shearer et al. [2003] showed not only that numerous short faults with different orientations ruptured following the Northridge earthquake, but also that the styles of faulting in the hanging wall and footwall differ. [17] Most aftershocks of major earthquakes on major strike-slip faults like the San Andreas fault [e.g., Eaton et

Figure 6. Map of aftershocks of the 1979 Homestead Valley earthquake in southern California, located by Stein and Lisowski [1983], and contours of Coulomb stress change from King et al. [1994]. The plot is from King et al. [1994], who used the aftershock locations to show that they occurred largely in regions where failure was facilitated by the change in Coulomb stress (the combination of increased shear stress on planes oriented favorably and decreased normal stress across such planes). Although the main shock ruptured a plane trending nearly north-south, aftershocks occurred throughout a large region far from the main rupture and hence require deformation within a finite volume, not just slip on a single plane.

5 of 12

F03014

MOLNAR ET AL.: TECTONICS, FRACTURING, AND EROSION

F03014

normal in place of thrust faulting (Figure 7b). Thus the wide variety of orientations of faults that slipped implies that some faults must intercept others, as Hatzfeld et al. [1997] and Resor et al. [2005] showed well for the Kozani-Grevena earthquake. In regions pervaded by aftershocks, we should expect the rock to be broken into blocks (whose dimensions we discuss below). [20] 2. Rupture areas of faulting in small earthquakes imply pervasive fragmentation. Enough is known of typical dimensions of ruptures in earthquakes with different magnitudes to extrapolate this relationship to estimate rupture dimensions for tiny earthquakes. Among the quantities that characterize an earthquake source, the one most easily and accurately measured is the seismic moment: Figure 7. Lower hemisphere stereographic projection of P and T axes for (a) the El Asnam aftershocks, determined by Ouyed et al. [1983], and (b) the Kozani-Grevena earthquake, determined by Hatzfeld et al. [1997]. Red and blue triangles show mean orientations of P and T axes, respectively. The scatter both of T axes for the El Asnam aftershocks and of P axes for the Kozani-Grevena earthquake require that for both sequences, faults with different orientations were activated. Therefore some of these faults must intercept others. rupture (Figure 6). Similar phenomena occurred following the 1992 Joshua Tree earthquake (M = 6.1) [Hauksson et al., 1993; King et al., 1994] and the 1994 Arthurs Pass earthquake in New Zealand [e.g., Abercrombie et al., 2000, 2001; Robinson and McGinty, 2000]. [18] In addition, in tectonically active regions, large perturbations to the tectonic stress field by nuclear explosions have triggered widespread aftershock activity [e.g., Hamilton and Healy, 1969; Hamilton et al., 1972] and the creation of new faults [McKeown and Dickey, 1969]. Regardless of whether dynamic or static changes in stress triggered such aftershocks, because the majority of aftershocks occurred far from the region where the explosions themselves created large permanent strain, only tectonic stresses could have caused slip on both preexisting and newly created faults. [19] Fault plane solutions of aftershocks commonly reveal a coherent average regional straining, but by no means do all such solutions resemble that of the main shock or each other [e.g., Bodin and Horton, 2004; Hamilton and Healy, 1969; Hamilton et al., 1972; Hardebeck et al., 2004; Hatzfeld et al., 1986, 1997, 2000; Ouyed et al., 1983; Shearer et al., 2003; Soufleris et al., 1982]. Such solutions commonly share a consistent P or T axis, axes of principal compressive or extensile strains (Figure 7), but faulting commonly occurs on planes with a variety of orientations. For instance, the consistent NW-SE orientation of P axes for the El Asnam earthquake [Ouyed et al., 1983], with T axes forming a girdle around the NW-SE orientation of maximum compressive strain, implies that a mixture of pure thrust, pure strike-slip, and oblique-slip faulting accommodated strain in the region of aftershocks (Figure 7a). The similar pattern, but with a consistent NNW-SSE orientation of T axes for the Kozani-Grevena earthquake [Hatzfeld et al., 1997], shows a similar variety of solutions, but with
MO mADu 1

where m is the shear modulus, A is the area that ruptured, and Du is the average slip on the fault during the earthquake [Aki, 1966]. Independent measurements of the rupture area, and, particularly for small events, of the average displacement are much harder to make. For earthquakes that rupture the Earths surface, surface faulting places a bound on one of the dimensions of rupture. Where aftershocks define clearly the fault that ruptured, their maximum depths, or that of background seismicity, constrain the other dimension of a rupture. For smaller earthquakes, the spectral content of the body waves from the earthquakes allow estimates of both seismic moment and source dimensions to be made.

Figure 8. Seismic moments versus estimated radii of equivalent circular ruptures for earthquakes spanning 9 orders of magnitude in seismic moment. Lines show relationships between radius and moment for constant stress drops. We replotted the data used by Hanks [1977]. The common relationship that stress drops lie within 1 order of magnitude of 0.1 MPa suggests that small and large earthquakes differ from one another only in dimensions and that scaling relationships can be extrapolated to small events.

6 of 12

F03014

MOLNAR ET AL.: TECTONICS, FRACTURING, AND EROSION

F03014

[21] Using estimates of rupture dimensions and seismic moments for earthquakes with magnitudes greater than $3, Hanks [1977] showed their logarithms plot on a straight line, implying a simple power law relationship (Figure 8). Subsequent analyses of similar data corroborate the pattern in Figure 8 and extend it to magnitudes as small as 1 [Abercrombie, 1995; McGarr et al., 1981; Spottiswoode and McGarr, 1975]. This scaling implies that the stress drop, the average decrease in shear stress across the fault that slips in an earthquake, typically ranges between 0.1 and 10 MPa, with an approximately lognormal distribution and a geometric average near 1 MPa. Exceptions do exist, for which source dimensions appear to be atypically small for the magnitudes and moments of the earthquakes, and stress drops are larger than 10 MPa; for such events, fracturing of intact rock seems inescapable [e.g., Nadeau and Johnson, 1994; Nadeau et al., 1994], as clearly occurs locally with some earthquakes due to excavations in mines [e.g., McGarr et al., 1975; Yamada et al., 2007]. Nevertheless, Ide and Beroza [2001] argued that if corrections were made to account for limited bandwidths of recording, the energy radiated by most earthquakes as seismic waves is proportional to the seismic moment over 17 orders of magnitude of moment; hence stress drops are indeed approximately constant for earthquakes with magnitudes smaller than 3 to larger than 8. As we show next, knowledge of both the stress drop and the seismic moment for smaller earthquakes allows an estimate of the dimensions of ruptures in such earthquakes. [22] As slip occurs during an earthquake, both the shear stress and the elastic strain across a fault drop. The strain drop is proportional to the ratio of average slip on the fault, Du, to the typical dimension of the rupture (or the smaller of the dimensions if the fault is long and narrow), so that the dimensionless constant depends on the geometry of faulting. The stress drop, in turn, is given by the product of the strain drop and the shear modulus, m. For small events, the simplest assumption is a circular rupture of radius r across which the slip varies smoothly from zero on the circumference to a maximum at the center. For this case the stress drop can be expressed as [Eshelby, 1957]
Ds 7p Du m 16 r 2

Kanamori [1979] originally tested (4) with magnitudes measured as Gutenberg and Richter had defined them earlier, but now, because seismic moments are commonly measured for moderate and large earthquakes, seismologists use MW defined by (4). In the following, we use MW for earthquake magnitude, regardless of how it was measured.) From (3) and (4) we can estimate typical spatial dimensions of faults or portions of faults that slip in earthquake with different magnitudes or seismic moments. [26] Suppose, for geomorphologic purposes, that faulting creates fracture spacing of 1 m. Because two faults cannot intersect one another without one displacing the other and hence creating 3 faults, for a network of faults to create blocks with dimensions of a meter, some faults must be only 1 m (or shorter) in dimension. The application of (3) and (4) suggests that an earthquake with MW = 2.4 will rupture a fault with radius r = 0.5 m (diameter of 1 m). Thus, where seismicity is common, as in tectonically active regions, we expect that blocks with dimensions of boulders, if not yet smaller, participate in regional deformation, insofar as such small earthquakes occur. [27] Most of the studies of aftershocks cited above consider earthquakes with magnitudes of 1 to 3 (or larger). Seismographs record even smaller earthquakes, however, where they can be installed with high sensitivity, such as on the seafloor [Butler, 2003], underground [e.g., Gibowicz et al., 1991; McGarr and Green, 1978], or in boreholes [e.g., Abercrombie, 1995; Teng and Henyey, 1981]. Borehole instruments in southern California, and those in mines in Canada and South Africa, have recorded earthquakes with magnitudes as small as 3, nanoearthquakes. Moreover, working in a deep mine, McGarr and Green [1978] could record essentially all earthquakes with magnitudes as small as 3. They showed that for MW > 3 the seismicity obeys the Gutenberg-Richter frequency magnitude relationship:
log10 N MW a bMW 5

[23] Using the seismic moment defined by (1) and A = pr2, (2) can be written as
r
3

r 7M0 16Ds

[24] Thus, if we know MO, and we assume that for very small earthquakes Ds % 1 MPa [e.g., Hanks, 1977; Ide and Beroza, 2001], we may deduce their typical rupture dimensions. [25] Magnitudes scale with the seismic moments [e.g., Hanks and Kanamori, 1979]:
log10 M0 1:5MW 9:0 4

where M0 is measured in Newton-meters, and MW is the magnitude calculated from the seismic moment. (Hanks and

where N(MW) is the number of earthquakes with magnitude greater than or equal to MW, and a and b are empirical constants. At a yet smaller dimension, Scholz [1968] found that the Gutenberg-Richter relationship in (5) applies also to acoustic emissions due to microcracking within samples with dimensions of a few centimeters that he stressed in the laboratory. Thus we have no reason to doubt that earth3 occur frequently in actively quakes with MW deforming regions. [28] We note that McGarr and Green [1978] found that b = 0.5 0.75 in (5), which is somewhat smaller than the typical values of b = 0.9 1.0 and hence might suggest a difference between nanoearthquakes and larger earthquakes. Scholz [1968] showed, however, that the value of b depends on the level of stress difference, the maximum minus minimum compressive stress applied to a sample in the laboratory; b is lower for samples deformed at large stress difference. Stresses that caused earthquakes in the mine are likely to have been larger than those causing earthquakes elsewhere [e.g., McGarr et al., 1975]. Thus the fact that the value of b found by McGarr and Green [1978] is smaller than that characteristic of larger earthquakes does not imply a fundamental difference between tiny earthquakes and those we feel.

7 of 12

F03014

MOLNAR ET AL.: TECTONICS, FRACTURING, AND EROSION

F03014

[29] Small earthquakes contribute less total strain to a region than large events, and might seem unimportant in this context. Because the scaling between frequency of occurrence and average area that ruptures in earthquakes, the total surface area that slips in earthquakes of any magnitude or moment is the same (one event of MW $ 7 ruptures roughly the same area as the 10 events of MW $ 6 expected from (5) with b = 1) [Hanks, 1992]. [30] Because only rare earthquakes seem to rupture new faults, we do not claim that the occurrence of a few small earthquakes with MW $ 3 implies that another block with dimensions of $1 m has been created. Most earthquakes occur on preexisting faults, and the total displacement on most faults accumulates with many earthquakes. Thus, to some extent, seismicity can give a biased measure of the rate of tectonic fracturing. Because faults can be reactivated at different times, some might argue that most faults on which earthquakes occur existed before the most recent phase of tectonic activity. Yet, because we commonly date tectonic activity by dating when faults rupture rock of known age, some, if not most, faults must become active when a region becomes active. Moreover, as relatively warm, unfractured rock in the lower crust is exhumed, if the region is tectonically active, new faults must form within that rock as it cools, and as brittle deformation becomes possible. In any case, although we associate tiny earthquakes with small rupture areas, only some small rupture areas bound small blocks, and only rare earthquakes extend or create new faults, and hence new blocks. 2.3. Earthquakes and Faults as Fractals and Self-Similarity [31] The power law distribution that characterizes the Gutenberg-Richter relationship for earthquake magnitudes in (5) implies scale invariance. Thus small earthquakes differ little from large ones, except in dimensions of ruptures and averaged displacements on faults. Analyses of populations of faults with a wide range of dimensions and offsets also show power law distributions and therefore also scale invariance [e.g., Hirata, 1989; Marrett and Allmendinger, 1990; Scholz and Cowie, 1990; Walsh and Watterson, 1987, 1992; Walsh et al., 1991]. A fractal distribution of fault dimensions implies a fractal distribution of dimensions of blocks that the fractures separate, and indeed, areas of lithospheric plates also obey power law, or fractal, distributions [Bird, 2003; Sornette and Pisarenko, 2003]. By analogy, we expect that the crust consists of a few large blocks, many smaller ones, and yet many more tiny blocks, such that the number of blocks larger than a certain size will scale with that size raised to a negative power. King [1983] showed that the commonly observed b % 1 in (5) can be explained by assuming, first, that for each block with a certain volume V or larger, there are another $10 blocks of volume 0.1 V or larger, and so on, and, second that a similar scaling applies to rupture dimensions of faults that slip in earthquakes with different magnitudes [see also Sammis and King, 2007]. [32] Scale invariance encourages extrapolations to dimensions of faults smaller than can be studied thoroughly in the field. Accordingly, much attention has been paid to the question of what fraction of regional deformation occurs because of slip on the largest faults, with views ranging

from most (perhaps 90%) occurring on the main faults [e.g., Scholz and Cowie, 1990; Scholz et al., 1993] to as much as 30 50% of the regional strain contributed by slip on small faults, those for which dimensions and amounts of slip are orders of magnitude smaller than the dimensions and offsets of the larger faults [e.g., Kautz and Sclater, 1988; Marrett and Allmendinger, 1992; Turcotte, 1986; Walsh et al., 1991]. Differences in deductions derive in large part from different estimates of the frequency-magnitude distributions of fault lengths and of offsets [e.g., Cowie and Scholz, 1992; Gillespie et al., 1992; Marrett and Allmendinger, 1991; Scholz et al., 1993]. Insofar as the discussion given above for earthquakes can be extrapolated to dimensions comparable to boulders, relatively small earthquakes, those whose magnitude is smaller than the largest possible earthquake in a region by one unit or more, contribute tens of percent of the strain, but almost surely less than 50% of the total strain [e.g., Chen and Molnar, 1977; Molnar, 1979]. [33] Analyses of faults in settings where strain is large suggest that the power law distribution that works so well for earthquakes gives way to an exponential distribution [e.g., Cowie et al., 1993; Spyropoulos et al., 1999]. As strain accumulates, stresses associated with slip on major faults interact so as to suppress continued slip on some minor faults [e.g., Gupta et al., 1998]. Analyses of clay gouges in fault zones also show that the fractal distribution fails when the dimensions of clay particles are reached [Wilson et al., 2005]. Thus, in regions of high strain, a power law relationship overestimates the contribution of slip on small faults to the strain rate. Spyropoulos et al. [2002] showed that as strain accumulates, faults that rupture the characteristic dimension of the deforming region, such as the thickness of the seismogenic layer of the crust, set the scale for the exponential distribution. Similarly, joint spacing in layered rock scales with layer thickness [e.g., Pollard and Segall, 1987; Wu and Pollard, 1995], and the degree to which such scaling develops quickly and accurately depends on the initial distribution of flaws and their fracture toughness [e.g., Fischer and Polansky, 2006; Olson, 2004]. If the patterns observed in the laboratory by Spyropoulos et al. [1999, 2002] and Wu and Pollard [1995], in computer simulations [e.g., Fischer and Polansky, 2006; Olson, 2004], at mid-ocean ridges by Cowie et al. [1993], and fold-and-thrust belts [e.g., Morellato et al., 2003] applied to the rest of the Earth, we might expect that in tectonically active regions, small fractures would become inactive as strain became increasingly concentrated on only a few major faults. Studies of seismicity seem to be inadequate to test whether small earthquakes also occur more rarely than the Gutenberg-Richter recurrence relationship in (5) suggests. The transfer of strain from developing partly by slip on small faults to almost entirely by slip only on large faults, however, need not affect the existence of small blocks; the creation of many small faults that bound small blocks in the initial stages of deformation before strain increases beyond a few percent, should still leave the region riddled with fractures (Figure 2). [34] Although fractures and small blocks of rock characterize the Earths surface in many regions, tectonic fracturing of rock occurs at depth, before those fractures and fragments are exhumed and below which direct observation is difficult. Nevertheless, small faults are particularly of

8 of 12

F03014

MOLNAR ET AL.: TECTONICS, FRACTURING, AND EROSION

F03014

Figure 9. Comparison of two domes of similar scale in massive rock. (left) North dome in Yosemite National Park, formed in Half Dome granodiorite, the westernmost unit of the Tuolumne Intrusive Series, and (right) dome in Archean rocks of the Periyar drainage in Kerala state, western Ghats, India. interest to some communities. Sibson [2001] pointed out that meshes of small faults are crucial for the fluid flow that occurs both when ores are deposited and in settings where earthquakes occur on faults with orientations such that in the absence of fluid pressure friction would prevent slip. Citing literature from the early 20th century, Sibson [2001] stressed that ore deposits are commonly associated with faults with small displacements, not with major faults. Exhumed mineral belts show that the fracturing of crust at depth, necessarily of tectonic origin, is not a rare phenomenon. 2.4. Flaw-Free Rocks [35] Here we speculate that a corollary to the arguments given above about the role of tectonics as a crusher of rock is that in those places where rock has dodged the rock crusher, it may be stronger and less easily removed by erosive agents. Granitic intrusions in arc settings are one example. Such rock has not been subjected to thrust faulting. In fact, the intrusive mechanisms now envisaged for such rock [e.g., Coleman et al., 2004; Glazner et al., 2004] may allow repeated annealing of incipient cooling joints, producing yet fewer flaws for erosion to gain purchase. In Yosemite National Park, the walls of El Capitan and other major climbing targets are particularly flaw-free. Another geomorphic feature may require flaw-free rock: domes, whose occurrence in strong or massive rocks has been long recognized. It has been pointed out since yet another of Gilberts [1904] publications that these features are characterized by sheet joints that form parallel to the local topographic surface. Although the mechanics of sheet joint formation has long been an enigma, recent work of Martel [2006] calls upon a combination of regional horizontal stresses and local topographic curvature to generate opening mode cracks. It is unlikely that these joints would form, if the rock were already riddled with fractures that could accommodate the horizontal stresses. The occurrence of similar domes in both the Sierran arc granite and the Archean deep crustal rock of the Western Ghats in India (Figure 9) can be explained by lack of significant tectonic (faulting) activity since they cooled sufficiently to lie in the brittle regime.

3. Summary
[36] Tectonics can affect erosion rates both by elevating terrain so that erosive agents gain potential energy and by fracturing rock. The latter effect increases the likelihood of transport by solution, by a variety of hillslope processes, and by rivers and glaciers. Recognition of either role played by tectonics is not new. We contend, however, that the former has been overly emphasized, and that the latter seems to be less well appreciated than it deserves. [37] Tectonic processes that operate in the brittle part of the crust facilitate erosion by fracturing rock. First, slip on faults that are not planar requires deformation of one or both of the blocks that slide past one another; with finite amounts of slip on major faults, the adjacent blocks must deform with permanent strain. Moreover, earthquakes within continental regions occur not just on major faults, but also on widely distributed, smaller faults. Both seismic moments of these earthquakes and dimensions of the faults that rupture obey power law frequency-magnitude distributions, which suggest self-similarity. Thus small earthquakes and small faults differ from large ones only in size, and we should expect much of the crust in tectonically active areas to be pervasively fractured to depths of $10 20 km. We contend that it is difficult to imagine tectonic deformation of the brittle upper crust that does not include pervasive fracturing. [38] Fracturing of the crust contributes to erosion in a number of ways. In the logic of Gilbert and Dutton, fractured rock has already undergone disintegration, the

9 of 12

F03014

MOLNAR ET AL.: TECTONICS, FRACTURING, AND EROSION

F03014

first stage of erosion. The erosion rates we measure in tectonically active regions are increased because the rock has been through this initial step before reaching depths accessed by surface processes. The creation of fragments amenable to transport down hillslopes or by rivers and glaciers eases the burden of rivers and glaciers. Processes such as landsliding, themselves facilitated by fractured rock at the surface, in turn, tend to break fragments into smaller pieces, making them yet easier to transport by rivers and glaciers below [e.g., Crosta et al., 2007]. The creation of fractures also increases the surface area of rock that can be attacked chemically, and hence enhances the rate of chemical erosion. Enhanced chemical erosion, in turn, contributes to the disintegration of rock into fragments easily transported by rivers, glaciers, and hillslopes. Correlations of chemical and physical erosion testify to the importance of enhanced surface area for both [e.g., Gaillardet et al., 1999]. [39] The association of rapid erosion where rock has been tectonically fractured gains support from the opposite case. High relief in tectonically quiescent regions, like the Western Ghats (Sahyadri) of India [Gunnell et al., 2003], the highlands of Sri Lanka [von Blanckenburg et al., 2004], and the Guyana Shield [Brown et al., 1992; Edmond et al., 1995; Stallard, 1985, 1988], does not lead to rapid erosion, despite high rainfall in all of these areas. As these authors recognized, the resistant intact rock of these regions retards its erosion. In these regions, tectonic activity has been mild, if present at all, since the high-grade rock was deformed by largely ductile processes in Precambrian time and later was exhumed. This rock may never have been extensively fractured. We have noted that a similar case may be made for the resistance to erosion of igneous rocks injected into the upper crust after the crust has undergone most of its strain. Some exhumed granitic plutons that form the cores of high mountain ranges may be massive (unfractured), and therefore sufficiently strong to stand in high local relief above the adjacent valleys because they have escaped the tectonic rock crusher that prepared the surrounding country rock for erosion. [40] All surely agree that geodynamic processes play a vital role in shaping the Earths surface, but not only by building mountains. At a small scale, these same processes that build mountains also sow the seeds of their destruction. If we are to benefit fully from the insights of our forbearers, Dutton and Gilbert, we must understand what they recognized as the first step in erosion: disintegration. [41] Acknowledgments. The writing of this paper was stimulated by field trips led by A. C. Narayana to the Western Ghats and by K. S. Kellogg to Colorados Front Range. One of us (P.M.) also discussed some aspects of tectonic fracturing with S. Willett in 2003. D. Hatzfeld helped with the preparation of Figures 4, 5, and 7. We thank J. D. Stock for two especially thorough, prompt, constructive reviews and S. Gupta, P. O. Koons, and K. X Whipple for additional helpful suggestions. R.S.A. acknowledges financial support of the NSF through grant EAR-0126253. Molnar thanks the Miller Foundation at the University of California in Berkeley for support during the final preparation of the manuscript. Our travel to the Western Ghats was supported in part by a University of Colorados Council on Research and Creative Works seed grant to S.P.A.

References
Abercrombie, R. E. (1995), Earthquake source scaling relationships from 1 to 5 ML using seismograms recorded at 2.5-km depth, J. Geophys. Res., 100, 24,015 24,036. Abercrombie, R. E., T. H. Webb, R. Robinson, P. J. McGinty, J. J. Mori, and R. J. Beavan (2000), The enigma of the Arthurs Pass, New Zealand,

earthquake: 1. Reconciling a variety of data for an unusual earthquake, J. Geophys. Res., 105, 16,119 16,137. Abercrombie, R. E., S. Bannister, A. Pancha, T. H. Webb, and J. J. Mori (2001), Determination of fault planes in a complex aftershock sequence using two-dimensional slip inversion, Geophys. J. Int., 146, 134 142. Aki, K. (1966), Generation and propagation of G waves from the Niigata earthquake of June 16, 1964, 2, Estimation of the earthquake moment, released energy, and stress-strain drop from G wave spectrum, Bull. Earthquake Res. Inst. Tokyo, 44, 73 88. Attal, M., and J. Lave (2006), Changes in bedload characteristics along the Marsyandi River (central Nepal): Implications for understanding hillslope sediment supply, sediment load evolution along fluvial networks, and denudation in active orogenic belts, Spec. Pap. Geol. Soc. Am., 398, 143 171. Beaumont, C., P. Fullsack, and J. Hamilton (1992), Erosional control of active compressional orogens, in Thrust Tectonics, edited by K. R. McClay, pp. 1 18, Chapman and Hall, New York. Bird, P. (2003), An updated digital model of plate boundaries, Geochem. Geophys. Geosyst., 4(3), 1027, doi:10.1029/2001GC000252. Bodin, P., and S. Horton (2004), Source parameters and tectonic implications of aftershocks of the Mw 7.6 Bhuj earthquake of 26 January 2001, Bull. Seismol. Soc. Am., 94, 818 827. Boulton, G. S., and R. Vivian (1973), Underneath the glaciers, Geogr. Mag., 45, 311 319. Briner, J. P., and T. W. Swanson (1998), Using inherited cosmogenic 36Cl to constrain glacial erosion rates of the Cordilleran ice sheet, Geology, 26, 3 6. Brown, E. T., R. F. Stallard, G. M. Raisbeck, and F. Yiou (1992), Determination of the denudation rate of Mount Roraima, Venezuela using cosmogenic 10Be and 26Al, Eos Trans. AGU, 73(43), 170. Burbank, D. W. (2005), Cracking the Himalaya, Nature, 434, 963 964. Burbank, D. W., J. Leland, E. Fielding, R. S. Anderson, N. Brozovic, M. R. Reid, and C. Duncan (1996), Bedrock incision, rock uplift, and threshold hillslopes in the northwestern Himalayas, Nature, 379, 505 510. Butler, R. (2003), The Hawaii-2 Observatory: Observation of nanoearthquakes, Seismol. Res. Lett., 74, 290 297. Chen, W.-P., and P. Molnar (1977), Seismic moments of major earthquakes and the average rate of slip in Central Asia, J. Geophys. Res., 82, 2945 2969. Coleman, D. S., W. Gray, and A. F. Glazner (2004), Rethinking the emplacement and evolution of zoned plutons: Geochronologic evidence for incremental assembly of the Tuolumne Intrusive Suite, California, Geology, 32, 433 436. Cowie, P. A., and C. H. Scholz (1992), Displacement-length scaling relationship for faults: Data synthesis and discussion, J. Struct. Geol., 14, 1149 1156. Cowie, P. A., C. H. Scholz, M. Edwards, and A. Malinvero (1993), Fault strain and seismic coupling on mid-ocean ridges, J. Geophys. Res., 98, 17,911 17,920. Cox, S. C., D. Craw, and C. P. Chamberlain (1997), Structure and fluid migration in a late Cenozoic duplex system forming the Main Divide in the central Southern Alps, New Zealand, N. Z. J. Geol. Geophys., 40, 359 373. Crosta, G. B., P. Frattini, and N. Fusi (2007), Fragmentation in the Val Pola rock avalanche, Italian Alps, J. Geophys. Res., 112, F01006, doi:10.1029/ 2005JF000455. Dahlen, F. A., and J. Suppe (1988), Mechanics, growth, and erosion of mountain belts, Spec. Pap. Geol. Soc. Am., 218, 161 178. Das, S., and C. Henry (2003), Spatial relation between main earthquake slip and its aftershock distribution, Rev. Geophys., 41(3), 1013, doi:10.1029/ 2002RG000119. Dutton, C. E. (1882), The Tertiary History of the Grand Canyon, U.S. Geol. Surv. Monogr., vol. 2, U.S. Gov. Print. Office, Washington, D. C. Eaton, J. P., M. E. ONeill, and J. N. Murdock (1970), Aftershocks of the 1966 Parkfield-Cholame, California, earthquakes: A detailed study, Bull. Seismol. Soc. Am., 60, 1151 1197. Edmond, J. M., M. R. Palmer, C. I. Measures, B. Grant, and R. F. Stallard (1995), The fluvial geochemistry and denudation rate of the Guyana Shield in Venezuela, Colombia, and Brazil, Geochim. Cosmochim. Acta, 59, 3301 3325. Eshelby, J. D. (1957), The determination of the elastic field of an ellipsoidal inclusion and related problems, Proc. R. Soc. London, Ser. A, 241, 376 396. Fedotov, S. A. (1965), Regularities of the distribution of strong earthquakes in Kamchatka, the Kurile Islands, and northeastern Japan (in Russian), Trans. Inst. Fiz. Zemli Akad. Nauk SSSR, 36, 66 93. Fischer, M. P., and A. Polansky (2006), Influence of flaws on joint spacing and saturation: Results of one-dimensional mechanical modeling, J. Geophys. Res., 111, B07403, doi:10.1029/2005JB004115.

10 of 12

F03014

MOLNAR ET AL.: TECTONICS, FRACTURING, AND EROSION

F03014

Gaillardet, J., B. Dupre, P. Louvat, and C. J. Allegre (1999), Global silicate ` weathering and CO2 consumption rates deduced from the chemistry of large rivers, Chem. Geol., 159, 3 30. Gibowicz, S. J., R. P. Young, S. Talebi, and D. J. Rawlence (1991), Source parameters of seismic events at the Underground Research Laboratory in Manitoba, Canada: Scaling relations for events with moment magnitude smaller than 2, Bull. Seismol. Soc. Am., 81, 1157 1182. Gilbert, G. K. (1877), Land sculpture, in Report on the Geology of the Henry Mountains: Geographical and Geological Survey of the Rocky Mountain Region, pp. 93 144, U.S. Gov. Print. Office, Washington, D. C. Gilbert, G. K. (1904), Dome and dome structures of the High Sierra, Geol. Soc. Am. Bull., 15, 29 36. Gillespie, P. A., J. J. Walsh, and J. Watterson (1992), Limitations of dimension and displacement data from single faults and the consequences for data analysis and interpretation, J. Struct. Geol., 14, 1157 1172. Glazner, A. F., J. M. Bartley, D. S. Coleman, W. Gray, and R. Z. Taylor (2004), Are plutons assembled over millions of years by amalgamation from small magma chambers?, GSA Today, 14, 4 11. Gunnell, Y., K. Gallagher, A. Carter, M. Widdowson, and A. J. Hurford (2003), Denudation history of the continental margin of western peninsular India since the early MesozoicReconciling apatite fission-track data with geomorphology, Earth Planet. Sci. Lett., 215, 187 201. Gupta, S., P. A. Cowie, N. H. Dawers, and J. R. Underhill (1998), A mechanism to explain rift-basin subsidence and stratigraphic patterns through fault-array evolution, Geology, 26, 595 598. Hack, J. T. (1957), Studies of longitudinal stream profiles in Virginia and Maryland, U.S. Geol. Surv. Prof. Pap., 294B, 45 97. Hack, J. T. (1973), Stream-profile analysis and stream-gradient index, J. Res. U.S. Geol. Surv., 1, 421 429. Hamilton, R. M., and J. H. Healy (1969), Aftershocks of the Benham nuclear explosion, Bull. Seismol. Soc. Am., 59, 2271 2281. Hamilton, R. M., B. E. Smith, F. G. Fischer, and P. J. Papanek (1972), Earthquakes caused by underground nuclear explosions on Pahute Mesa, Nevada Test Site, Bull. Seismol. Soc. Am., 62, 1319 1341. Hancock, G. S., R. S. Anderson, and K. X. Whipple (1998), Beyond power: Bedrock river incision process and form, in Rivers Over Rock: Fluvial Processes in Bedrock Channels, Geophys. Monogr. Ser., vol. 107, edited by K. J. Tinkler and E. E. Wohl, pp. 35 60, AGU, Washington, D. C. Hanks, T. C. (1977), Earthquake stress drops, ambient tectonic stresses, and stresses that drive plate motions, Pure Appl. Geophys., 115, 441 458. Hanks, T. C. (1992), Small earthquakes, tectonic forces, Science, 256, 1430 1432. Hanks, T. C., and H. Kanamori (1979), A moment magnitude scale, J. Geophys. Res., 84, 2348 2351. Hardebeck, J. L., et al. (2004), Preliminary report on the 22 December 2003 M6.5 San Simeon, California, earthquake, Seismol. Res. Lett., 75, 155 172. Hatzfeld, D., A. A. Christodoulou, E. M. Scordilis, D. Papagiotopoulos, and P. M. Hatzidimitriou (1986), A microearthquake study of the Mygdonian graben (northern Greece), Earth Planet. Sci. Lett., 81, 379 396. Hatzfeld, D., et al. (1997), The Kozani-Grevena (Greece) earthquake of May 13, 1995 revisited from a detailed seismological study, Bull. Seismol. Soc. Am., 87, 463 473. Hatzfeld, D., V. Karakostas, M. Ziazia, I. Kassaras, E. Papadimitriou, K. Makropoulos, N. Voulgaris, and C. Papaioannou (2000), Microseismicity and faulting geometry in the Gulf of Corinth (Greece), Geophys. J. Int., 141, 438 456. Hauksson, E., L. M. Jones, K. Hutton, and D. Eberhart-Phillips (1993), The 1992 Landers earthquake sequence: Seismological observations, J. Geophys. Res., 98, 19,835 19,858. Hirata, T. (1989), Fractal dimension of fault systems in Japan: Fractal structure in rock fracture geometry at various scales, Pure Appl. Geophys., 131, 157 170. Holmes, A. (1944), Principles of Physical Geology, 532 pp., Thomas Nelson, London. Holmes, A. (1965), Principles of Physical Geology, 2nd ed., 1288 pp., Ronald Press, New York. Ide, S., and G. C. Beroza (2001), Does apparent stress vary with earthquake size?, Geophys. Res. Lett., 28, 3349 3352. Ishiyama, T., K. Mueller, M. Togo, A. Okada, and K. Takemura (2004), Geomorphology, kinematic history, and earthquake behavior of the active Kuwana wedge thrust anticline, central Japan, J. Geophys. Res., 109, B12408, doi:10.1029/2003JB002547. Kautz, S. A., and J. G. Sclater (1988), Internal deformation in clay models of extension by block faulting, Tectonics, 7, 823 832. Kelleher, J., L. R. Sykes, and J. Oliver (1973), Possible criteria for predicting earthquake locations and their application to major plate boundaries of the Pacific and Caribbean, J. Geophys. Res., 78, 2547 2585.

Kellogg, K. S. (2001), Tectonic controls on a large landslide complex: Williams Fork Mountains near Dillon, Colorado, Geomorphology, 41, 355 368. King, G. (1983), The accommodation of large strains in the upper lithosphere of the Earth and other solids by self-similar fault systems: The geometrical origin of b-value, Pure Appl. Geophys., 121, 761 815. King, G. C. P., Z. X. Ouyang, P. Papadimitriou, A. Deschamps, J. Gagnepain, G. Houseman, J. A. Jackson, C. Soufleris, and J. Virieux (1985), The evolution of the Gulf of Corinth (Greece): An aftershock study of the 1981 earthquakes, Geophys. J. R. Astron. Soc., 80, 677 693. King, G. C. P., R. S. Stein, and J. Lin (1994), Static stress changes and the triggering of earthquakes, Bull. Seismol. Soc. Am., 84, 935 953. Koons, P. O. (1989), The topographic evolution of collisional mountain belts: A numerical look at the Southern Alps of New Zealand, Am. J. Sci., 289, 1041 1069. Koons, P. O. (1990), Two-sided orogen: Collision and erosion from the sandbox to Southern Alps, New Zealand, Geology, 18, 679 682. Koons, P. O., and D. Craw (2002), Beyond steady state in threedimensional orogens, Eos Trans. AGU, 83(47), Fall Meet. Suppl., Abstract T72B-06. Little, T. A., R. J. Holcombe, and B. R. Ilg (2002), Kinematics of oblique collision and ramping inferred from microstructures and strain in middle crustal rocks, central Southern Alps, New Zealand, J. Struct. Geol., 24, 219 239. Lliboutry, L. (1994), Monolithologic erosion of hard beds by temperate glaciers, J. Glaciol., 40, 433 450. Loveless, J. P., G. D. Hoke, R. W. Allmendinger, G. Gonzalez, B. L. Isacks, and D. A. Carrizo (2005), Pervasive cracking of the northern Chilean Coastal Cordillera: New evidence for forearc extension, Geology, 33, 973 976. Marrett, R., and R. W. Allmendinger (1990), Kinematic analysis of faultslip data, J. Struct. Geol., 12, 973 986. Marrett, R., and R. W. Allmendinger (1991), Estimates of strain due to brittle faulting: Sampling the fault populations, J. Struct. Geol., 13, 735 738. Marrett, R., and R. W. Allmendinger (1992), Amount of extension on small faults: An example from the Viking Graben, Geology, 20, 47 50. Martel, S. J. (2006), Effect of topographic curvature on near-surface stresses and application to sheeting joints, Geophys. Res. Lett., 33, L01308, doi:10.1029/2005GL024710. McGarr, A., and R. W. E. Green (1978), Microtremor sequences and tilting in a deep mine, Bull. Seismol. Soc. Am., 68, 1679 1697. McGarr, A., S. M. Spottiswoode, and N. C. Gay (1975), Relationship of mine tremors to induced stresses and rock properties in the focal region, Bull. Seismol. Soc. Am., 65, 981 993. McGarr, A., R. W. E. Green, and S. M. Spottiswoode (1981), Strong ground motion of mine tremors: Some implications for near-source ground motion parameters, Bull. Seismol. Soc. Am., 71, 295 319. McKeown, F. A., and D. D. Dickey (1969), Fault displacements and motions related to nuclear explosions, Bull. Seismol. Soc. Am., 59, 2253 2269. Molnar, P. (1979), Earthquake recurrence intervals and plate tectonics, Bull. Seismol. Soc. Am., 69, 115 133. Molnar, P. (1987), Inversion of profiles of uplift rates for the geometry of dip-slip faults at depth, with examples from the Alps and the Himalaya, Ann. Geophys., 5B, 663 670. Morellato, C., F. Redini, and C. Doglioni (2003), On the number and spacing of faults, Terra Nova, 15, 315 321. Nadeau, R. M., and L. R. Johnson (1994), Seismological studies at Parkfield VI: Moment release rates and estimates of source parameters for small repeating earthquakes, Bull. Seismol. Soc. Am., 88, 790 814. Nadeau, R., M. Antolik, P. A. Johnson, W. Foxall, and T. V. McEvilly (1994), Seismological studies at Parkfield III: Microearthquake clusters in the study of fault-zone dynamics, Bull. Seismol. Soc. Am., 84, 247 263. Negishi, H., J. Mori, T. Sato, R. Singh, S. Kumar, and N. Hirata (2002), Size and orientation of the fault plane for the 2001 Gujarat, India earthquake (Mw7.7) from aftershock observations: A high stress drop event, Geophys. Res. Lett., 29(20), 1949, doi:10.1029/2002GL015280. Olson, J. E. (2004), Predicting fracture swarmsThe influence of subcritical crack growth and the crack-tip process zone on joint spacing in rock, Spec. Publ. Geol. Soc., 231, 73 87. Ouyed, M., G. Yielding, D. Hatzfeld, and G. C. P. King (1983), An aftershock study of the El Asnam (Algeria) earthquake of 1980 October 10, Geophys. J. R. Astron. Soc., 73, 605 639. Pandey, M. R., R. P. Tandukar, J. P. Avouac, J. Lave, and J. P. Massot (1995), Interseismic strain accumulation on the Himalayan crustal ramp (Nepal), Geophys. Res. Lett., 22, 751 754.

11 of 12

F03014

MOLNAR ET AL.: TECTONICS, FRACTURING, AND EROSION

F03014

Pettifer, G. S., and P. G. Fookes (1994), A revision of the graphical method for assessing the excavatability of rock, Q. J. Eng. Geol., 27, 145 164. Pollard, D. D., and P. Segall (1987), Theoretical displacements and stresses near fractures in rock: With applications to faults, joints, veins, dikes, and solution surfaces, in Fracture Mechanics of Rock, edited by B. K. Atkinson, pp. 277 349, Elsevier, New York. Resor, P. G., D. D. Pollard, T. J. Wright, and G. C. Beroza (2005), Integrating high-precision aftershock locations and geodetic observations to model coseismic deformation associated with the 1995 Kozani-Grevena earthquake, Greece, J. Geophys. Res., 110, B09402, doi:10.1029/ 2004JB003263. Robinson, R., and P. J. McGinty (2000), The enigma of the Arthurs Pass, New Zealand, earthquake: 2. The aftershock distribution and its relation to regional and induced stress fields, J. Geophys. Res., 105, 16,139 16,150. Sammis, C. G., and G. C. P. King (2007), Mechanical origin of power law scaling in fault zone rock, Geophys. Res. Lett., 34, L04312, doi:10.1029/ 2006GL028548. Schmidt, K. M., and D. R. Montgomery (1995), Limits to relief, Science, 270, 617 620. Scholz, C. H. (1968), The frequency-magnitude relation of microfracturing in rock and its relation to earthquakes, Bull. Seismol. Soc. Am., 58, 399 415. Scholz, C. H., and P. A. Cowie (1990), Determination of total strain from faulting using slip measurements, Nature, 346, 837 839. Scholz, C. H., N. H. Dawers, J.-Z. Xu, M. H. Anders, and P. A. Cowie (1993), Fault growth and fault scaling laws: Preliminary results, J. Geophys. Res., 98, 21,951 21,961. Schulte-Pelkum, V., G. Monsalve, A. Sheehan, M. R. Pandey, S. Sapkota, R. Bilham, and F. Wu (2005), Imaging the Indian subcontinent beneath the Himalaya, Nature, 435, 1222 1225. Segall, P. (1984), Formation and growth of extensional fracture sets, Geol. Soc. Am. Bull., 95, 454 462. Selby, M. J. (1980), A rock mass strength classification for geomorphic purposes: With tests from Antarctica and New Zealand, Z. Geomorphol., 24, 31 51. Shaw, J. H., and J. Suppe (1994), Active faulting and growth folding in the eastern Santa Barbara Channel, California, Geol. Soc. Am. Bull., 106, 607 626. Shearer, P. M., J. L. Hardebeck, L. Astiz, and K. B. Richards-Dinger (2003), Analysis of similar event clusters in aftershocks of the 1994 Northridge, California, earthquake, J. Geophys. Res., 108(B1), 2035, doi:10.1029/2001JB000685. Sibson, R. H. (2001), Seismogenic framework for hydrothermal transport and ore deposition, Soc. Econ. Geol. Rev., 14, 25 50. Sklar, L., and W. E. Dietrich (2001), Sediment and rock strength controls on river incision into bedrock, Geology, 29, 1087 1090. Sklar, L. S., and W. E. Dietrich (2004), A mechanistic model for river incision into bedrock by saltating bed load, Water Resour. Res., 40, W06301, doi:10.1029/2003WR002496. Sleep, N. H. (1971), Thermal effects of the formation of Atlantic continental margin by continental break up, Geophys. J. R. Astron. Soc., 24, 325 350. Sornette, D., and V. Pisarenko (2003), Fractal plate tectonics, Geophys. Res. Lett., 30(3), 1105, doi:10.1029/2002GL015043. Soufleris, C., J. A. Jackson, G. C. P. King, C. P. Spencer, and C. H. Scholz (1982), The 1978 earthquake sequence near Thessaloniki (northern Greece), Geophys. J. R. Astron. Soc., 68, 429 458. Spottiswoode, S. M., and A. McGarr (1975), Source parameters of tremors in a deep-level gold mine, Bull. Seismol. Soc. Am., 65, 93 112. Spyropoulos, C., W. J. Griffith, C. H. Scholz, and B. E. Shaw (1999), Experimental evidence for different strain regimes of crack populations in a clay model, Geophys. Res. Lett., 26, 1081 1084. Spyropoulos, C., C. H. Scholz, and B. E. Shaw (2002), Transition regimes for growing crack populations, Phys. Rev. Lett., 65, 056105, doi:10.1103/ PhysRevE.65.05610584. Stallard, R. F. (1985), River chemistry, geology, geomorphology, and soils in the Amazon and Orinoco basins, in The Chemistry of Weathering, edited by J. I. Drever, pp. 293 316, Springer, New York. Stallard, R. F. (1988), Weathering and erosion in the humid tropics, in Physical and Chemical Weathering in Geochemical Cycles, edited by A. Lerman and M. Meybeck, pp. 225 246, Springer, New York.

Stein, R. S., and M. Lisowski (1983), The 1979 Homestead Valley earthquake sequence, California: Control of aftershocks and postseismic deformation, J. Geophys. Res., 88, 6477 6490. Stock, J. D., and W. E. Dietrich (2006), Erosion of steepland valleys by debris flows, Geol. Soc. Am. Bull., 118, 1125 1148. Stock, J. D., D. R. Montgomery, B. D. Collins, W. E. Dietrich, and L. Sklar (2005), Field measurements of incision rates following bedrock exposure: Implications for process controls on the long profiles of valleys cut by rivers and debris flows, Geol. Soc. Am. Bull., 117, 174 194. Suppe, J. (1983), Geometry and kinematics of fault-bend folding, Am. J. Sci., 90, 648 721. Suppe, J. (1985), Principles of Structural Geology, 537 pp., Prentice-Hall, Englewood Cliffs, N. J. Sykes, L. R. (1971), Aftershock zones of great earthquakes, seismicity gaps, and earthquake prediction for Alaska and the Aleutians, J. Geophys. Res., 76, 8021 8041. Teng, T. L., and T. H. Henyey (1981), The detection of nanoearthquakes, in Earthquake Prediction: An International Review, Maurice Ewing Ser., vol. 4, edited by D. W. Simpson and P. G. Richards, pp. 533 542, AGU, Washington, D. C. Thuro, K. (1997), Drillability prediction: Geological influences in hard rock drill and blast tunneling, Geol. Rundsch., 86, 426 438. Turcotte, D. L. (1986), A fractal model for crustal deformation, Tectonophysics, 132, 261 269. Vanacker, V., F. von Blanckenburg, T. Hewawasam, and P. W. Kubik (2007), Constraining landscape development of the Sri Lankan escarpment with cosmogenic nuclides in river sediment, Earth Planet. Sci. Lett., 253, 402 414. Vervoort, A., and K. De Wit (1997), Correlation between dredgeability and mechanical properties of rock, Eng. Geol., 47, 259 267. von Blanckenburg, F., T. Hewawasam, and P. W. Kubik (2004), Cosmogenic nuclide evidence for low weathering and denudation in the wet, tropical highlands of Sri Lanka, J. Geophys. Res., 109, F03008, doi:10.1029/2003JF000049. Wager, L. R. (1937), The Arun River drainage pattern and the rise of the Himalaya, Geogr. J., 89, 239 250. Walsh, J. J., and J. Watterson (1987), Distributions of cumulative displacement and seismic slip on a single normal fault surface, J. Struct. Geol., 9, 1039 1046. Walsh, J. J., and J. Watterson (1992), Populations of faults and fault displacement and their effects on estimates of fault-related regional extension, J. Struct. Geol., 14, 701 712. Walsh, J., J. Watterson, and G. Yielding (1991), The importance of smallscale faulting in regional extension, Nature, 391, 381 393. Weissel, J. K., and M. A. Seidl (1997), Influence of rock strength properties on escarpment retreat across passive continental margins, Geology, 25, 631 634. Wilson, B., T. Dewers, Z. Reches, and J. Brune (2005), Particle size and energetics of gouge from earthquake rupture zones, Nature, 434, 749 752. Wobus, C., A. Heimsath, K. Whipple, and K. Hodges (2005), Active outof-sequence thrust faulting in the central Nepal Himalaya, Nature, 434, 1008 1011. Wu, H.-Q., and D. D. Pollard (1995), An experimental study of the relationship between joint spacing and layer thickness, J. Struct. Geol., 17, 887 905. Yamada, T., J. J. Mori, S. Ide, R. E. Abercrombie, H. Kawakata, M. Nakatani, Y. Iio, and H. Ogasawara (2007), Stress drops and radiated seismic energies of microearthquakes in a South African gold mine, J. Geophys. Res., 112, B03305, doi:10.1029/2006JB004553. Yielding, G., J. A. Jackson, G. C. P. King, H. Sinvhal, C. Vita-Finzi, and R. M. Wood (1981), Relations between surface deformation, fault geometry, seismicity, and rupture characteristics during the El Asnam (Algeria) earthquake of 10 October 1980, Earth Planet. Sci. Lett., 56, 287 304.

R. S. Anderson and S. P. Anderson, Institute of Arctic and Alpine Research, University of Colorado, Boulder, CO 80309-0450, USA. P. Molnar, Department of Geological Sciences, University of Colorado, Boulder, CO 80309-0399, USA. (molnar@colorado.edu)

12 of 12

You might also like