You are on page 1of 12

INTERNATIONAL JOURNAL OF CLIMATOLOGY Int. J. Climatol. 23: 14531464 (2003) Published online 6 August 2003 in Wiley InterScience (www.interscience.wiley.com).

DOI: 10.1002/joc.938

THE CENTRAL ANDEAN WEST-SLOPE RAINSHADOW AND ITS POTENTIAL CONTRIBUTION TO THE ORIGIN OF HYPER-ARIDITY IN THE ATACAMA DESERT
JOHN HOUSTONa, * and ADRIAN J. HARTLEYb Nazca S.A., Avenida Las Condes 10373, Ocina 60, Santiago, Chile Department of Geology and Petroleum Geology, University of Aberdeen, Aberdeen AB24 3UE, UK
a

Received 31 January 2003 Revised 19 May 2003 Accepted 22 May 2003

ABSTRACT The west slope of the central Andes exhibits a pronounced rainshadow effect. Precipitation between 15 and 27 S is dominated by summer convective activity from Amazonia, and data analysis shows that the increase in precipitation with elevation due to the rainshadow effect best ts an exponential correlation. Coupling with limited data from high elevations suggests that the correlation is accurate to 4500 m above sea level (m a.s.l.) and perhaps to 5500 m a.s.l., suggesting that increased precipitation goes unrecorded over the peaks of the western Cordillera. South of 27 S the precipitation is dominated by winter frontal sources and shows no well-dened relationship with elevation. The core zone of hyper-aridity in the Atacama Desert extends from 15 to 30 S at elevations from sea level to 3500 m a.s.l. Although the Atacama Desert has existed since at least 90 Ma, it is considered that the initial onset of hyper-aridity was most likely to have developed progressively with the uplift of the Andes as they reached elevations between 1000 to 2000 m a.s.l. coupled with the intensication of a cold, upwelling Peruvian Current between 15 and 10 Ma. Also apparent in the palaeogeographic record are subsequent uctuations between (semi-) arid to hyper-arid conditions that were probably largely controlled by changes in orbital and oceanic forcing. Copyright 2003 Royal Meteorological Society.
KEY WORDS:

precipitation; rainshadow; hyper-aridity; palaeoclimate; Atacama Desert; Chile

1. INTRODUCTION The Atacama Desert lies on the west slope of the central Andes between 15 and 30 S at elevations between sea level and 3500 m a.s.l., at the core of the South American dry diagonal, which extends from 5 S on the west coast to nearly 50 S on the east coast, over 4000 km with less than 200 mm mean annual rainfall (MAR) (Figure 1). The controls on moisture availability in the core area of the Atacama Desert in northern Chile are of paramount importance to the water resources, and hence to the economic development of the area. Yet, although considerable recent work has focused on the current (e.g. Markgraf, 2001; Garreaud et al., 2003) and Late Pleistocene to Holocene palaeoclimate (e.g. Betancourt et al., 2001), there is still little quantitative data on spatial variability and the origin of hyper-aridity. Following a review of the controls on precipitation, the west slope precipitationelevation effects are analysed and the possible origins of hyper-aridity are reviewed and discussed. Hyper-aridity is not solely due to a lack of rainfall (P ); it is also related to potential evapotranspiration (PET: the maximum evapotranspiration rate given no soil moisture decit) and may be dened as a ratio of P /PET of less than 0.05 (UNEP, 2001). By this denition, virtually the whole of the area between 15 and
* Correspondence to: John Houston, Nazca S.A., Avenida Las Condes 10373, Ocina 60, Santiago, Chile; e-mail: houston@entelchile.net

Copyright 2003 Royal Meteorological Society

1454

J. HOUSTON AND A. J. HARTLEY


70W 50W

Quelccaya
15S

e Lin

of

on cti se

g Fi

Sajama

Iquique
25S

El Laco

35S

45S

Hyper-Arid zone Precipitation < 200 mm/a

Figure 1. Location map of hyper-arid South America showing highest stations where precipitation data is available

30 S at elevations from sea level to 3500 m a.s.l. may be considered as hyper-arid. Nevertheless, within this area there are signicant spatial and temporal variations in intensity.

2. REGIONAL CLIMATE AND CONTROLS ON PRECIPITATION Four factors contribute to the aridity of the Atacama Desert. Firstly, it lies in the subtropical high-pressure belt where descending stable air produced by the Hadley circulation signicantly reduces convection and hence precipitation (zonal effect). Secondly, it lies at a considerable distance (up to 2000 km) from the AmazoniaAtlantic moisture source (continentality effect). Thirdly, the proximity of the Andes upwind prevents moisture advection from the east (rainshadow effect). Finally, the cold Peruvian Current (PC), which upwells along the Atacama coast, inhibits the moisture capacity of onshore winds and creates a persistent inversion that traps any moisture below 800 m a.s.l. (the oceanic effect). The impact of the high Andes on the climate of the Atacama Desert is critical. Regional uplifts such as the Andes have been shown to: (a) disrupt zonal circulation, thus creating a barrier to global winds that is particularly acute where the barrier is perpendicular to the zonal ow, creating orographic precipitation on the windward side and a rainshadow on the leeward side (Manabe and Broccoli, 1990; Basist et al., 1994, Figueroa et al., 1995); (b) alter the radiation balance as a result of increased insolation at high elevations, thus increasing evaporation (Hay and Wold, 1998); and (c) displace air masses, modifying and intensifying pressure contrasts, frequently leading to monsoonal circulation (Hay et al., 1990; Zhou and Lau, 1998), all of which effects are exacerbated at low latitudes. What moisture does arrive in the Atacama Desert originates from three sources. During the austral summer, northeasterly (monsoonal) airows bring convective precipitation from Amazonia (Garreaud, 1999; Garreaud et al., 2003), whereas during the winter, southerly (westerly) airows bring frontal precipitation from extratropical cyclones (Vuille, 1999). Summer precipitation is the main source of moisture; this is usually in the form of rain at elevations below 4500 m a.s.l. and snow above 5300 m a.s.l. Winter precipitation has been signicantly underestimated, mainly south of 20 S (Vuille, 1996), because it frequently occurs as snow, which is poorly quantied and can fall as low as 1000 m a.s.l. The mean annual boundary between the two moisture
Copyright 2003 Royal Meteorological Society Int. J. Climatol. 23: 14531464 (2003)

CENTRAL ANDEAN WEST-SLOPE RAINSHADOW


3

1455

Altiplano

18S

Ce

-1

Iquique
20S 22S
3 2

ntr al V alle y

J F M A M J J A S O N D

6000

Antofagasta
24oS 26S 70W 28S 68W

4000 2000 masl

-1

J F MA M J J A S O N D

Figure 2. The core area of the Atacama Desert showing topography and the division between stations dominated by summer or winter precipitation. Insets show mean standardized monthly precipitation for each group of stations

systems may be dened as that between the areas that have monthly precipitation maxima in summer or winter (Figure 2) and is diagonal with respect to the Andes due to the different source areas. The third source of moisture is fog and drizzle below the inversion, which is more common in winter and associated with the passage of frontal systems. Data presented by Schemenauer and Cereceda (1991; Cereceda and Schemenauer, 1991) suggest that precipitation rates as high as 100 mm year1 may occur at the inversion layer (800 m a.s.l.) between 26 and 28 S, although this decreases northward and is probably overestimated due to recycling by evapotranspiration processes. The zonal impact of the subtropical high-pressure belt produced by the Hadley circulation generates a strong precipitation decit as a result of descending stable air, centred between 24 and 28 S (Figure 3), although the lowest precipitation actually occurs between 18 and 20 S. Several features should be noted. Firstly, station elevations reduce considerably towards the south, although this is not true of the Andean mountain range, thus causing a bias in any possible precipitationelevation relationship. Secondly, envelope curves indicate a wide scatter of precipitation values in the north but much less variation in the south. These characteristics are related to the two principle sources of precipitation, the northern monsoon and southern frontal systems. The zone of minimum precipitation between 18 and 28 S corresponds with the boundary between the two systems where their effects are both minimized. The continentality and orographic effects can be seen in a southwestnortheast section from Iquique to Belem (Figure 4). Between longitudes 54 and 67 W, there is a gradual reduction in precipitation associated with the increasing effects of distance from the Atlantic Ocean moisture source. Superimposed on this gradual reduction is a signicant increase due to the orographic effect on the eastern (windward) slope of the Andes, starting at 300 m a.s.l., reaching a maximum at 1000 m a.s.l. and declining to background levels by 5000 m a.s.l. By contrast, on the Andean western (lee) slope, the marked reduction in precipitation is a direct result of the rainshadow effect. 3. PRECIPITATIONELEVATION RELATIONS IN THE CORE REGION OF THE ATACAMA DESERT Precipitation data for northern Chile are available from the Direcci on Meteorol ogica de Chile and the Direcci on General de Aguas. These databases contain 50 stations, maintained since the mid 20th century, spread over
Copyright 2003 Royal Meteorological Society Int. J. Climatol. 23: 14531464 (2003)

1456
1500

J. HOUSTON AND A. J. HARTLEY


8000

1200 6000 Precipitation (mm/a) 900 4000 600 2000 300 Elevation (masl) Precipitation (mm)

0 15 20 25 Latitude 30 35

Figure 3. Longitudinal section of precipitation from 15 to 35 S based on 195084 data from Servicio Nacional de Meteorologla e Hidrologia de Peru and 195180 data from Direcci on General de Aguas de Chile. (Black diamonds: precipitation; grey squares: station elevations; grey line: generalized crest line of Andes)

10000

10000

orographic effect rainshadow effect


8000 Elevation (masl) 1000

continentality effect
6000 100

4000

10

topographic profile
2000 1

70

65

60 Longitude

55

0.1 50

Figure 4. Generalized latitudinal section of precipitation and elevation from 70 to 50 W based on Philips World Atlas (Fullard and Darby, 2001), Lizarazu et al. (1997) and Direcci on General de Aguas de Chile. (Note that the data points are not station data but arbitrary points from contoured precipitation maps)

the western slopes of the central Andes between 18 and 24 S. The period 197891 has good quality data from a wide geographical spread of stations, with means close to the long-term average, and including both wet (198486) and dry years (197879 and 198890). Precipitation records are largely of rainfall; but as noted above, both snowfall at high elevations and fog at low elevations are under-recorded. Raw data from the databases were validated and adjusted as described below, and it is considered that the processed data represent a reliable record of the rainfall variability within the region. Of the original data, over 20 stations were rejected for analysis because they did not contain an uninterrupted time series within the periods of analysis or they suffered from data inadequacies. Monthly data were checked for missing or incomplete values. Any station with more than 2% of missing monthly data was excluded from
Copyright 2003 Royal Meteorological Society Int. J. Climatol. 23: 14531464 (2003)

CENTRAL ANDEAN WEST-SLOPE RAINSHADOW

1457

the analysis. The remaining missing data were estimated by inserting the mean monthly value, factored by the annual rainfall for the year compared with the long-term station mean. Monthly values were converted to water years (OctoberSeptember) and the data for each standardized over the period of record. Standardization (z score transformation with units of standard deviation) is essential for time series that have means varying over more than two orders of magnitude; otherwise, comparative results are biased towards those time series with high means. After standardization, each station was regressed against the mean of all stations. Five stations had a correlation coefcient that was not signicant at 10% and three stations had a slope that was signicantly (at 10%) different from unity. Since these stations appeared to be randomly distributed across the area it was considered likely that the data were in error and the stations, amounting to less than 10% of the original database, were not considered further. Specic annual outliers were investigated for anomalous monthly values and corrected in the same way as missing data. Correction of monthly data amounted to less than 1% of all months for all stations used in the analysis. The resulting MAR (OctoberSeptember) for 28 stations between sea level and 4320 m a.s.l. are detailed in Table I and shown in Figure 5(b). The rainshadow effect is pronounced, with precipitation decreasing rapidly with decreasing elevations. However, the relationship is not simple and three zones may be distinguished.
Table I. Station data for the core area of the Atacama Desert between 1977 and 1991 (the precipitationelevation correlation given in the text is based on asterisked stations only those above 2300 m a.s.l. on the Andean west-slope) Station Long. S (decimal deg) 23.45 18.50 21.68 22.28 22.47 19.32 22.32 22.33 20.95 22.45 20.02 22.37 20.93 22.03 20.22 22.20 21.93 21.22 20.02 23.68 26.40 19.13 21.63 20.18 23.58 19.28 22.25 20.97 Lat. W (decimal deg) 70.45 70.27 68.28 68.32 68.92 69.42 68.22 68.65 68.90 69.65 68.82 68.03 69.05 68.07 70.13 67.98 68.33 68.25 69.20 68.07 69.47 69.12 69.52 69.33 67.88 68.65 68.17 68.65 Elevation (m a.s.l.) 10 5 3956 3031 2260 2380 3260 2524 3490 1290 3990 4320 2460 4100 10 4096 3700 3650 2570 2480 2850 4200 802 1815 3350 3965 3350 4200 Annual precipitationa (mm) 4.1 0.9 78.8 36.8 3.2 13.4 65.6 4.2 44.4 0.4 124.9 164.1 11.5 103.2 1.2 173.0 65.1 85.9 26.8 23.1 22.4 140.3 0.15 0.9 52.6 111.6 109.0 127.6 Summerb precipitation (%) 13 50 88 88 42 94 89 63 83 0 92 87 84 94 1 91 89 93 90 69 19 97 0 84 73 93 91 91

Antofagasta Arica Ascotan Ayquina Calama Camigna Caspana Chiu Chiu Copaquire Coya Sur Coyacagua El Tatio Guatacondo Inacaliri Iquique Linzor Ojos de SP Ollague Parca Peine Potrerillos Pumire Quillagua Sagasca Socaire T Isluga Toconce Ujina
a 15

year mean.

b NovemberMarch.

Copyright 2003 Royal Meteorological Society

Int. J. Climatol. 23: 14531464 (2003)

1458
10000

J. HOUSTON AND A. J. HARTLEY

Precipitation (mm/a)

1000

Quelccaya

100

10

(a) 15 - 17 S
1 0 1000 2000 3000 4000 5000 6000

Elevation (masl)
1000
COASTAL ZONE CENTRAL VALLEY

Caquena ElLaco

Sajama

Precipitation (mm/a)

100

10
Antofagasta Iquique Arica Quillagua Calama Chiu chiu Sagasca Coya Sur

(b) 17 - 27 S
CORDILLERA ZONE

0.1 0 1000 2000 3000 4000 5000 6000

Elevation (masl)
10000

Precipitation (mm/a)

1000

100

10

(c) 27 - 35 S
1 0 1000 2000 3000 4000 5000 6000

Elevation (masl)

Figure 5. Precipitationelevation relationships for the Atacama Desert: (a) Peruvian west slope between 15 and 17 S; (b) north Chile west slope between 17 and 27 S; (c) north Chile west slope between 27 and 35 S. Single point data for Quelccaya (Thompson et al., 1985), Sajama (Thompson et al., 1999), El Laco (Vuille, 1996) and Caquena (Fuenzalida and Rutllant, 1986) added

In the Cordillera Zone, from 4300 m a.s.l. down to 2300 m a.s.l., there is a well-developed decline in MAR, which is relatively little affected by location and aspect, and which is related to elevation (A, metres a.s.l.) according to the following best t relationship: MAR = e0.0012A (r = 0.94, signicant at 99.9%)

For the Cordillera Zone, all stations are dominated by summer rainfall and the exponential decline can be assigned to the barrier effects of the Andes preventing the passage of moisture from Amazonia. Separating out the seasonal winter component shows a similar relationship, but there is wider scatter due to the frontal or anticyclonic nature of the moisture source. Precipitation data in Fuenzalida and Rutllant (1986) for northern
Copyright 2003 Royal Meteorological Society Int. J. Climatol. 23: 14531464 (2003)

CENTRAL ANDEAN WEST-SLOPE RAINSHADOW


10000 10000

1459

8000 Elevation (masl)

precipitation profile from Fig 3 volcanic summit


Precipitation (mm) 1000

6000

4000

mean elevation profile precipitation-elevation relationship based on MAR = e 0.0012A


71 70 69 68 67 66 65 64 63 62

100

10

2000

0 Longitude

Figure 6. Precipitation prole across the Andes between 18 and 22 S, indicating that the exponential regression is a good t to the west slope rainshadow data as far east as the peaks of the Western Cordillera. Increased precipitation over the Altiplano is due to convective activity

Chile during 1984 (a relatively wet year) have the same exponential relationship, with a multiplier of 1.14 (i.e. 14% more precipitation due to the wet year). The average elevation of the Altiplano in this region is around 4000 m a.s.l., with volcanic peaks rising to 6000 m a.s.l.; it would be useful to know, since there are very little data at these elevations, whether the correlation breaks down above the level of the Altiplano or extends to the volcanic summits. A study of winter snowfall at El Laco (24 S, 44005000 m a.s.l.) between 1990 and 1993 suggests that annual precipitation may be underestimated by as much as 25%, with MAR equivalent to 215230 mm year1 (Vuille, 1996). A second study by Thompson et al. (1999) indicates that averaged data from Cerro Sajama, Bolivia, in the Western Cordillera at 18 S, suggests an MAR equivalent to 378 mm year1 at 5380 m a.s.l. These data conrm the precipitationelevation correlation described above (and see Figure 5(b)) to at least 4500 m a.s.l. and possibly as high as 5500 m a.s.l. This correlation, however, is only applicable to the west-slope rainshadow zone and breaks down over the Altiplano to the east (Figure 6), where higher levels of precipitation are associated with convective activity derived from Amazonia. Below 2300 m a.s.l. is a zone of extreme hyper-aridity associated with the Central Valley, which also appears to be directly related to elevation according to a different relationship: MAR = 0.03 e0.002A (r = 0.99, signicant at 99.9%)

It is not clear why this occurs, but 60% of the stations in this extreme hyper-arid zone are dominated by winter rainfall, and this may help to explain the different relationship with elevation, since they are on the extreme northerly limit of frontal precipitation sourced from the south. The Coastal Zone receives increased precipitation compared with the Central Valley and this increased precipitation apparently coincidentally conforms in the north (Arica, Iquique) with the Cordillera Zone rainshadow correlation for the 1727 S zone. Coastal zone precipitation also increases in a southerly direction as the stations become increasingly subject to southerly derived winter frontal systems.

4. PRECIPITATIONELEVATION RELATIONS OUTSIDE THE CORE AREA North of the core area described above, data from the Servicio Nacional de Meteorolog a e Hidr ologia de Peru for the southern Peruvian Andean west slope between 15 and 18 S, although not subjected to the same
Copyright 2003 Royal Meteorological Society Int. J. Climatol. 23: 14531464 (2003)

1460

J. HOUSTON AND A. J. HARTLEY

rigorous validation process, demonstrate a precipitationelevation correlation of the same form (Figure 5(a)), conrming the northward extension of the rainshadow effect: MAR = 4.7 e0.0012A (r = 0.95, signicant at 99.9%)

The multiplier suggests a progressive increase in precipitation towards the north, due to the zonal effect and closer proximity to the AtlanticAmazonian source area. The same data hint at the possibility of a northward extension of the extremely hyper-arid zone between 10002000 m a.s.l., and intriguingly that the Coastal Zone, which in this area also suffers from coastal fogs due to the strong inversion, apparently conforms to the Cordillera Zone rainshadow correlation. South of the core area of the Atacama Desert, precipitation data are biased by the preponderance of stations below 3000 m a.s.l. (Figure 3). Nevertheless, it is apparent that precipitationelevation effects break down (Figure 5(c)) and no statistically signicant relationship is present. This conrms the results of Basist et al. (1994), who found that orientation to prevailing wind and exposure were more important determining factors for precipitation in central Chile than elevation.

5. PALAEOGEOGRAPHIC ASPECTS AND THE ONSET OF HYPER-ARIDITY Of the four factors contributing to aridity, namely zonal, continental, rainshadow, and oceanic, it is important for palaeoclimatic reconstructions to evaluate those that have contributed to hyper-aridity. 5.1. Zonal forcing Palaeomagnetic data show that South America has remained within 9 of its current latitude on the planet since at least 165 Ma in the Late Jurassic (Beck, 1999) and has thus been largely within the subtropical highpressure belt of the Hadley circulation since that time, despite uctuations in its extent and strength (Williams, 1994). Stratigraphical and sedimentological evidence conrm that aridity existed in the Atacama Desert since at least 90 Ma in the Upper Cretaceous (Hartley et al., 1992). Nevertheless, the same sedimentological evidence indicates an arid to semi-arid environment with signicant uvial activity and that hyper-aridity did not develop until much later. 5.2. Continental forcing Prior to the Early Cretaceous (140130 Ma), the Gondwana landmass created a supercontinent with a fetch of 5000 km from a Tethys Ocean moisture source, creating a potentially huge continentality effect. With the onset of deep circulation through the Atlantic in the Late Cretaceous (90 Ma) the distance to an oceanic moisture source was greatly reduced and can thus be discounted as a cause for hyper-aridity. 5.3. Elevation forcing Regional uplifts, such as the Andes, have been shown unequivocally to cause increasing aridity (Manabe and Broccoli, 1990; Ruddiman et al., 1997). At elevations of 1000 m the effects of topographic forcing begin to be felt (Browning, 1980), with increasing effect by the time elevation has reached 2000 m (Hay and Wold, 1998; Otto-Bliesner, 1998), and palaeoclimate modelling of the Himalayas suggests that the impacts on climate may develop progressively and in step with increasing uplift Zhiseng et al. (2001). Based on estimations of palaeoelevation for the Andes (Kennan, 1999; Gregory-Wodzicki, 2000), 1000 m a.s.l. was reached by perhaps 20 Ma and 2000 m a.s.l. by 10 Ma. Thus, a rainshadow effect is likely to have been present by at least 10 Ma and probably earlier. Supporting this, Alpers and Brimhall (1988) suggest a progressive desiccation in northern Chile from 15 to 9 Ma based on the cessation of supergene enrichment in porphyry copper deposits, which require precipitation in excess of 100 mm year1 in order to form (Clark et al., 1990). Assuming zonal evaporation rates of between 2000 and 3000 mm year1 and precipitation of
Copyright 2003 Royal Meteorological Society Int. J. Climatol. 23: 14531464 (2003)

CENTRAL ANDEAN WEST-SLOPE RAINSHADOW

1461

100 mm year1 at cessation of supergene enrichment, the P /PET ratio would be between 0.05 and 0.03, indicative of hyper-arid conditions. Additionally, Sillitoe et al. (1968) and Mortimer and Saric (1972) suggest that the cessation of pediplanation at 1211 Ma indicates a change to hyper-arid conditions. Thus, the uplift of the Andes between 15 and 10 Ma and consequent creation of a proto-monsoonal circulation with a pronounced rainshadow on the central Andean west slope is considered to have contributed signicantly to the onset of increasing aridity. 5.4. Oceanic forcing The early Circum-Equatorial Current (CEC), 18060 Ma, caused by the prevailing continental conguration, generated a large-scale gyre in the South Pacic and created a north owing PC; but it is unlikely that this was very cold, because global temperatures were high and no Antarctic Circumpolar Current (ACC) existed (Gerhard and Harrison, 2001). Hence, this is unlikely to have contributed to zonal aridity at this time. The CEC was greatly reduced by 6040 Ma with the northward movement of India and initial closure of Tethys forcing the CEC southward around South Africa. Antarctic Bottom Water (AABW) started to form around 30 Ma (Sykes et al., 1998); this was followed by formation of the ACC around 23 Ma (Kennett, 1982) and major increases in the East Antarctic Ice Sheet during the Middle Miocene between 15 and 12 Ma (Shackleton and Kennett, 1974; Flower and Kennett, 1994), causing global climatic deterioration and aridication and the development of the deep cold-water PC with enhanced upwelling 1411 Ma (Crowley and North, 1991; Dunbar et al., 1990, Tsuchi, 1997). Thus, global climate deterioration and intensication of cold-current upwelling at 1511 Ma is also considered to have contributed to the early development of increasing aridity. Similar conclusions have been drawn for the Namib Desert in respect of the Benguela Current (Siesser, 1980; Van Zinderen Bakker and Mercer, 1986). 5.5. Discussion Arguments for a later onset of hyper-aridity based on lacustrine and uvial sedimentation in the Atacama Desert during 83 Ma have been attributed to global climate change rather than uctuations in the strength of the PC or Andean uplift (Hartley and Chong, 2002). Supporting this conclusion, nal closure of the Panama seaway at 4.6 Ma (Haug and Tidemann, 1998) and closure of the Indonesian seaway at 53 Ma (Cane and Molnar, 2002) have both been cited as causes for changes in the global thermohaline circulation and consequent Northern Hemisphere glaciation and intensied global climate aridication. On the other hand, late-stage lacustrine and uvial sedimentation may not be inconsistent with an early onset to hyperaridity, since they may represent periods of greater humidity originating from increased climate variability. For example, Hayward et al. (2002) and Billups (2002) demonstrate Late Miocene to Pliocene (73 Ma) climate variability also associated with orbital forcing. Widespread canyon incision and associated erosion in northern Chile between 3 and 1 Ma was undoubtedly initiated by tectonic uplift (Sebrier et al., 1988), but is probably also associated with increased humidity in the Altiplano and consequent runoff in the Atacama. Such increased humidity post 3 Ma is inconsistent with a single late-stage onset of hyper-aridity at 3 Ma, but not with a uctuating climate scenario. Similar canyon incision has also been demonstrated in the Namib Desert between 2.8 and 1.30.4 Ma and has been attributed to a period of increased humidity at a time of Antarctic Ice Sheet expansion (Cieselski and Grinstead, 1986; Van de Wateren and Dunai, 2001). Additional evidence for climate uctuations is provided by Thompson et al. (1999) and Amman et al. (2001), who demonstrate that stadial (glacial) conditions during the Pleistocene and Holocene led to increased precipitation and reduced evaporation over the Altiplano, although this is not universally accepted (e.g. Dettinger et al., 2000). Conversely, recent studies by Stott et al. (2002) and Koutavas et al. (2002), based on oxygen isotope ( 18 O) composition and the magnesium/calcium ratios of planktonic foraminifera in the Pacic, have correlated stadials with enhanced El Ni no-like circulations (which lead to droughts in the Altiplano and Atacama) and interstadials with enhanced La Ni na-like conditions (which are associated with increased humidity over the Altiplano and the Atacama).
Copyright 2003 Royal Meteorological Society Int. J. Climatol. 23: 14531464 (2003)

1462

J. HOUSTON AND A. J. HARTLEY

Despite the lack of congruence in these latter studies, signicant uctuations in climate have been demonstrated almost universally from the Oligocene to the present, with the PliocenePleistocene registering increased frequencies of change.

6. CONCLUSIONS The Atacama Desert experiences a hyper-arid climate in the core zone between 15 and 30 S at elevations between sea level and 3500 m. Within this zone there are three moisture sources: the northeast summer monsoon, southwest frontal precipitation and fog below the 800 m a.s.l. inversion. Of these, the most effective is the northeast summer monsoon, which creates a rainshadow effect on the west (lee) slope of the Andes and is well developed north of 27 S. Decreasing precipitation with decreasing elevation has an exponential relationship in a form that is maintained from 27 S at least as far north as 15 S. Nevertheless, although the exponent remains stable, the multiplier varies latitudinally and temporally, increasing in a northward direction and during wet years. The relationship appears to be reliable to at least 4500 m a.s.l. and may extend to 5500 m a.s.l. Data at such high elevations are extremely difcult to obtain, but such data as exists suggest that increased precipitation goes unrecorded, and may have important water resource implications. South of 27 S, where winter frontal precipitation dominates and hyper-aridity gives way progressively to arid and semi-arid zones, there is apparently no strong relationship between precipitation and elevation. Four principal factors control aridity within the Atacama Desert: its zonal position in relation to the Hadley circulation; its distance from the main AtlanticAmazonian source of moisture; the elevation of the Altiplano, creating a rainshadow effect and altering the radiation balance; and its proximity to the cold PC. Largely due to the fact that South America has maintained its latitudinal position for the last 165 Ma, the Atacama Desert may be amongst the oldest on Earth due to its zonal position relative to the Hadley circulation. The continentality effect suffered a major reduction with the opening of the Atlantic (90 Ma), and thus cannot be considered as a potential cause of hyper-aridity. Evidence for the development of a rainshadow effect between 15 and 10 Ma suggests that this may have contributed signicantly to the development of hyper-aridity. Major cooling (aridication) associated with changes in global oceanic circulation and climate at 1512 Ma may also have contributed to the development of hyper-aridity. Such an early onset to hyperaridity need not be inconsistent with later climate uctuations causing alternations between relatively humid (semi-arid to arid) and hyper-arid periods which would give rise to the observed sedimentation and erosion record. Thus, it is considered that the evidence is weighted in favour of an early onset of hyper-aridity. This is most likely to have occurred progressively with the uplift of the Andes and the development of the west-coast rainshadow coupled with Middle Miocene global climate deterioration and the onset of the upwelling cold PC, but later climate uctuations permitted periodic amelioration of the intense aridity.
ACKNOWLEDGEMENTS

Funding for this study was provided by Nazca S.A.


REFERENCES Alpers CN, Brimhall GH. 1988. Middle Miocene climatic change in the Atacama Desert, northern Chile: evidence from supergene mineralisation at La Escondida. Geological Society of America Bulletin 100: 16401656. Amman C, Jenny B, Kammer K, Messerli B. 2001. Late Quaternary glacier response to humidity changes in the arid Andes of Chile (1829 S). Palaeogeography, Palaeoclimatology, Palaeoecology 172: 313326. Fullard H, Darby HC (eds). 2001. World Atlas. George Philip and Co.: London. Basist A, Bell GD, Meentemeyer V. 1994. Statistical relationships between topography and precipitation patterns. Journal of Climate 7: 13051315. Beck ME. 1999. Jurassic and Cretaceous apparent polar wander relative to South America: some tectonic implications. Journal of Geophysical Research 104: 50635068. Copyright 2003 Royal Meteorological Society Int. J. Climatol. 23: 14531464 (2003)

CENTRAL ANDEAN WEST-SLOPE RAINSHADOW

1463

Betancourt J, Quade J, Seltzer G (eds). 2001. Paleoclimatology of the Central Andes. PEPI Workshop, Tucson, Arizona. Available at http://wwwpaztcn.wr.usgs.gov/pcaw/ [Last accessed March 2002]. Billups K. 2002. Late Miocene through early Pliocene deep water circulation and climate change viewed from the sub-Antarctic South Atlantic. Palaeogeography, Palaeoclimatology, Palaeoecology 185: 287307. Browning KA. 1980. Structure, mechanism, and prediction of orographically enhanced rain in Britain. In Orographic Effects in Planetary Focus, Hide R, White PW (eds). Global Atmospheric Research Programme Publications Series no. 23. WMO: Geneva; 85114. Cane MA, Molnar P. 2002. Closing of the Indonesian seaway as a precursor to East African aridication around 34 million years ago. Nature 411: 157162. Cereceda P, Schemenauer RS. 1991. The occurrence of fog in Chile. Journal of Applied Meteorology 30: 10971105. Cieselski PF, Grinstead GP. 1986. Pliocene variations in the position of the Antarctic Convergence in the southwest Atlantic. Palaeoceanography 1: 197232. Clark AH, Tosdal RM, Farrar E, Plazolles VA. 1990. Geomorphic environment and age of supergene enrichment of the Cuajone, Quellaveco and Toquepala porphyry copper deposits, southern Peru. Economic Geology 85: 16041628. Crowley TJ, North GR. 1991. Palaeoclimatology. Oxford Monographs on Geology and Geophysics 18. Oxford University Press: Oxford. Dettinger MD, Cayan DR, McCabe GS, Marengo J. 2000. Multiscale streamow variability associated with El Ni nosouthern oscillation. In El Ni no and the Southern Oscillation; Multiscale Variability and Global and Regional Impacts, Diaz HF, Markgraf V (eds). Cambridge University Press: Cambridge; 113148. Dunbar RB, Marty RC, Baker PC. 1990. Cenozoic marine sedimentation in the Sechura and Pisco basins, Peru. Palaeogeography, Palaeoclimatology, Palaeoecology 77: 235261. Figueroa SN, Satyamurty P, da Silva PL. 1995. Simulations of the summer circulation over the South American region with an eta coordinate model. Journal of Atmospheric Sciences 52: 15731584. Flower BP, Kennett JP. 1993. Middle Miocene ocean climate transition: high resolution oxygen and carbon isotopic records from Deep Sea Drilling Project Site 588A, SW Pacic. Palaeoceanography 8: 811843. Fuenzalida H, Rutllant J. 1986. Estudio sobre el origen del vapor de agua que precipita en el invierno altiplanico. Convenio de Cooperacion Direcci on General de Aguas y Universidad de Chile. Unpublished Informe Final. Garreaud RD. 1999. Multiscale analysis of the summertime precipitation over the central Andes. Monthly Weather Review 127: 900921. Garreaud RD, Vuille M, Clement AC. 2003. The climate of the Altiplano: observed current conditions and mechanisms of past changes. Palaeogeography, Palaeoclimatology, Palaeoecology 194: 522. Gerhard LC, Harrison WE. 2001. Distribution of oceans and continents: a geological constraint on global climate variability. In Geological Perspectives on Global Climate Change, Gerhard LC, Harrison WE, Hanson BM (eds). AAPG Studies in Geology 47. American Association of Petroleum Geologists: Tulsa OK. Gregory-Wodzicki KM. 2000. Uplift history of the central and northern Andes: a review. Geological Society of America Bulletin 112: 10911105. Hartley AJ, Chong. G. 2002. Late Pliocene age for the Atacama Desert: implications for the desertication of western South America. Geology 30: 4346. Hartley AJ, Flint S, Turner P, Jolley EJ. 1992. Tectonic controls on the development of a semi-arid basin as reected in the stratigraphy of the Purilactis Group (Upper CretaceousEocene), northern Chile. Journal of South American Earth Sciences 5: 275296. Haug GH, Tidemann R. 1998. Effect of the formation of the Isthmus of Panama on Atlantic Ocean thermohaline circulation. Nature 393: 673676. Hay WW, Wold CM. 1998. The role of mountains and plateaus in a Triassic climate model. In Tectonic Boundary Conditions for Climate Reconstructions, Crowley TJ, Burke KC (eds). Oxford University Press: Oxford. Hay WW, Barron EJ, Thompson SL. 1990. Global atmospheric circulation experiments on an Earth with meridional pole-to-pole continent. Journal of the Geological Society of London 147: 385392. Hayward AM, Valdes PJ, Sellwood BW. 2002. Magnitude of climate variability during middle Pliocene warmth: a palaeoclimate modeling study. Palaeogeography, Palaeoclimatology, Palaeoecology 188: 124. Kennan L. 1999. Large-scale geomorphology of the Andes: interrelationships of tectonics, magmatism and climate. In Geomorphology and Global Tectonics, Summereld MA (ed.). Wiley: Chichester; 167199. Kennett J. 1982. Marine Geology. Prentice Hall: New Jersey. Koutavas A, Lynch-Stieglitz J, Marchitto TM, Sachs JP. 2002. El Ni no-like pattern in ice age tropical Pacic sea surface temperature. Science 297: 226230. Lizarazu J, Soria A, Cortes L. 1997. Mapa Hidroge ologico de Bolivia. Servicio Nacional de Geolog a y Miner a: La Paz. Manabe S, Broccoli AJ. 1990. Mountains and arid climate of middle latitudes. Science 247: 192195. Markgraf V (ed.). 2001. Interhemispheric Climate Linkages. Academic Press: New York. Mortimer C, Saric N. 1972. The Cenozoic evolution of northern Chile. Unpublished Report Instituto de Investigaciones Geol ogicas, Santiago, Chile. Otto-Bliesner BL. 1998. Effects of tropical mountain elevations on the climate of the Late Carboniferous: climate model simulations. In Tectonic Boundary Conditions for Climate Reconstructions, Crowley TJ, Burke KC (eds). Oxford University Press: Oxford. Ruddiman WF, Raymo ME, Prell WL, Kutzbach JE. 1997. The upliftclimate connection: a synthesis. In Tectonic Uplift and Climate Change, Ruddiman WF (ed.). Plenum: New York; 471516. Schemenauer RS, Cereceda P. 1991. The quality of fog water collected for domestic and agricultural use in Chile. Journal of Applied Meteorology 31: 275290. Sebrier M, Lavenu A, Fornari M, Soulas JP. 1988. Tectonics and uplift in central Andes (Peru, Bolivia and northern Chile) from Eocene to present. G eodynamique 3: 85106. Shackleton NJ, Kennet JP. 1974. Palaeotemperature history of the Cenozoic and the initiation of the Antarctic glaciation: oxygen and carbon isotope analysis in the DSDP Sites 277, 279 and 281. In Initial Reports of the Deep Sea Drilling Program, 29, Kennet JP (ed.). US Government Printing Ofce: Washington, DC; 743755. Siesser WG. 1980. Late Miocene origin of the Benguela upwelling system off northern Namibia. Science 208: 283285. Copyright 2003 Royal Meteorological Society Int. J. Climatol. 23: 14531464 (2003)

1464

J. HOUSTON AND A. J. HARTLEY

Sillitoe RH, Mortimer C, Clark AH. 1968. A chronology of landform evolution and supergene mineral alteration, southern Atacama Desert, Chile. Transactions of the Institute of Mining and Mineralogy 77: 166169. Stott L, Poulsen C, Lund S, Thunell R. 2002. Super ENSO and global climate oscillation at millennial time scales. Science 297: 222226. Sykes TJS, Ramsay ATS, Kidd RB. 1998. Southern Hemisphere Miocene bottom-water circulation: a palaeobathymetric analysis. In Geological Evolution of Ocean Basins: Results from the Ocean Drilling Program, Cramp A, MacLeod CJ, Lee SV, Jones EJW (eds). Geological Society, London, Special Publication, 131. Geological Society Press: London; 4354. Thompson LG, Mosley-Thompson E, Bolzan JF. 1985. A 1500 year record of tropical precipitation in ice cores from the Quelccaya Ice Cap, Peru. Science 229: 971973. Thompson LG, Davis ME, Mosley-Thompson E, Sowers TA, Henderson KA, Zagorodnov VS, Lin PN, Mikhalenko VN, Campen RK, Bolzan JF, Cole-Dai J, Francou B. 1999. A 25 000 year tropical climate history from Bolivian ice cores. Science 282: 18581864. Tsuchi R. 1997. Marine climatic responses to Neogene tectonics of the Pacic Ocean seaways. Tectonophysics 281: 113124. UNEP. 2001. Dry and sub-humid lands biodiversity denitions. Available at http://www.biodiv.org/programmes/areas/ dryland/denitions.asp [Last accessed December 2002]. Van de Wateren FM, Dunai TJ. 2001. Late Neogene passive margin denudation history cosmogenic isotope measurements from the central Namib desert. Global and Planetary Change 30: 271307. Van Zinderen Bakker EM, Mercer JH. 1986. Major Late Cenozoic climatic events and palaeoenvironmental changes in Africa viewed in a world wide context. Palaeogeography, Palaeoclimatology, Palaeoecology 56: 217235. Vuille M. 1996. Zur raumzeitlichen Dynamik von Schneefall und Ausaperung im Bereich des s udlichen Altiplano, S udamerika. Geographica Bernensia 45: 1118. Vuille M. 1999. Atmospheric circulation over the Bolivian Altiplano during dry and wet periods and extreme phases of the southern oscillation. International Journal of Climatology 19: 15791600. Williams MAJ. 1994. Cenozoic climate changes in deserts: a synthesis. In Geomorphology of Desert Environments, Abrahams AD, Parsons AJ (eds). Chapman and Hall: London; 644670. Zhiseng A, Kutzbach JE, Prell WL, Porter SC. 2001. Evolution of Asian monsoons and phased uplift of the HimalayaTibetan plateau since Late Miocene times. Nature 411: 6266. Zhou JY, Lau KM. 1998. Does a monsoon climate exist over South America? Journal of Climate 11: 10201040.

Copyright 2003 Royal Meteorological Society

Int. J. Climatol. 23: 14531464 (2003)

You might also like