<<

Interactions • Volume 11 (2007) • Paper No. 19 • Page 1

Copyright © 2007, Paper 11-019; 10,756 words, 15 Figures, 0 Animations, 1 Tables. http://EarthInteractions.org

Landslide Rate in the Eastern Cordillera of Northern Bolivia

Troy A. Blodgett* and Bryan L. Isacks Department of Earth and Atmospheric Sciences, Cornell University, Ithaca, New York

Received 28 October 2006; accepted 1 October 2007

ABSTRACT: The northeastern edge of the Bolivian Eastern Cordillera is an example of a tectonically active plateau margin where orographically enhanced precipitation facilitates very high rates of erosion. The topography of the steep- est part of the margin exhibits the classic signature of high erosion rates consisting of high-relief V-shaped valleys where landsliding is the dominant process of hillslope erosion and bedrock are incising into the landscape. The authors mapped landslide scars on multitemporal aerial photographs to estimate hillslope erosion rates. Field surveys of landslide scars are used to calibrate a landslide volume versus area relationship. The mapped area of landsliding, in combination with an estimate of the time for landslide scars to revegetate, leads to an erosion rate estimate. The estimated revegetation time, 10–35 yr, is based on analysis of multitemporal aerial photographs and tree rings. About 4%–6% of two watersheds in the region considered were affected by landslides over the last 10–35 yr. This result implies an erosion rate of 9±5mmyr−1 assuming that 90% of a single landslide reaches the on average. Classified Landsat Thematic Mapper images show that landslides are occurring at approximately the same rate all across an approximately 40-km- wide swath within the high-relief zones of the cordillera. KEYWORDS: Landslide; Erosion; Central Andes

* Corresponding author address: Troy A. Blodgett, GIA, 5355 Armada Dr., Carlsbad, CA 92008. E-mail address: [email protected]

DOI: 10.1175/2007EI222.1

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC

EI222 Logo live 4/C Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 2 1. Introduction In tectonically active mountain belts, uplift and precipitation can produce very high rates of erosion. The removal of mass by erosion can be a major factor in the structural evolution of the mountain belt, as argued in numerous papers since the seminal work of Suppe (Suppe 1981) on the Central Range of Taiwan, and Adams (Adams 1980) on the Southern Alps of New Zealand. The study area is located in a part of the Bolivian Eastern Cordillera along the northeastern flank of the Central Andean Plateau. This plateau edge is characterized by orographic focusingof precipitation and Mid- to Late Cenozoic compressional (e.g., Isacks 1988; Masek et al. 1994; Horton 1999; Barnes and Pelletier 2006). Estimatesof erosion rates have been derived from measurements of suspended load in streams (Guyot et al. 1988; Guyot 1992; Guyot et al. 1993; Summerfield and Hulton 1994; Aalto et al. 2006), cosmogenic radionuclides (Safran et al. 2005), measurements of the cooling ages of minerals (Benjamin et al. 1987; Masek et al. 1994; Safran 1998; Moore and England 2001; Anders et al. 2002; Barnes et al. 2006; Gillis et al. 2006), and estimates of the volume of eroded sediments from stratigraphic and structural measurements (e.g., Masek et al. 1994; Roeder and Chamberlain 1995; McQuarrie 2002). These estimates integrate denudation over time scales ranging from tens of years for the suspended sediment measurements, to hundreds to thousands of years for the cosmogenic nuclides, and to millions of years for the geological and geochemical methods. Also, the erosion rate estimates varyinthe spatial extent over which they are integrated as well as temporally. For example, comogenics and suspended load are point measurements that are converted to integrate over the entire upstream drainage area while a cooling age point mea- surement is probably applicable to only the local sampled structure. Our landslid- ing estimate also occupies a unique place among all of the other estimates in the spatial extent over which it is integrated. However, all the methods are subject to a variety of sampling and interpretative problems, so that additional estimates of erosion rate with a different method are of great value. In this paper we introduce the regional setting of the Bolivian Eastern Cordillera including the influence of structure and climate on erosion. We then describe three distinct geomorphologic belts: glaciated zone, high-relief zone, and the sub- Andean zone. Then we explain how we derive an erosion rate estimate based on field and satellite mapping of landsliding scars observed during a recent 30–35-yr period. This method, rarely applied, can be quite effective in regions where land- sliding is the dominant erosional mechanism (e.g., Simonett. 1967; Hoviusetal. 1997; Hovius et al. 2000). We apply it to a zone of high relief located where the mean slope is 30°–35°, indicative of landsliding as the dominant hillslope ero- sional process. This zone is located at lower elevations and northeast of the heavily glaciated part of the Eastern Cordillera (the Cordillera de la Paz). In this zone, fluvial incision and hillslope landsliding are the dominant erosional processes. The zone also coincides with the steepest part of the northeastern plateau edge. The erosion rate we estimate with landslides is nearly an order of magnitude higher than previously reported measurements and is thus of considerable interest. We examine the result in respect to the reliability of the different methods of estimat- ing erosion rates, the assumptions used in the estimates, and the implications regarding temporal and spatial variations in erosion.

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 3 2. Regional setting The Central Andes, extending across Peru, Bolivia, Chile, and Argentina, con- sist of two mountain chains, the Eastern Cordillera and the Western Cordillera (Figure 1). The Central Andean Plateau includes the Bolivian Altiplano and the Argentine Puna, which are located between the two cordilleras with elevations averaging about 4 km. In this paper we consider the “Eastern Cordillera” as a generic term for the mountain ranges located along the edge and high-elevation flanks of the Central Andean Plateau. In our region (Figure 2), the Eastern Cor- dillera is subdivided into the heavily glaciated, high-elevation Cordillera Real (including the Cordilleras Apolobamba, Munecas de La Paz, and Tres Cruces); and the Yungas, comprising the wet, forested ridges, valleys, and bedrock rivers lo- cated just northeast of, and at lower elevations than, the Cordillera Real. Farther northeast, and at still lower elevations, is the sub-Andes, a zone of ridges and valleys striking northwest–southeast parallel to the plateau boundary. In Figure 2 we show the main regions as 1) Altiplano, 2) Cordillera Real, 3) Yungas, and 4)

Figure 1. Topography of the Northern Andean Plateau and Bolivian Eastern Cordil- lera (from USGS GTOPO30 1-km DEM). The Landsat TM image used in the landslide study (outlined in light blue) falls within the Beni River drainage, which drains the central region at about 15°S. Lake Titicaca is outlined in white and overlaps TM outline.

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC

Fig 1 live 4/C Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 4

Figure 2. Landsat TM image showing the main physiographic regions of northern Bolivia centered on the Eastern Cordillera. The image is produced by combining red, green, and blue images that correspond to TM bands 5, 4, and 2, respectively. The regions numbered in yellow are 1) Altiplano, 2) Cordillera Real, 3) Yungas, 4) sub-Andes. Two tributaries of the Challana River where landslides were mapped are outlined in white. The sites used for slope analysis and training areas for landslide classification are out- lined and numbered in red. The region covered by multitemporal photo- graphs is outlined in blue. The SIRC-C DEM swath is outlined in magenta. Codes labeled in white mark the locations of el Programa Hidrolo´gio y Climatolo´gico de Bolivia (PHICAB) stations that recorded sediment load measurements (Aalto et al. 2006). The codes correspond to station names found in Table 1 of Aalto et al. (2006). sub-Andes. We will further divide the Yungas into a higher elevation, high-relief zone and a lower elevation, low-relief zone. The eastern flanks of the plateau include major Late Cenozoic compressional structures thought to be important in the uplift of the plateau (e.g., Isacks 1988; Allmendinger et al. 1997; Lamb and Hoke 1997; Kley et al. 1999; McQuarrie et

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC

Fig 2 live 4/C Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 5 al. 2005). The highly simplified structural cross section in Figure 3 indicates the main geological characteristics of these physiographic zones. The high peaks of the glaciated Cordillera Real are, for the most part, Mesozoic and Cenozoic in- trusives exhumed partly in the Oligocene and partly since mid-Miocene (Barnes et al. 2006; Gillis et al. 2006, and references therein). The batholiths intrude early Paleozoic metasediments also comprising the bedrock of the Yungas. The sub-

Figure 3. Simplified structural cross section showing the relationship between to- pography, structure, and precipitation within the Beni watershed in the Eastern Cordillera (modified from Masek et al. 1994). The structural fea- tures identified are the Altiplano backthrust zone (ABT), the main Andean thrust (MAT), and the deformation front (DF).

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 6 Andean zone is a classic –thrust belt involving Paleozoic through Late Ceno- zoic rocks active since the Late Miocene (Roeder and Chamberlain 1995; Babyet al. 1993). Although the structure of the sub-Andes is fairly well known, Late Cenozoic structures of the Yungas and Cordillera Real in relation to plateau uplift and development of the sub-Andean fold–thrust belt remain poorly resolved (e.g., compare Roeder and Chamberlain 1995; Baby et al. 1993; Masek et al. 1994). Baby et al. (Baby et al. 1993) and Kley et al. (Kley et al. 1999) distinguish an “inter-Andean” zone for the southern Bolvian Eastern Cordillera, located between the sub-Andes and the higher part of the Eastern Cordillera, but the application of this differentiation to the northern Bolivian Eastern Cordillera is not so clear, as noted by Kley. Masek et al. (Masek et al. 1994), Montgomery et al. (Montgomery et al. 2001), and Barnes and Pelletier (Barnes and Pelletier 2006), among others, describe the strong northward increase in moisture flux impinging on the eastern plateau edge and the implications of this variation regarding a northern increase in erosion rates and Late Cenozoic denudation. The region of interest in this paper is located within the northern region of orographically enhanced precipitation from the large Ama- zonian moisture flux. The region includes the upland of the Beni River whose waters eventually flow into the Amazon via the Madeira River. The dominant mechanisms of erosion differ across the region. The extensively hollowed cirques and glaciated valleys in Cordillera Real, above about 3500 m, indicate that Quaternary glacial erosion has had a profound effect on the denuda- tion of that zone. In the Yungas, landslides are the dominant hillside erosional mechanism. Landsliding is common throughout the Andes, and and hydrologic conditions are the usual triggers (Blodgett et al. 1998). Landslides in the Andes range in size from the immense Huascarán debris avalanche that dev- astated Yungay, Peru, in 1970 to the numerous small landslides that frequently block transportation systems connecting the highlands to the lowlands (Ericksen et al. 1989). Landslides are especially common within the steep forested region on the northeastern flank of Bolivian Yungas. Here swarms of landslides are observed. Some key factors that contribute to the abundant landslides in the Beni basin are 1) high relief, which provides ample gravitational potential energy; 2) relatively warm and very moist conditions conducive to ; and 3) sufficient stream power to eventually remove the landslide material and to incise into bedrock.

3. Geomorphic zones and erosional processes A number of different data sources were utilized to make observations about topography and erosional processes in the area of study, including Shuttle Radar Topography Mission (SRTM) 90-m digital elevation model (DEM), aerial photo- graphs, field observations, Landsat Thematic Mapper images (TM), and a 20-m DEM derived from the National Aeronautics and Space Administration’s (NASA’s) Spaceborne Imaging Radar C (SIR-C) radar interferometry (Fielding et al. 1995). There is a large variability in slope distributions in this study area across the Eastern Cordillera. Slopes range from 0° to near vertical, with varying con- centrations of steep slopes. The region can be divided into geomorphic zones largely based on slope distributions. The following zones are labeled in Figure 4 (glaciated, high relief, and the sub-Andes).

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 7

Figure 4. Map of the distribution of slopes across the Eastern Cordillera that is de- rived from SRTM data. The relatively low-relief Altiplano and Amazon basin region are colored green. The zone of highest relief is expressed as the white region between the plateau and the orange and yellow bands of the sub-Andes, which protrude from the Amazon basin. The white region includes both the glaciated high peaks and high-relief forested and ridges of the Yungas.

3.1. Glaciated zone The steep slopes of the glaciated high peaks have a mean slope of about 30° and are at least in part a consequence of the more resistant crystalline plutons eroding more slowly relative to the surrounding metasediments (Figure 4). Similar obser- vations have been made for the boundary between crystalline units and enclosing metasediments of various metamorphic grades in the (Ni and Barazangi 1984; Kalvoda 1992; Shroder 1993; Duncan 1997). Most of the present glaciers are too small to have much significant erosional consequence, but the former glaciers were much more extensive (Klein 1997). The extent of the glaciers mainly coincides with the tree line in Figure 2. Below the high peaks, the formerly glaciated region is characterized by U-shaped valleys with gentle stream gradients and broad ridge crests with cirques cut in their flanks. The pervasive glacial erosion evident in the Cordillera Real may account for a substantial fraction of the

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC

Fig 4 live 4/C Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 8 nearly 5 km of Neogene exhumation reported by Barnes et al. (Barnes et al. 2006) in their “Eastern Cordillera” (including the Cordillera Real plus the western, higher-elevation part of our Yungas).

3.2. High-relief zone Below the glaciated zone, the topography of the Yungas is rugged with narrow floors and sharp ridge crests that have average slopes of more than 33°. The western limit of the zone is an approximate boundary between the clearly glaciated and terrain dominated by fluvial erosion without a glacial topographic signature, as determined by examination of TM images in conjunction with the SRTM. The eastern limit of the zone is delineated by the change in average slope and texture of the sub-Andes (Figure 4). The distribution of slopes in the high- relief zone is shown in Figure 5, where the peak slope is about 33°. The combination of moist and warm conditions of the forested canyons of the Yungas is favorable for weathering. Slope destabilization combined with steep slopes provides a setting for frequent landslides. Within the two watersheds out- lined in Figure 2, 5%–6% of the slopes are visibly scarred by landslides. Land- slides are not just occurring on the steepest slopes, but over a large distribution of local slopes in the high-relief zone. Figure 6 shows two histograms of slope distributions calculated from the SIR-C DEM. The magenta-colored histograms shows the distributions of slopes where landslides scars have been mapped on two sites outlined in red on Figure 2. The green-colored histograms in Figure 6 show

Figure 5. The distribution of slopes in the high-relief zone are shown where the peak slope is about 33°. Slopes are derived from SRTM data.

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 9

Figure 6. Distribution of slope angles for landslide pixels and forest pixels mapped on TM and verified with aerial photography for (a) site 1 and (b) site 2. Sites 1 and 2 are shown in Figure 2. Slopes are derived from interferometric analyses of SIR-C data. the slope distributions of the nonscarred portions of the hillslopes (forested slopes). The landslide areas were verified with aerial photographs. It is apparent both in the field and from aerial photographs that landslide failures not only occur in hollows, but they are also common on the flanks and spurs of hillslopes. The peak of slope values for both regions, between 35° and 40°, is consistent with measurements of the angle of repose for landsliding in other regions (Reneau and Dietrich 1987a; Reneau and Dietrich 1987b; Dietrich et al. 1993). The lower peak value for the larger region derived from the 90-m DEM results from the larger sampling area as well as the difference in DEM resolution, but is consistent with hillslope erosion dominated by landsliding. Field observations indicate that bedrock exposure is typical of low-order streams within the high-relief zone (Figure 7a). Because bedrock is exposed in channels in this zone, we infer that the system is weathering limited rather than transport limited. In short, all mass that is lost from the hillslopes in the high-relief zone gets transported downslope. The residence time of the mass shed to the main river channels is probably dependent on the magnitude and frequency of large floods. Hovius et al. (Hovius et al. 2000) found in bedrock-floored streamsof Taiwan that sediment load maxima correlated with peaks in water discharge closely followed by a rapid decay of sediment load, indicating the effective re- moval of most landslide deposits.

3.3. Sub-Andean zone During past periods of Andean evolution some mass entering the river systems was deposited before reaching the Amazon basin. These relict alluvial conglom-

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC

Fig 6 live 4/C Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 10

Figure 7. (a) Landslide and bedrock exposed in a Challana tributary typicalof low-order streams in high-relief zone with view downvalley. (b) Landslide and associated debris flow in high-relief zone with view upvalley from the Challana. erates are mainly located between the zone of high slopes shown in Figure 4 and the clearly delineated, predominantly northwest–southeast striking ridges and val- leys of the sub-Andes. Within this zone of relatively low relief and slopes (mean of 21°), the river systems are now cutting canyons. This strike parallel zone is identified by Strub et al. (Strub et al. 2005, and references therein) as the “internal sub-Andean belt” with a Mid- to Late Miocene fold–thrust deformation partially covered by the 8-Ma conglomeratic Cangallí Formation. The Cangallí Formation is constrained to be about 8.0 Ma (Herail et al. 1995). According to Strub et al. (Strub et al. 2005), since the Late Miocene the deformation has developed farther to the northeast in the “external sub-Andes” with inferred duplex thrust structures uplifting the Eastern Cordillera beneath and southwest of the internal sub-Andean belt. Note that the high slope zone (the upper Yungas) is coincident with the steepest regional slope (Figure 4), while the internal sub-Andean belt is located northeast of the base of this steep upper Yungas. The structural scenario of Strub et al. (Strub et al. 2005) implies that the present position of the steep part of the plateau edge in this region is not the result of erosional retreat or backcutting of the plateau edge as suggested by Isacks (Isacks 1988) and Masek et al. (Masek et al. 1994), but may

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC

Fig 7 live 4/C Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 11 have resulted from cumulative uplift of at least the northeastern side of the Alti- plano during the Neogene by duplex structures. These duplex structures, located at depth beneath the northeast plateau edge, are associated with the sub-Andean fold–thrust development. The Cangallí Formation is characterized by a finely textured stream pattern visible on both the TM image and both DEMs (Figure 2 south and west of station SRC). The Late Miocene conglomerate is presently being eroded and remobilized. This relative lower-relief zone marks the transition between erosion of Paleozoic rocks and old structures and the erosion of Tertiary rocks and newer structures. Evidence of channel incision through the surface is apparent (Blodgett 1998). Shallow landslides, typical of the high-relief zone, are commonplace in the unin- habited areas of this region. In contrast, voluminous rotational failures are more prevalent on slopes with clear-cut ridges, indicating that human disturbance may be accelerating erosion rates in this region (Blodgett 1998). The active fold and thrust belt northeast of the Cangallí Formation is unmis- takable because the long wavelength topography of troughs and ridges run parallel to the trend of the Andes that reflect the geologic structure that controls the surface topography (Figures 1 and 4). The age of the structures is post–Late Miocene (Strub et al. 2005). Most of the rivers through the parallel to the sub-Andes, but the larger rivers have cut through the structures (Figures 1 and 2). 4. Landslide mapping and erosion rates 4.1. Landslide dominant landscape shaping process Landslides mainly transport rock, soil, and vegetation from slopes into low- order channels. Because landslides are infrequently occurring phenomena, under- standing the diverse factors contributing to slope failure in a region cannot be accomplished at one site (Dietrich and Dunne 1978). Scars revegetate in the course of a few years, stabilizing the scar surface and allowing soil to thicken during the interval between slope failures. The accumulation rate is nonlinear (Lehre 1982; Shimokawa 1984; Trustrum and De Rose 1988). An analysis of dated landslide scars in New Zealand showed that soil depth increased at a rate of 3.5 mm yr−1 over the first 40 yr after slipping, but dropped to 1.2 mm yr−1 for the following 50 yr (Trustrum and De Rose 1988). The difference in soil thickening rates is not only a result of changes in bedrock weathering rates, but also because additional col- luvium is derived from surface fragmentation of exposed bedrock and crumbling scar margins during the early years after slide failure (Trustrum and De Rose 1988; Lehre 1981). After the collapse of the scar head and sidewalls, other processes such as bioturbation (especially tree throw) and soil creep continue to smooth out the scars produced by landslides and subsequent gullying (Trustrum and De Rose 1988). Although processes such as bioturbation and soil creep are important, in regions where there are many visible landslide scars, the effects of landslides outweigh the mass transport accomplished by continuous processes (Saunders and Young 1983). Erosion rates are expected to be high in the high-relief zone because of favor- able weathering conditions, downcutting streams, and steep slopes. It is therefore more important to evaluate the erosion rates in this zone. The residence time of

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 12 landslide mass in the fluvial system is unknown, but this study focuses on the first link in the transportation system, getting mass from the slopes to the channel system. If landsliding is the dominant process in such steep and forested regions, the volume of mass removed by landslides can be used to infer erosion rates. Often, however, landslides in the high-relief zone have been ob- served to only partially evacuate the colluvial mantle, and bedrock is frequently left unexposed by shallow landsliding. Evidence of post-landslide gullying on the exposed slide scar is commonplace, and in some cases may contribute a high proportion of the total mass lost from the landslide. The landslide colluvium and sediment from the gullying that reaches the steep low-order channels is later remobilized during high rainfall events, which causes debris flows to sweep the landslide colluvium from low-order to higher-order river channels (Costa and Wieczorek 1987; Selby 1993; Van Asch and Van Steijn 1991). Evidence of two debris flows were observed near the Challana field sites (Figure 7b). Although the destination of most of the landslide mass is in the low-order channels, the exposed bedrock of low-order stream channels proves that debris flows will eventually transport the mass to a higher-order channel. In the main channel, the mass will be further weathered and abraded to smaller particles. Floods will sweep the mass basinward, where by the time it reaches the foreland it has been converted almost completely to dissolved and suspended load (Dietrich and Dunne 1978).

4.2. Landslide and drainage area mapping Two relatively uninhabited heavily vegetated watersheds on the eastern flank of the Andes were chosen for study (Figure 2), where there is complete coverageof both drainage basins by aerial photographs and a TM image. Multitemporal pho- tographs and field measurements were also of paramount importance in develop- ing an erosion rate estimate. Mapping of landslides from satellite images and aerial photographs together with new topographic data have allowed remote and inaccessible portions of the Beni basin to be studied. The available images include TM images, three setsof aerial photographs (from 1964, 1975, 1993), a 25-m DEM derived from interfero- metric analyses of the SIR-C data obtained during 1994, and the 90-m DEM of the entire region available from the SRTM. In addition to serving as a verification of features identified on the TM image, the aerial photographs were used as a tem- plate for mapping landslides. The multitemporal coverage of the photographs combined with the TM provided a useful time interval over which to examine changes. The resolution of the photographs is sufficient to distinguish individual trees in most areas. Field work was also done to estimate the 3D geometry and revegetation stages of landslide scars. Excursions were made to collect measure- ments and observations in the Yungas in the years 1994–96. Landslide scars mapped within the two watersheds of the Eastern Cordillera, Bolivia, were located at elevations between about 3300 and 800 m (locations shown in Figure 2). The main rivers flowing through each of the watersheds are the Challana and an unnamed tributary of the Challana that will be referred to here as “Challana tributary.” The total area of each drainage basin and the aerial extent of landslide scars were mapped on scanned aerial photographs (Figure 8). The

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 13

Figure 8. Landslide scar distribution in the forested portions of two watersheds drain- ing into the Challana River. Areas shaded gray mark the largest regions unsuitable for mapping. Zones of relatively high landslide density are con- centrated in the high-relief zone. entire visible scar area was mapped for each slide including the runout zone. Because all area–volume measurements include the runout zone in the field mea- surements, the runout zone on aerial photographs was included. Although the runout zone seems to be a region of net gain from the slide, vegetation is usually

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 14 destroyed by runout allowing subsequent gullying and sheetwash to occur before revegetation stabilizes the slope. Any mass lost from the hillslope to a stream before revegetation is complete is considered here as volume loss associated with the landslide. Scanned aerial photographs at approximately 1:50 000 scale were used for mapping. Each photographic image was georegistered to the TM image, which had been georectified using 1:50 000 and 1:1 000 000 scale topographic maps. Land- slide mapping was only undertaken in suitable portions of the watershed. Suitable areas for mapping exclude regions above tree line, cloud covered or shadowed, and agricultural areas. Unsuitable areas represent about 15% of the watershed area. A shadowed region is defined as an area where individual trees cannot be discerned. Because very few people live below tree line in this watershed, only 1% of the watershed below tree line was mapped as anthropogenically affected. Because one of the requisites for suitability is that individual trees must be discerned, the resolution of the photographs is about 3–10 m. Although shadowed areas were removed, only a slight bias toward lower erosion rates is expected because there is no strong relationship between slope angle and landsliding in the high-relief zone (Figure 6). The percentages of landslide area for the two watersheds are 4.5 and 5.9 for the Challana and its tributary watershed, respectively.

4.3. Landslide area–volume relationship The landslide area–volume relationship used to convert from landslide area to volume and was established by Simonett (Simonett 1967) for landslides triggered by an in New Guinea. To ensure that Simonett’s curve is a good approximation for landslides in the high-relief zone, seven landslides were sur- veyed in the field using standard surveying equipment (theodolite and stadia rod). The location of the seven slides is shown in Figure 9. The theodolite station was most often located near the base of each slide. Although the objective was to place the stadia rod on the surface of the scar and the surrounding forested slope in well-distributed locations, the steepness, the amount of loose rock on the scar, and the dense undergrowth around the scar margin usually discouraged any optimal survey plan. Reaching as many parts of the slides as possible without ropes and then using ropes to help survey the gaps was the chosen strategy. The survey points were combined with field observations and photographs to estimate present and former topographic contours and to reconstruct each scar surface (Figure 10). The contours and points were used as inputs for an ARC/INFO GIS utility called Topogrid. The locations of gullied channels on the slides were used as additional constraints in the utility. A reconstruction of the original slope prior to landsliding was inferred from a best-fitted surface of the points located on the slopes adjacent to the scar. The gullying was particularly deep on landslide 6, and bedrock was exposed on landslides 6 and 7. An immense boulder projects out of the upper center of landslide 2. Although there are insufficient quantities of landslide reconstructions to estab- lish a statistically sound area–volume relationship for the Yungas, the trend of the mapped data is consistent with Simonett’s New Guinea relationship (Figure 11). The similarity of the two areas shows that the area–volume relationships are perhaps independent of the triggering mechanism, because landslides measured by

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 15

Figure 9. Locations of landsides surveyed in field. Landslides 1–6 are located within the Challana Valley. Landslide 7 is located in the Tipuani Valley.

Simonett were triggered by earthquakes while the ones measured in this study were most likely hydrologically triggered. Moreover, two of the Bolivian land- slides contained exposed bedrock on part of their surface, but their volumes and areas were consistent with Simonett’s curve. Haigh et al. (Haigh et al. 1993) also found that the outfall volumes of bedrock versus nonbedrock seeded landslides were similar in the analysis of 88 slides in the Lesser Himalayas. We conclude that Simonett’s curve adequately represents the landslide area–volume relationship in the Yungas and is used to infer the volume for each landslide area measured from aerial photographs. 4.4. Revegetation time

To determine time interval (t␷) in the erosion rate equation, a combination of multitemporal aerial photographs and dendrochronology was used. In other studies where complete aerial photograph coverage exists spanning a sufficient length of time, the time period of landslide occurrence is simply the period covered by the photographs (Hovius et al. 1997). Because the temporal resolution of aerial pho-

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC

Fig 9 live 4/C Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 16

Figure 10. Reconstructed contours of the scar surface of landslide 3 are based on survey points and field observations. Elevations are in meters above sur- vey station.

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 17

Figure 11. New Guinea slide area–volume relationship (Simonett 1967, Figure 4.2) with Bolivia data (red triangles) superimposed. tography is limited for Bolivia, scar revegetation time is used to estimate the duration of time sampled by scars on any photograph. Revegetation time is there- fore defined as the duration of time between a landslide’s occurrence and when the landslide’s scar can no longer be identified on aerial photographs.

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC

Fig 11 live 4/C Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 18 Figure 2 shows the extent of an area covered by three time periods: 1964, 1975, and 1993. Figure 12 shows the 1964 and 1993 photographs. Four fresh slides are identified on the 1964 photograph (Figure 12). All four landslide scars are estimated to be less than 3 yr old because herbs have not yet established them- selves as they have on landslides observed in the field of an older age. In the 1975 photographs all landslides are still identifiable making the revegetation time greater than 12–15 yr for this area. By 1993, one of the smaller slides was no longer recognizable, and only a small portion of another slide was visible. However, both larger slides are still discernible. The estimated revegetation time at this site near the tree line is about 30 yr for the smaller slides, but longer for the larger slides. However, the older slides are nearing the stage where it is difficult to recognize the scar boundaries. The scars would therefore no longer be recognizable and excluded from the landslide maps in less than about 4 yr. The seasonal differences between the time periods represented in Figure 12 are no- ticeable, and reproduction quality is also poor relative to the original scanned

Figure 12. Multitemporal aerial photographs showing relatively slow revegetation rate (about 29 yr) near the tree line. Photographs were taken on 6 May 1964 at the end of the wet season and on 9 July 1993 in the middle of the dry season.

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC

Fig 12 live 4/C Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 19 photographs; therefore, strict comparisons should not be made without seeing the original. Because of the distinct seasonal precipitation pattern of relatively wet summers and dry winters, tree rings develop in some of the tree species like the alder. A comparison between tree rings and aerial photographs supports the assertion that the alder’s rings are annual (Blodgett 1998). The alders have well-defined rings and are a common tree found between about 2000 m and the tree line. Alders are an invader of disturbed areas and are commonly found on previously cleared land, flood deposits, and landslide scars. Because alders are fast-growing short-lived trees of forest clearings, there is a good correlation between tree girth and age. Cores from a landslide scar exhibiting complete canopy closure at an altitude of about 3000 m in the Challana Valley indicate that canopy closure occurs prior to 14–16 yr after the trees revegetated the upper scarp. However, a correction needs to be made, because saplings do not reach the coring height until after 2–3 yr. Slides 1–5 in Figure 9, for example, are all younger than the July 1993 aerial photograph and only a few saplings were just approaching coring height in the August field season in 1996. If 2–3 yr are added as a correction for the initial establishment of trees, then closure on the 3000-m-elevation scar is estimated to have occurred about 16–19 yr after the landslide failed. From the observations of dated young landslides, alder seedlings on the toeof the slide scars tend to establish themselves faster than alder seedlings on the upper scar face. Water supply during the dry season and nutrients are probably the major factors influencing revegetation rates (Dalling and Tanner 1995). Because it is important to estimate the fastest revegetation times possible, we sampled in several optimal locations for landslide revegetation. Four sites were sampled at elevations below 3000 m. In each of the following four sites, conditions are in theory more favorable for revegetation than the upper scar slope because nutrients and water supply are more readily available. Two sites were located on alluvium and showed corrected canopy closure at about 8–11 yr. Two other sites located on the toe of a landslide scar showed corrected closure rates of about 11–14 yr. Optimal revegetation times at lower elevations are estimated to be 11–14 yr. Upper scarp measurements (poor revegetation conditions) yield a range of 16–19 yr. The average of the two types of conditions at lower elevations is 13.5–16.5 yr, while the revegetation time for higher elevations was estimated to be 30–35 yr. Higher elevations are likely to require more time to revegetate because of greater aridity and generally colder conditions. Because most of the high-relief zone lies well below the tree line, the actual revegetation time is probably closer to the lower elevation range of revegetation time, but a conservative range of 10–35 yr is used here to calculate erosion rates.

4.5. Volume delivery factor and landslide erosion rate estimates for two watersheds All landslide volume reaching a stream is considered “delivered” from the hillslope to the streams if eventually remove the material, even if the removal occurs decades later. To determine the landslide volume delivery factor (Fd), “fresh” landslides on two aerial photographs were examined to deter- mine which reached streams and which remained on the hillslopes. Only relatively

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 20 fresh slides were counted to avoid the possibility of revegetation isolating the landslide. About 90% of the slides reach a stream making the delivery factor 0.9; nevertheless, volume delivery for certain cases is unclear. Landslides with long runout may lose more volume on route. On the other hand, runout can cause subsequent gullying that can increase volume lost to the streams. The landslides measured in the field all had nearly 100% delivery, and toes near streams may lose even more volume over time because of bank erosion, which was not considered when measuring the landslide area-to-volume ratios. However, there is also a possibility that the initial landslide toe may be lost in a subsequent landslide and should not in that case be counted as delivered to the streams until then. Given the complexity of these considerations, we retain the initial delivery factor of 0.9 until more comprehensive studies of landslide delivery can be made. We note that halving the delivery ratio would halve the erosion rate.

4.6. Landslide erosion rate equation We developed the following equation for determining erosion rates from land- ס slides: Re Vs/(Att␷), where Re is the erosion rate, Vs is the volume of landslide sediment, At is the mapable area of the drainage basin, and t␷ is the time interval over which the volumetric contribution of landslides is measured. The landslide sediment volume is determined by summing the volumes of all the individual ⌺␷ ס landslides that reach the stream network and can be expressed as Vs Fd s, ␷ where Fd is the landslide volume delivery factor and s is the volume of each landslide scar. Landslide scar volume is derived from the planimetric measurement of landslide scar area converted to volume using an area–volume relationship established by Simonett (Simonett 1967) in New Guinea and field checked in Bolivia. Because a portion of the landslides do not appear from aerial photographs to reach a stream, the correction Fd is applied to account for the landslide volume that remains on the slopes until the next landslide occurs in the same location. The bulk of the colluvium delivered to the streams is probably removed by debris flows and/or large flood events. Evidence of alluvial processes is also visible on aerial photographs and satellite images, but bank erosion and alluvium 35–10 ס were excluded from the landslide mapping. With a revegetation time T␷ the erosion rate estimate is 10±5mmyr−1 ,90% ס yr and a delivery factor of F for the Challana and8±4mmyr−1 for the tributary basin. The average of these two estimates is9±5mmyr−1.

4.7. Extrapolation to upper Beni basin region An alternative to mapping on aerial photographs is to classify satellite images because large datasets are needed to account for spatial and temporal heterogene- ity. Two types of classification methods, a decision tree classifier (DTC) and a maximum likelihood classifier (MLC), were applied to a Thematic Mapper image. Precise details of the classification procedures are described in Blodgett (1998). With the help of aerial photographs for verification, the following feature types were sampled on TM image 001_71, which was acquired in 1987: landslide scars, mature forest, paramo, debris flow, and fresh burn. Landslide scars were then subdivided and sorted based on the extent and type of revegetation associated with

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 21 the age of the scar. Forest samples were taken in different geographical regions to capture the heterogeneity of the cloud forest. Figure 13 shows the mean spectral reflectances of several age classes of land- slides and mature forest. Band 4 shows the most prominent spectral differences between landslide classes, but bands 3, 5, and 7 are more useful for discriminating vegetation from landslides. A quantitative study that links the age of the scar to specific spectral characteristics has not been accomplished, but a numberofob- servations indicate that a succession of plant communities is the likely cause for the observed spectral differences. The spectra of the fresh landslide class resemble alluvium in stream channels. The “juvenile” scar corresponds to a mixture of herbs

Figure 13. Spectral reflectance curves of landslide age classes and cloud forest constructed by determining the mean and standard deviation of reflec- tance values sampled from 6 TM bands. Cloudforest 1 was sampled from the high-relief zone in the Challana drainage basin.

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC

Fig 13 live 4/C Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 22 and saplings, which are the colonizers of the bare soil. Usually ferns and bamboo grasses become the dominant species after 2–3 yr, which is expressed as a higher reflectance of band 4 on older landslide scars. A comparison of the spectral signature of old landslide scars with mature forest indicates that the reflectance of band 4 actually decreases as forest canopy closure occurs, which is not optimal for the normalized difference vegetation index (NDVI) classifier that is frequently used for classifying vegetation on TM (Wolter et al. 1995). Empirical analysis shows that the greatest difference is between bands 5 and 7, one of the ratios applied in the DTC (Figure 13) Without knowing the size of the entire landslide scar, it is difficult to associate a landslide volume with any given landslide pixel. For this reason, the goalofthe classifiers are to compare a classified image of the Challana drainage basins already mapped to a much larger territory to see if the ratio of landslide scars to forest is consistent or different. Two types of classifiers (DTC and MLC) were applied to ensure that the results are not unique to one type of classifier. The DTC serves as the conservative classifier. Only the most distinct landslides will be counted as landslides. Restrictions will be loosened for the MLC, allowingitto include more pixels as landslides. All streams, agricultural areas, regions above tree line, and clouds were clipped out. Shadowed areas were also removed using a band 5 threshold. Although the DTC underestimates the landslide scar area, the underestimate is consistent enough to calibrate the classified image. The Challana watershed mapped from aerial photographs detected landslides on 4.2% of the watershed, while the classified image was only able to detect 0.7% of the landslides, which is about 17% less than the manual mapping. Classification of the other watershed, the tributary watershed, yielded a 1.2% landslide area, which is about 20% less than what was mapped. The factors of 4.9 and 6 were applied to the classified image of a much larger dataset that includes all of the high-relief zones on the TM image 001_71 that have not been disturbed by agriculture. After calibrating the raw results with the Chal- lana and tributary watershed estimates, a landslide area of 5.1%–6.2% for the entire high-relief zone was determined. Because this estimate is similar to the estimate for the two individual watersheds of 4.2% and 5.1%, the erosion rates estimated for the two watersheds apply to the whole high-relief zone on the TM image. Results for MLC produced the following landslide scar-to-forest ratios: Chal- lana (6.3%), tributary (9.6%), and whole high-relief zone (7.6%). After applying the factors of 0.67 and 0.53, the entire high-relief zone has a landslide scar-to- forest ratio of 4.0%–5.1%, which is almost identical to the Challana landslide map estimates. The MLC estimate further supports the validity of using the landslide erosion rates of9±5mmyr−1 for the whole high-relief zone.

5. Discussion The main factors determining the distribution of landslides within the two regions is unknown, but there are some likely candidates. Prolonged, intense rainfall events in many tropical climates are the triggers for most landslide occur- rences (Thomas 1994). If the landslides were hydrologically triggered, then the

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 23 distribution would reflect those places where pore-water pressure was sufficiently high to overcome the soil’s shear strength. However, a second factor is also important. Several studies have indicated that a threshold in soil/colluvium thick- ness must be attained before a landslide can occur (Reneau and Dietrich 1987a; Crozier et al. 1990); therefore, a period of time must lapse before a landslide will occur again in the same location. This factor has even been incorporated into several shallow-landsliding models (Dietrich et al. 1995; Okimura and Kawatani 1987; Okimura 1989). This time-dependent threshold has been attributed to the ability of tree roots to penetrate into bedrock and anchor the soil (Dietrich and Dunne 1978; Reneau and Dietrich 1987a). Roots also influence landslide size. Because tree roots have strength, small slide scarps in the forest have a difficult time overcoming the cohesive strengthof arboreal root mats. Small landslides are therefore expected to be rare. The distribution of landslide scar areas displayed in Figure 14 support this hypothesis. The fractal breakdown is illustrated by the distribution of landslides plotted on a cumulative area versus size plot (Pelletier et al. 1997). The areas sorted by size are clearly not fractal down to the smallest scales resolvable (Figure 14). The resolution of the photograph is adequate to detect landslides with sizes well below the break in slope shown in Figure 14, because even individual trees

Figure 14. Plot of the individual landslide area as a function of the cumulative number of landslide scars from the Challana Valley, Bolivia (modified from Pelletier et al. 1997).

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 24 can be discerned across much of the photograph. Notice that the number of landslides smaller than about 4000 m2 is not sufficient to maintain the power-law distribution. Based on landslide mapping and TM image classification, erosion rates in the high relief of the upper Beni basin are estimated to be about9±5mmyr−1. Erosion rates in Southern Alps of New Zealand have yielded similar erosion rate estimates of9±4mmyr−1 when landslide area maps were used as the method for determining erosion rates (Hovius et al. 1997). Despite the similarity between these erosion rates, the Andean landslide erosion estimates exceed most previous estimates in the Andes by almost a factor of 10 (Barnes and Pelletier 2006). One possible reason for the discrepancy is that our Andean landslide rate is dependent on a volume delivery factor of 0.9, which could be overestimated as discussed earlier. Nevertheless, some differences are expected between the various methods. The landslide study focuses on the region with the highest relief. The study appears to integrate over enough time (decades) to capture the big events on the hillslopes that are doing most of the work. In contrast, sediment flux studies in the Andes tend to integrate over larger areas, including the lower-relief and gla- ciated areas, and may potentially miss the large flood events that may occur less frequently. Fission track studies integrate over long time scales at each location where samples are collected but, as mentioned previously, the narrow band of intense erosion is not expected to remain stationary over these time scales. Climate and glacial cycles are also changing significantly over time. Therefore, at any given location where fission track samples are collected, the rate of erosion would include slow periods too. Cosmogenic radionucleides measuring rates over 400–15 ka should be the most compatible with the landslide rates, but the rates are still integrated over the whole basin including low-relief zones above and below the high-relief zones. Other methods used to measure erosion rates in the Central Andes mainly apply to the drier region of the south where slower rates are ex- pected (Barnes and Pelletier 2006). A possible cause for such a high landslide erosion rate is that the weather was unusually wet during the years sampled by the landslides. More landslides would be expected under wetter conditions. Figure 15 shows the relationship between El Niño events and the level of Lake Titicaca. Because the origin of Lake Titicaca’s moisture is from the same source as the upper Beni basin (Zhao and Lau 1998), the lake levels are used here as a proxy for precipitation. El Niño events are most often associated with lower lake levels and therefore less precipitation (Rowe and Dun- 2004). The El Niño records over the last 467 yr indicate that moderate, severe, and very severe El Niño events occur about 17% of the time (Quinn et al. 1987). Between 1973 and 1993, there were moderate and stronger El Niño events during 20% of the time and during 17% of the time over 30 yr prior to 1993. Therefore, the number of strong and very strong El Niño events was typical during either of those time periods. Based on the El Niño records, the number of landslides gen- erated over the time period sampled is not likely to be unusually high because of atypical weather patterns. Guyot (Guyot 1992) determined denudation rates from suspended load to be 1.4 mm yr−1 for the Beni River and 1.3 mm yr−1 for all the Andean rivers in Bolivia that drain into the Amazon. The sediment load measurements collected in the Amazon basin underestimate the erosion occurring in the mountains because they

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 25

Figure 15. Lake Titicaca level and its relationship to El Niño events. The relative strength of El Niño events are indicated as follows: M = moderate, S = strong, VS = very strong, and the plus symbol (+) = slightly stronger than others of the same category. are integrating over both high- and low-relief portions of the basin. The denudation rate of 1.4 mm yr−1 for the Beni basin was determined from the station Angosta de Bala near Rurrenabaque, where the suspended load in the Beni was measured from 1983 to 1989 at an elevation of 280 m. The best comparison between sediment load measurements and the landslide erosion rate estimates is made by comparing measurements at the same elevation and in the same catchment, but there are no daily records of sediment load for the Challana River yet. Neverthe- less, it is worthwhile comparing the sediment load measurements in the high-relief zone farther south in the Beni with the landslide erosion rate estimates. Table 1 lists the erosion rates based on sediment load measurements (Aalto et al. 2006). The measurements at station CAJ yield an erosion rate of 6.9 mm yr−1, which is within the range of the landslide erosion estimates. However, manyofthe rates at other stations in the high-relief zone were calculated to be much slower: SIR (TAM) at 2.8 mm yr−1 (0.95 mm yr−1). The extreme variability of erosion rates between catchments in the high-relief zone is an indication that measure- ments have not been integrated over a long enough time period. A period of a decade or less is probably insufficient to capture the larger flood events, thus severely underestimating the average suspended load carried by the river over hundreds of years.

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 26 Table 1. Erosion rates determined from sediment load measurements (Aalto et al. 2006). Station Elevation (m) Erosion rate (mm yr−1) SIR 1640 2.8 TAM 1185 9.5 × 10–1 VBA 1050 1.6 CAJ 760 6.9 AQM 500 1.5 SRC 440 5.7 × 10–1 AIN 400 1.4

Fission track measurements have determined that ∼10 km have been eroded off the upper Beni slope over the past ∼15 m.y., which corresponds to exhumation rates of about 0.7 mm yr−1 (Barnes et al. 2006). The disparity between our mea- surements and the fission track estimates is perhaps because the erosion front (the high-relief zone) may not have remained in precisely the same geographic location during the last 10–15 m.y. period relative to the rock sampled. This is already true of the fission track samples collected from the glaciated zone where the current erosion rate is slow (Guyot 1992). A second explanation is that the fast erosion rates of the humid present are atypical of the Neogene. Paleoclimate evidence indicates much drier periods during the Late Pleistocene (Clapperton 1993). It is conceivable that long dry periods also were the norm during the Neogene, which may have reduced the long-term erosion rates. Cosmogenic radionuclides, measured in river sediment in the upper Beni, indi- cated erosion rates of 0.04–1.35 mm yr−1 over time scales of 400–15 ka (Safran et al. 2005). However, these measurements assume that sediment delivered to chan- nels becomes rapidly mixed and that long-term valley storage is minimal (Safran et al. 2005). We have noted several examples of long-term valley storage that may impact these results. Examples include remnants of glacial outwash terraces such as those in the headwaters of the Consata River near the city of Sorata, landslides and debris flows that linger in stream channels unmixed for an undetermined number of years, and the Late Miocene Cangallí conglomerate that is presently being eroded and remobilized. Again, these rates derived from cosmogenic radio- nuclides integrate over both high- and low-relief portions of the basin, which may also explain the slower rates relative to the landslide study.

6. Conclusions In the tectonically active and relatively moist flank of the Eastern Cordillera in the Central Andes, erosion rates are high. Glaciation, especially in the Late Pleis- tocene is the dominant erosion mechanism in the uplands whereas landsliding is dominant in the Yungas downstream. In the landslide-dominated high-relief zone of the Yungas, landslides occur not only in hollows but on also on the spurs and flanks of hillslopes. However, landslides occur more frequently on steeper hillslopes. About 90% of the landslide scars mapped on aerial photographs reach streams. Because the streams have bedrock channels in many places, the fluvial system is considered weathering

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 27 limited. Based on dendrochronology and multitemporal images, the time for land- slide scars to revegetate is relatively rapid at 10–35 yr. The landslide area and volume measurements of seven landslide scars are consistent with Simonett’s curve regardless of whether or not bedrock was exposed on the scar surface. From landslide area maps, Simonett’s area–volume relationship, revegetation time, and the percentage of landslide material reaching the streams, the average erosion rate for the Challana River and one of its tributaries is calculatedtobe 9±5mmyr−1. Satellite image classifications show that the landslide erosion rate is not only applicable to the two watersheds, but also applies to the rest of the high-relief zones of the upper Beni. The erosion rates determined from this method are consistent with landslide erosion estimates in the New Zealand Southern Alps and with some sediment load measurements collected in the Beni headwaters.

Acknowledgments. This project was funded by grants from GSA and NASA-topog- raphy, and NASA (EOS) Contract NAGW-2638. We thank the following people for help in the field: Brandel Rock; Jeff Masek; Juanita Mcgarrigle; Richard Sherry; Seth Blodgett; Clifford Blizard; Andrew Warner; and especially Maximo Pilco Quispe, our guide from Pablo Amaya, Bolivia. Eric Fielding processed the SIRC-C radar data. We thank Chris Duncan and Shane Deitweiler for helping analyze various topographic data. We appreciate the helpful comments and suggestions from Andrew Klein, Bruce Malamud, Jeff Masek, Liz Safran, Tom Dunne, Art Bloom, and the reviewers.

References Aalto, R., T. Dunne, and J. L. Guyot, 2006: Geomorphic controls on Andean denudation rates. J. Geol., 114, 85–99. Adams, J., 1980: Contemporary uplift and erosion of the Southern Alps, New Zealand. Bull. Geol. Soc. Amer., 91, 1–114. Allmendinger, R. W., T. E. Jordan, S. M. Kay, and B. L. Isacks, 1997: The evolution of the Altiplano-Puna plateau of the Central Andes. Annu. Rev. Earth Planet. Sci., 25, 139–174. Anders, M. H., K. M. Gregory-Wodziki, and M. Spiegelman, 2002: A critical evaluation of late Tertiary accelerated uplift rates for the Eastern Cordillera, central Andes of Bolivia. J. Geol., 110, 89–100. Baby, P., B. Guillier, J. Oller, and G. Montemurro, 1993: Modele cinematique de la zone subandine du coude de Santa Cruz (entre 16 degrees S et 19 degrees S. Bolivie) deduit de la construction de cartes equilibrees. [Kinematic model of the subandean zone of the Santa Cruz Bend (between 16 degrees S and 19 degrees S. Bolivia) deduced from balanced mapping]. C. R. Acad. Sci., Ser. 2, 317, 1477–1483. Barnes, J. B., and J. D. Pelletier, 2006: Latitudinal variation of denudation in the evolution of the Bolivian Andes. Amer. J. Sci., 306, 1–31. ——, T. A. Ehlers, N. McQuarrie, P. B. O’Sullivan, and J. D. Pelletier, 2006: Eocene to recent variations in erosion across the central Andean fold-thrust belt, northern Bolivia: Implica- tions for plateau evolution. Earth Planet. Sci. Lett., 248, 118–133. Benjamin, M. T., N. M. Johnson, and C. W. Naeser, 1987: Recent rapid uplift in the Bolivian Andes: Evidence from fission-track dating. Geology, 15, 680–683. Blodgett, T., 1998: Erosion rates on the NE escarpment of the Eastern Cordillera, Bolivia derived from aerial photographs and thematic mapper images. Ph.D. thesis, Cornell University, 130 pp. ——, C. Blizard, and B. Isacks, 1998: Andean landslide hazards. Geomorphological Hazards in High Mountainous Areas, Kluwer Academic, 211–227.

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 28 Clapperton, C., 1993: Quaternary Geology and of South America. Elsevier Sci- ence Publishers, 779 pp. Costa, J. E., and G. F. Wieczorek, Eds., 1987: Debris Flows/Avalanches: Process, Recognition, and Mitigation. Geological Society of America, 239 pp. Crozier, M. J., E. E. Vaughan, and J. M. Tippett, 1990: Relative instability of colluvium-filled bedrock depressions. Earth Surf. Processes Landforms, 15, 329–339. Dalling, J. W., and V. J. Tanner, 1995: An experimental study of regeneration on landslides in montane rain forest in Jamaica. Ecology, 83, 55–64. Dietrich, W. E., and T. Dunne, 1978: Sediment budget for a small catchment in mountainous terrain. Z. Geomorph., 29 (Suppl. B), 191–206. ——, C. J. Wilson, D. R. Mongomery, and J. McKean, 1993: Analysis of erosion thresholds, channel networks and landscape morphology using a digital terrain model. J. Geol., 101, 161–180. ——, R. Reiss, M. Hsu, and D. R. Montgomery, 1995: A process-based model for colluvial soil depth and shallow landsliding using digital elevation data. Hydrol. Processes, 9, 383–400. Duncan, C., 1997: Tectonics, topography, climate and erosion: Analysis of Himalayan digital elevation data and numerical modeling. Ph.D. thesis, Cornell University, 200 pp. Ericksen, G. E., C. F. Ramirez, J. F. Concha, T. M. Gilberto, and U. B. Fernando, 1989: Landslide hazards in the Central and Southern Andes. Landslides: Extent and Economic Significance, E. E. Brabb, and B. L. Harrold, Eds., Balkema, 111–117. Fielding, E. J., C. L. Werner, T. G. Farr, and G. F. Peltzer, 1995: High-resolution DEMs of the Altyn Tagh in the Kunlun of Tibet. Abstract, Geologic Society of America Annual Meetings, New Orleans, LA, Abstr. No. 15642. Gillis, R., B. K. Horton, and M. Grove, 2006: Thermochronology, geochronology, and upper crustal structure of the Cordillera Real: Implications for Cenozoic exhumation of the central Andean plateau. Tectonics, 25, 1–22. Guyot, J. L., 1992: Hydrogéochemimie de fleuves de l’Amazonie bolivienne. Ph.D. dissertation, Université de Bordeaux, 362 pp. ——, J. Bourges, R. Hoorelbecke, H. Calle, J. Cortes, and M. C. Barragan Guzman, 1988: Sediment discharge from the Andes to the Amazonia along the Beni River. Sediment Bud- gets, IAHS Publ. 174, 443–451. ——, J. M. Jouanneau, J. Quintanilla, and J. G. Wasson, 1993: Dissolved and suspended sediment loads exported from the Andes by the Beni river (Bolivian Amazonia), during a flood. Geodyn. Acta, 6, 233–241. Haigh, M. J., J. S. Rawat, S. K. Bartarya, and M. S. Rawat, 1993: Environmental influences on landslide activity: Almora bypass, Kumaun Lesser Himalaya. Nat. Hazards, 8, 153–170. Herail, G., W. Sharp, G. Viscarra, and M. Fornari, 1995: La edad del la formacion Cangallí: Nuevos datos geocronologicos y su significado geologico. Memorias del XI Congreso geo- logico de Bolivia, Santa Cruz de la Sierra, Bolivia, Sociedad Geologica de Bolivia, 361–365. Horton, B. K., 1999: Erosional control on the geometry and kinematics of thrust belt development in the central Andes. Tectonics, 18, 1292–1304. Hovius, N., C. P. Stark, and P. A. Allen, 1997: Sediment flux from a mountain belt derived by landslide mapping. Geology, 25, 231–234. ——, ——, C. H. Chu, and L. Jiun-Chuan, 2000: Supply and removal of sediment in a landslide- dominated mountain belt: Central Range, Taiwan. J. Geol., 108, 73–89. Isacks, B. L., 1988: Uplift of the central Andean plateau and bending of the Bolivian orocline. J. Geophys. Res., 93, 3211–3231. Kalvoda, J., 1992: Geomorphological Record of the Quaternary in the Himalaya and the Karakoram. Elsevier, 315 pp. Klein, A. G., 1997: Modern and late Pleistocene glacial studies in the central Andes of Peru and Bolivia. Ph.D. dissertation, Cornell University, 191 pp.

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 29 Kley, J., C. R. Monaldi, and J. A. Salfity, 1999: Along-strike segmentation of the Andean foreland: Causes and consequences. Tectonophysics, 301, 75–94. Lamb, S., and L. Hoke, 1997: Origin of the high plateau in the Central Andes, Bolivia, South America. Tectonics, 16, 623–649. Lehre, A., 1981: Sediment budget of a small California coast range drainage basin near San Francisco. Erosion and in Pacific Rim Steeplands, IAHS Publ. 132, 123–139. ——, 1982: Sediment mobilization and production from a small mountain catchment: Lone Tree Creek, Marin County, California. Ph.D. dissertation, University of California, Berkeley, 235 pp. Masek, J. G., B. L. Isacks, T. L. Gubbels, and E. J. Fielding, 1994: Erosion and tectonics at the margins of continental plateaus. J. Geophys. Res., 99, 13 941–13 956. McQuarrie, N., 2002: The kinematic history of the central Andean fold–thrust belt, Bolivia: Implications for building a high plateau. Geol. Soc. Amer. Bull., 114, 950–963. ——, B. K. Horton, G. Zandt, S. Beck, and P. G. DeCelles, 2005: Lithospheric evolution of the Andean fold-thrust belt, Bolivia, and the origin of the central Andean Plateau. Tectonophys- ics, 399, 15–37. Montgomery, D. R., G. Balco, and S. D. Willet, 2001: Climate, tectonics, and the morphology of the Andes. Geology, 29, 579–582. Moore, M., and P. C. England, 2001: On the inference of denudation rates from cooling ages of minerals. Earth Planet. Sci. Lett., 185, 265–284. Ni, J., and M. Barazangi, 1984: Seismotectonics of the Himalayan collision zone: Geometry of the underthrusting Indian Plate beneath the Himalaya. J. Geophys. Res., 89, 1147–1164. Okimura, T., 1989: Predication of slope failure using the estimated depth of the potential failure layer. J. Nat. Disaster Sci., 11, 67–79. ——, and T. Kawatani, 1987: Mapping of the potential surface-failure sites on granite slopes. International Geomorphology 1986, Part I, V. Gardiner, Ed., Wiley, 121–138. Pelletier, J., B. Malamud, T. Blodgett, and D. Turcotte, 1997: Scale-invariance of soil moisture variability and the frequency-size distribution of landslides. Eng. Geol., 48, 255–268. Quinn, W. H., V. T. Neal, and S. E. Antunez de Mayolo, 1987: El Niño occurrences over the past four and a half centuries. J. Geophys. Res., 92C, 14 449–14 461. Reneau, S. L., and W. E. Dietrich, 1987a: Size and location of colluvial landslides in a steep forested landscape. Int. Assoc. Hydrol. Sci. Bull. Publ., 165, 39–49. ——, and ——, 1987b: The importance of hollows in debris flow studies; examples from Marin County, California. Debris Flows/Avalanches: Process, Recognition and Mitigation, J. E. Costa and G. F. Wiezcorek, Eds., Vol. VII, Reviews in Engineering Geology, Geological Society of America, 165–180. Roeder, D., and R. L. Chamberlain, 1995: Structural geology of sub-Andean fold and thrust belt in northwestern Bolivia. Petroleum Basins of South America: AAPG Memoir 62, A. J. Tankard et al., Eds., American Association of Petroleum Geologists, 459–479. Rowe, H. D., and R. B. Dunbar, 2004: Hydrologic-energy balance constraints on the Holocene lake-level history of Lake Titicaca, South America. Climate Dyn., 23, 439–454. Safran, E. B., 1998: Channel network incision and patterns of mountain geomorphology. Ph.D. thesis, University of California, Santa Barbara, California, 327 pp. ——, P. R. Bierman, R. Aalto, T. Dunne, K. X. Whipple, and M. Caffee, 2005: Erosion rates driven by channel network incision in the Bolivian Andes. Earth Surf. Processes Landforms, 30, 1007–1024. Saunders, I., and A. Young, 1983: Rates of surface processes on slopes, slope retreat and denu- dation. Earth Surf. Processes Landforms, 8, 473–501. Selby, M., 1993: Hillslope Materials and Processes. Oxford University Press, 289 pp. Shimokawa, E., 1984: Natural recovery process of vegetation on landslide scars and landslide

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC Earth Interactions • Volume 11 (2007) • Paper No. 19 • Page 30 periodicity in forested drainage basins. Proc. Symp. on the Effects on Forest Land Use on Erosion and Slope Stability, Honolulu, HI, University of Hawaii, 99–107. Shroder, J. F., Jr., 1993: Himalaya to the sea: Geomorphology and the Quaternary of Pakistan in the regional context. Himalaya to the Sea: Geology, Geomorphology, and the Quaternary, J. F. Schroder, Jr., Ed., Routledge, 1–42. Simonett, D. A., 1967: Landslide distribution and earthquakes in the Bewani and Torricelli Moun- tains, New Guinea. Landform Studies from Australia and New Guinea, J. N. Jennings and J. A. Mabbutt, Eds., Cambridge University Press, 64–84. Strub, M., G. Hérail, J. Darrozes, R. García-Duarte, and G. Astorga, 2005: Neogene to Present tectonic and orographic evolution of the Beni Subandean Zone. Extended Abstracts, Sixth Int. Symp. on Andean Geodynamics, Barcelona, Spain, Institut de Recherche pour le De´veloppement, Paris, 709–713. Summerfield, M. A., and N. J. Hulton, 1994: Natural controls of fluvial denudation in major world drainage basins. J. Geophys. Res., 99, 13 871–13 884. Suppe, J., 1981: Mechanics of mountain building and metamorphism in Taiwan. Mem. Geol. Soc. China, 4, 67–89. Thomas, M. F., 1994: Geomorphology in the Tropics. John Wiley and Sons, 460 pp. Trustrum, N. A., and R. C. De Rose, 1988: Soil depth-age relationship of landslides on deforested hillslopes, Taranaki, New Zealand. Geomorphology, 143–160. Van Asch, T. W. J., and H. Van Steijn, 1991: Temporal patterns of mass movements in the French Alps. Catena, 18, 515–527. Wolter, P. T., D. J. Mladenoff, G. E. Host, and T. C. Crow, 1995: Improved forest classification in the northern lake states using multi-temporal Landsat imagery. Photogramm. Eng. Remote Sens., 61, 1129–1143. Zhao, J., and K.-M. Lau, 1998: Does a monsoon climate exist over South America? J. Climate, 11, 1020–1040.

Earth Interactions is published jointly by the American Meteorological Society, the American Geophysical Union, and the Association of American Geographers. Permission to use figures, tables, and brief excerpts from this journal in scientific and educational works is hereby granted provided that the source is acknowl- edged. Any use of material in this journal that is determined to be “fair use” under Section 107 or that satisfies the conditions specified in Section 108 of the U.S. Copyright Law (17 USC, as revised by P.IL. 94-553) does not require the publishers’ permission. For permission for any other form of copying, contact one of the copublishing societies.

Unauthenticated | Downloaded 10/06/21 03:24 AM UTC