applications of climatic resources in mountainous areas · however, mountain chains and peaks cause...

48
CHAPTER 15 Applications of Climatic Resources in Mountainous Regions By Guillermo A. Baigorria and Consuelo C. Romero This chapter was externally reviewed by Kulasekaran Ramesh, Hans Schreier, Jean-Joinville Vacher and Jin Yun 15.1 Introduction 15.2 Why is the climate different in mountainous regions? 15.3 Meteorological variables in mountainous regions 15.3.1 Incoming solar radiation 15.3.1.1 Facing the sun 15.3.1.2 Atmospheric transmissivity 15.3.1.3 Favorable use of incoming solar radiation knowledge in mountains 15.3.2 Temperature 15.3.2.1 Temperature versus altitude? 15.3.2.2 Frost events 15.3.2.3 Daily temperature swing and thermo regulator agents 15.3.2.4 Favorable use of mountainous temperature knowledge 15.3.3 Rainfall 15.3.3.1 Rainfall processes in mountainous regions 15.3.3.2 Rainfall, altitude and orography 15.3.3.3 Monitoring rainfall in mountainous regions 15.3.3.4 Favorable use of mountainous rainfall knowledge 15.3.4 Atmospheric humidity and Foehn effect 15.4 Applications of the spatio-temporal climate variability knowledge in mountainous regions 15.4.1 Influence of climate on soil formation: A case study 15.4.2 Soil erosion in mountain regions: A case study 15.5 Risk associated to mountain climate processes 15.6 Conclusions References 15.1. Introduction As endorsed in 1992 at the Earth Summit in Rio de Janeiro (in Agenda 21, Chapter 13, ‘Managing Fragile Ecosystems: Sustainable Mountain Development.’), ‘Mountains are an important source of water, energy and biological diversity. Furthermore, they are a source of such key resources as minerals, forest products and agricultural products and of recreation. As a major ecosystem representing the complex and interrelated ecology of our planet, mountain environments are essential to the survival of the global ecosystem.’ 1

Upload: others

Post on 19-Oct-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

CHAPTER 15

Applications of Climatic Resources in Mountainous Regions

By Guillermo A. Baigorria and Consuelo C. Romero

This chapter was externally reviewed by Kulasekaran Ramesh, Hans Schreier, Jean-Joinville Vacher and Jin Yun

15.1 Introduction

15.2 Why is the climate different in mountainous regions?

15.3 Meteorological variables in mountainous regions15.3.1 Incoming solar radiation

15.3.1.1 Facing the sun15.3.1.2 Atmospheric transmissivity15.3.1.3 Favorable use of incoming solar radiation knowledge in

mountains15.3.2 Temperature

15.3.2.1 Temperature versus altitude?15.3.2.2 Frost events15.3.2.3 Daily temperature swing and thermo regulator agents15.3.2.4 Favorable use of mountainous temperature knowledge

15.3.3 Rainfall15.3.3.1 Rainfall processes in mountainous regions15.3.3.2 Rainfall, altitude and orography15.3.3.3 Monitoring rainfall in mountainous regions15.3.3.4 Favorable use of mountainous rainfall knowledge

15.3.4 Atmospheric humidity and Foehn effect

15.4 Applications of the spatio-temporal climate variability knowledge inmountainous regions

15.4.1 Influence of climate on soil formation: A case study15.4.2 Soil erosion in mountain regions: A case study

15.5 Risk associated to mountain climate processes

15.6 Conclusions

References

15.1. Introduction

As endorsed in 1992 at the Earth Summit in Rio de Janeiro (in Agenda 21, Chapter 13, ‘Managing Fragile

Ecosystems: Sustainable Mountain Development.’), ‘Mountains are an important source of water, energy and

biological diversity. Furthermore, they are a source of such key resources as minerals, forest products and

agricultural products and of recreation. As a major ecosystem representing the complex and interrelated ecology of

our planet, mountain environments are essential to the survival of the global ecosystem.’

1

Page 2: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Usually considered as ‘Bad Lands’ in most developed countries, mountainous regions are the homeland for

around 270 million rural mountain people in developing and transition countries. According to Huddleston et al.

(2003), 78 percent of the world's mountain area is considered marginally- or unsuited for arable agriculture and only

7 percent is currently classified as cropland. In contrast, rainfed agriculture constitutes around 83% of global

agriculture (Borlaug and Dowswell, 2000), but this percentage rises to almost 100% when considering resource-poor

farmers in mountainous regions.

Mountains are responsible for delivering water and sediment distribution to lowlands. Fertile and deep soil

profiles in lowlands valleys are usually formed at the expense of soils eroded from hillsides, leaving behind erosional

features and shallow soil (Romero et al., 2007b). An estimated 80% of fresh water consumption by people comes

from rainfall and melting glaciers in the mountainous areas (Schreier et al., 2002) and 12% of the global population

is directly supported from resources in the mountains while 50% are indirectly affected by mountains. However many

of the world’s most impoverished and food-insecure people live in these regions (UN, 2003).

Most of the abiotic risk in mountainous regions is associated to the effect of the interannual climatic

variability that severely affects the fragile mountain environment. The high climatic variability affects all human

activities in mountains. The focus of this chapter is to document the climatic impacts on resources in mountains with

an emphasis on small scale issues. In doing so, we can explore ideas on how people can adapt to these climatic

conditions and how they can manage the climatic resources in these regions. First we need to describe the effects of

the mountains on the general atmospheric circulation patterns and vice versa, and discuss the implications these

have on decision-making processes at the regional and farm levels.

15.2. Why is the climate different in mountainous regions?

The main and most evident difference between mountain climate and other regions is altitude (Figure 1).

The atmospheric density over the mountains is reduced and this increases atmospheric transmissivity (section

15.3.1.2; Baigorria et al., 2004; Whiteman, 2000). Surface incoming solar radiation is equivalent to the

extraterrestrial solar radiation (which depends on the relative position of the earth and sun and the latitude)

multiplied by the filtering coefficient of atmospheric transmissivity. In mountainous areas, these filtering coefficient

are commonly greater than 0.75 on an average monthly basis during dry season (clear skies), and daily atmospheric

transmissivity coefficients can reach values up to 0.95 (Baigorria, 2005; Lee, 1978; SENAMHI-MEM, 2003).

2

Page 3: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Figure 1: Mountainous regions in the world (based on data from U.S. Geological Survey’s EROS Data Center, 1996)

3

Page 4: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Mountains are also dominated by complex topography which is a major factor of climate in mountainous

regions, but this is not always true. For example in the Andean High Plateau (14°S Latitude, 71°’ W Longitude to

20°S Latitude, 67° W Longitude; Figure 2) which ranges from 3,600 to 5,000 m above the sea level, the mean slope

angle is 2 degrees. At a first sight, this 135,910 km2 plateau should be a desert in terms of human presence;

however, agriculture and livestock practices are performed in many areas since centuries ago due to certain special

climatic characteristics which will be discussed in section 15.3.2.3.

Figure 2: Location and altitude maps of the Andean High Plateau.

In mountainous areas, topography is a major factor that determines the amount of incoming solar energy

received at the surface. Variability in elevation, slope, aspect (angle formed by the geographic north and the

mountain hillside), and shade can create strong local gradients in the incoming solar radiation that directly and

indirectly affect such biophysical processes such as air and soil heating, energy and water balances, and primary

production Dubayah and Rich, 1995).

Altitude has been used as the main factor affecting the spatial distribution of rainfall and temperature.

However, most of the time the combination of topography and wind circulation are the main factors responsible for

rainfall distribution, whereas temperature variation are mainly explained by the same variables plus altitude and the

low water vapor content in the atmosphere.

Global Circulation Models (GCMs) are usually unaffected by the high elevation gradient (Figueroa et al.,

1995). A Cumulus nimbus cloud could reach 15 km height (Whiteman, 2000) while the Everest peak reaches 8,846

m altitude (Fellman et al., 1992). However, mountain chains and peaks cause the air to lift and then condensation

processes in the lower troposphere occurs, which, under certain atmospheric humidity conditions, causes small

scale clouds formations and even rainfall. These movements affect not only the condensation level of the mountain,

but due to the atmospheric wave’s formation, affect areas far away parallel to the lee side of mountains.

4

Page 5: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Water vapor constitutes around 4% in volume of the atmosphere (3% in weight) close to land surface, and is

almost absent beyond 10 km. Water vapor is available to the atmosphere by evaporation from water and land

surfaces and from transpiration of plants, which then raises by air turbulence. According to the Clausius-Clapeyron

equation, saturation vapor pressure depends on air temperature, and because air temperature decrease with

altitude, the water vapor content in the atmosphere decrease inversely to altitude (Barry and Chorley, 1998). Water

vapor content is another important characteristic in mountainous areas since it is responsible for the sensible heat

level which is directly related to the daily temperature swing. Water in mountainous areas is not only important for

the crop water requirements, but also as a thermo-regulator agent. As we will see later, presence of water bodies

affects the spatial distribution of temperature especially in mountainous areas.

15.3 Meteorological variables in mountainous regions

15.3.1. Incoming solar radiation

Interactions between incoming solar radiation and hillside orientation as well as the effects of altitude in the

optical thickness of the atmosphere will be described in this section. Effects of global circulation patterns in the

atmospheric transmissivity in mountainous regions, as well as the relationships between the spatial and temporal

variability of incoming solar radiation with crops, and plague and diseases will be emphasized.

15.3.1.1 Facing the sun

An important factor is the orientation of the mountain change. East-oriented hillsides receive direct and

diffuse incoming solar radiation during the morning, whereas west-oriented hillsides receive diffuse and part-time

direct incoming solar radiation later in the morning (Figure 3a). The process reverses during the afternoon. However,

land warming during the morning produces convective (vertical) movements thus forming convective clouds. The

albedo of these clouds reflects part of the direct incoming solar radiation into space, which acts as a filter. As a

consequence, during the afternoon, the amount and quality of incoming solar radiation decreases in comparison to

that received during morning (Figure 3b). As a result, west-oriented hillsides receive less amount of the total daily

incoming solar radiation than those east-oriented hillsides. This fact is emphasized in deep and narrow mountain

passes, gorges and canyons.

Hillsides oriented towards the equator receive more total annual incoming solar radiation than those oriented

to the Polar Regions. The higher the latitude the larger the difference found. This difference increases while crossing

the 23.45° N and 23.45° S latitudes, which are the extreme latitudes where the relative movement of the sun reach

the tropics of Cancer and Capricorn respectively during the solstices. As an historical example of the hillside

orientation effects we can cite the city of Shechem that was located between the vegetated north-facing slope of Mt.

Gerizim and the arid south-facing slope of Mt. Ebal. This ecological difference was due to the almost double

evaporative rate originated by the direct incoming solar radiation over the south-facing slope Mt. Ebal (Hillel, 2006).

5

Page 6: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

This geographical orientation creates climatic differences which are especially important in mountainous

chains, like the Himalayas which follow an east-west direction, whereas the Andes follow a north-south direction.

Due to difference in the radiative budget, latitudinal changes generate more climate variability than longitudinal

changes. This creates a combination of climates and microclimates that differ between mountain chains that dissect

the Earth’s surface. This creates such extremes as desert lands in the southern Andes of Chile and Argentina, and

rain forests in the northern Andes of Colombia and Venezuela. Geographical orientation at local scales also has a

big influence on climate as will be shown in sections 15.3.2, 15.3.3 and 15.3.4.

Figure 3: Effect of the hillside orientation in receiving incoming solar radiation. (a) morning, and (b) afternoon.

15.3.1.2 Atmospheric transmissivity

Most of the energy available in the environments is due to the amount of incoming solar radiation. Gases

and condensating nuclei in the atmosphere filter some of the energy available in the top of the atmosphere before

reach the Earth’s surface. The percentage of energy that goes across this filter on its travel to ground surface is

called atmospheric transmissivity (τ) and the calculation is shown in Equation 1. The τ have temporal and spatial

variations, however because mountains reduce the optical thickness of the atmosphere, under clear sky conditions,

this values can reach values as higher as 95% (Lee, 1978).

6

Page 7: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

100×=oH

Hτ (1)

Where: H and Ho (MJ m-2 d-1) are the measured and the extraterrestrial incoming solar radiation respectively.

Figure 4: Climate maps of atmospheric transmissivity (%) in different seasons in Peru: (a) March, (b) June, (c)

September; and (d) December (adapted from Baigorria et al., 2004)

7

Page 8: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Baigorria et al., (2004) found a gradient of τ parallel to the Andes, as a result of altitude and orographic

barrier effects over clouds and aerosols in the atmosphere. Thus, mountains are major topographic barriers to

weather systems and airflow below 2500 m a.s.l., preventing the regular exchange of air masses, in case of the

Andes, between the Pacific and Atlantic.

However, high altitudes do not mean always high τ as shown in Figure 4. Interactions between climatic

controls and atmospheric circulations patterns create unique seasonality effects like in the case of the Andean High

Plateau (Figure 2). In this area, precipitation is concentrated over a the two- to three-month wet season during the

austral summer, which is associated with the development of convective clouds over the Central Andes and the

southwestern part of the Amazon Basin (Horel et al., 1989). As a consequence, during these wet episodes, around

50% of the area is covered by clouds during the afternoons whereas convective clouds are almost non-existent

during the dry episodes (Garreaud, 1999). These observations explain both the occurrence of monthly average τ

values that are lower than 65% during winter and the occurrence of rainfall during the summer, despite the fact that

in some areas the altitude is more than 4000 m a.s.l. However, monthly average τ values higher than 75% can be

reached during spring, due to the sun being positioned over the Southern Hemisphere and the cloud systems not yet

being formed. As a conclusion we can say that τ do not increases with elevations, but the potential of getting higher

τ values is increased with elevation.

15.3.1.3 Favorable use of incoming solar radiation knowledge in mountains

Due to the potentially high transmissivity levels of the atmosphere at high altitudes (Baigorria et al., 2004;

Lee, 1978), incoming solar radiation at the crop surface, exceeds the maximum threshold of Photosynthetic Active

Radiation (PAR) needed for photosynthesis. This is especially true when receiving direct incoming solar radiation. As

explained in section 15.3.1.1, north- and south-oriented hillsides receive the incoming solar radiation of medium to

high quality during the morning, but medium to low quality during the afternoon which is related to weather

conditions. Changes of ratios of PAR to H with altitude have been reported in the range of +3.6% per km under

cloudless weather and -1.8% per km under cloudy weather in the firsts 1500 m of altitude in the Naeba Mountain of

Japan (Wang et al., 2007). Thus, daily integration of incoming solar radiation for crops is higher in north- and south-

oriented hillsides, followed by east-oriented and finally west-oriented hillsides.

This information is important for crop zoning. If the interest is on fruits where markets favor color

appearance, sugar concentration levels (°Brix), as for instance apples (Merzlyak et al., 2005; Reay and Lancaster

2001), grapes (Kliewer, 1977; Kliewer, et al., 1967), peaches (Erez and Flore, 1986; Layne et al., 2001), etc., then

east-oriented hillsides are recommended. However, extreme conditions of incoming solar radiation cause sunburn or

sunscald (Merzlyak et al., 2002; Piskolczi, et al., 2004). Therefore, crop and cultivar selection, especially in fruit

plantations with several years before the first harvest, is one of the most important factors to cosider before planting

in mountainous regions.

8

Page 9: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Planting at higher densities than normal can be favorable under conditions of high incoming solar radiation

(Dosio et al., 2000). If the interest is focused on yield, C4 crops, (maize, sorghum, etc.) then north- and south-

oriented hillsides are recommended. Finally, because under low water vapor contents (commonly found at high

altitudes) incoming solar radiation is directly related to temperature, crops like potato, native roots, tubers and

pastures, as well as green or yellow fruits cultivars are suitable for these conditions.

This zoning approach is based only on quantity and quality of incoming solar radiation. This must be

combined with other variables, especially rainfall regimes or water availability for a better crop selection.

Another favorable use of potentially high levels of incoming solar radiation in mountainous regions is the

control of plagues and diseases. As in the case of crops, majority of plagues and diseases are affected by

constrains on growth and development caused by climatic conditions. It means that the most important plagues and

diseases in the lowlands, hardly survive under the climatic conditions in high mountainous regions. For example,

incoming solar radiation have a detrimental effect on the germination of Phytophthora infestans sporangia (Jaime-

Garcia et al., 1999; Mizubuti et al., 2000; Rotem and Aust, 1991; Rotem et al., 1985) which is responsible for the late

potato blight, a major constraint to potato production worldwide (Forbes et al., 2001).

A good example is the district of Huasahuasi – Junín in Peru (14°23’ S latitude, 71°19’ W longitude, 4046 m

a.s.l. altitude) which is located in the climatic boundary area where commercial potato varieties can be cultivated but

where almost no plague and disease can survive. There, high quality asexual potato seeds are produced for highly

productive fields in the lowlands, with low needs for pesticides, fungicides and nematicides (Bentley et al., 2001;

Moreno, 1985). However some changes in climate have been affected this region in the last years, so slightly

increasing the incidence of plagues and diseases.

15.3.2. Temperature

15.3.2.1 Temperature vs. altitude?

When unsaturated air (relative humidity less than 100%) is lifted, it cools at the thermodynamic rate of 9.8°C

per 1 km altitude, which is called dry adiabatic lapse rate. However, this happens in a free atmosphere, far away

from the ground surface effect and absence of temperature inversions1. The direct relationship between temperature

and altitude is only shown when the vertical structure of the atmosphere is analyzed but not necessary when

topography and land characteristics are taken into account. For example, Tang and Fang (2006) reported different

temperature lapse rates at different hillslope orientation in Mt. Taibai (Qinling Mountains, China). South-oriented

hillslopes registered a temperature lapse rate of 0.34 ± 0.05 °C/100 m whereas north-oriented hillslopes 0.50± 0.02

°C/100 m. These values showed a large seasonal difference within a maximum range of 59% and 42% for the

southern and northern hillslopes respectively. François et al. (1999) reported higher temperatures in the hillsides of

mountains and volcanoes in comparison to the flat lower areas of the Andean High Plateau. Thus, this theoretical

relationship between altitude and temperature is far from be linear specially when applied on large mountain areas,

1 Layers in which temperature increases with height instead of decreasing (Whiteman, 2000)

9

Page 10: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

due to the effect oceans and continents have on the atmosphere that diminishes with height (Baigorria, 2005;

Peixoto and Oort, 1992).

This dry adiabatic lapse rate is usually misunderstood and applied to the mountains in the wrong way. Many

highland valleys above 2500 meters a.s.l. have higher temperatures than its neighbors in adjacent valleys at the

same altitude. These are usually fertile valleys supporting high density of livestock and intense agriculture. These

important areas are exceptions to the rule of the inverse relationship between altitude and temperature.

At local scales, spatial variations of minimum temperatures are closely related to the terrain type due to cold

air accumulation. Thus, frost occurs most frequently in narrow valleys and concave locations, whereas peaks and

convex areas are found to have very few radiative frost events (Lindkvist et al., 2000). This is why we are going to

describe the spatial and temporal distribution of temperature in mountainous areas in more detail.

15.3.2.2 Frost events

We have already described the importance of the high levels of atmospheric transmissivity in mountainous

regions, and how these are related to high values of incoming solar radiation. All bodies with temperatures above 0°

Kelvin generate radiation in a wavelength inverse to the body’s temperature (Peixoto and Oort, 1992). Because

Earth’s temperature is higher than 0°K (assumed to have a temperature of 255°K) it generates the maximum

emission in the infrared (thermic) range (~10 µm). In mountainous regions, due to the low atmospheric optical

thickness and the lack of water vapor in the atmosphere (which captures the long-wave radiation and maintains the

sensible heat), the downward long-wave radiation is reduced and the Earth’s radiation disperses over space under

cloudless weather nights. This loss of energy diminises surface temperature, which depending on other atmospheric

and land characteristics, can reach temperatures below 0°C. This is called radiative frost and often occurs during

night-time clear skies, including in the most equatorial mountainous regions of the Ecuadorian Andes (Baigorria,

2005; Baigorria et al., 2007; Crissman et al., 1998). The major portion of radiative frosts occurs during the dry

season when rainfed-agricultural fields are usually fallow. However, frost events can occur early and/or later during

the cropping season seriously affecting emerging crops and/or harvests respectively. Knowledge of the free-frost

period supports decision makers in selecting planting dates as well as in selecting from short to long term crops and

varieties. Where water is available for irrigation, a combination of frost resistant crop varieties and previous clear-sky

night-time irrigation are the key for a success harvest.

Air temperature decrease gradually during radiative frost during night-time but this is not necessary the case

of advective frost, which occurs when low-temperature air masses come from cooler regions by global atmospheric

circulation. These advective frosts do not depend on clear sky conditions, and temperatures decrease drastically in

minutes to hours sometimes creating temperatures below 0°C. There are two important differences between freeze

events: (i) frost and (ii) duration of the event. Air temperature gradually decrease during radiative frosts and then

plant cells begin to release intracellular water, trying to generate heat from the water state change from liquid to solid

(latent heat of fusion: 333 J gr-1 of water). If the frost event is long enough then the plant cells die by dehydration.

10

Page 11: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Irrigation under these circumstances makes available extra energy from water state changes and do not supply

water to the plant. In case of a rapid temperature decreases during advective frost, temperatures fall below 0°C and

the intercellular water will freeze, forming ice which eventually will breaks the cellular membrane. Irrigation and/or air

movement does not help at all. Advective frost kills the crop, and necrotic tissues appear immediately as black-

colored tissues replacing the green-chlorophyll.

Again, knowledge of topography and global atmospheric circulation patterns will support the decision

making when planting different kinds of crop and variety. There are several native crops and varieties that are frost

resistant in some or all their phenological stages; but usually they have low productivity capacity, as characterized

by their shape, sizes, taste or colors with low market values. However, these native germoplasms are an asset to

marginal agriculture because they guarantee food security for farmers.

Examples of resistant crops are native Andean roots and tubers. One of these examples is known as ‘maca’

(Lepidium meyenii Walp.) which is typically cultivated above 4400 m a.s.l. and grow in areas of -1.5°C mean

temperature, reaching -10°C as minimum temperature (Quiroz and Cárdenas, 1997). ‘Oca’ (Oxalis tuberosa) can

grow up to 4100 m a.s.l and can yield from 35 to 55 tons per hectare with adequate management

(http://www.cipotato.org/artc/artc.htm); ‘Ulluco’ (Ullucus tuberosum) and, ‘Mashua’ (Tropaeolum tuberosum) both

contain up to 75% of dry matter in their tubers as well as high levels of isothiocyanates known for their insecticidal,

nematicidal, and bactericidal properties, and can tolerate temperatures from -2 to -4°C (Romero et al., 1989). Native

grains such as ‘Quinoa’ (Chenopodium quinoa Willd.) survive in temperatures up to –8°C and

‘Canihua’ (Chenopodium pallidicaule) survive in temperatures of up to -10°C (Mujica, 1994).

Figure 5: Comparison of simulated potato yield distribution between two sub-species of Solanum under rainfed

conditions and low N fertilization rates. La Encañada – Peru. (a) Solanum tuberosum L. subsp. tuberosum

var. Mariva, and (b) Solanum tuberosum L. subsp. andigenum var. Imilla Negra.

Some cultivated bitter potatoes as Solanum juzepczukii, Solanum ajanhuiri and Solanum tuberosum

andigenus tolerate lower temperatures than commercial varieties. Such bitter varieties are cultivated in high

mountainous niches were commercial varieties can not grow. Figure 5 shows simulated differences in the

11

Page 12: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

distribution and yields between a commercial variety (Mariva) and bitter potato variety (Imilla Negra). Some native

bitter potato varieties resist temperatures as low as -4°C (Solanum chomatophilum, Solanum multidissectum) and

even -12°C (Solanum commersonii; Chen et al., 1976). These species are used as the genetic source of frost

tolerance of commercial potato varieties (Baudo et al., 1996; Cardi et al., 1993; Van Swaaij et al., 1987; Wallis et al.,

1997). These germoplasms are considered a major asset to our societies and fortunately most of them under care of

international and local efforts.

15.3.2.3 Daily temperature swing and thermo regulator agents

Vapor water in the atmosphere store the sensible heat flux, which is directly reflected in the air temperature.

Because of the low level of water vapor at high altitudes and in deserts, the variations in air temperature mainly

depend on the incoming solar radiation. Low water vapor content combined with the high atmospheric transmissivity,

give rise to wide daily temperature swing.

Meteorologically speaking, air temperature is measured at the meteorological shelter. However, the same

body (same albedo), will record large differences in temperatures depending whether it is expose to direct sunlight.

A typical example is observed in towns located in mountainous areas; people walking on footpaths under house’s

shades are usually wearing coats, while people across the street and under direct sunlight conditions wear T-shirts.

According to Toudert and Mayer (2007) while studying thermal comfort in an east-west oriented street canyon under

hot summer conditions in Freiburg (Germany), there was a small increase of air temperature in the irradiated

surfaces of the street canyon. This was due to the direct impact of the incoming solar radiation and the heat gained

because of the geometry and orientation of the canyon. According to this study thermal stress was mostly attributed

to solar exposure, and on average, a standing body absorbed 74% and 26% of heat in the form of long- and short-

wave irradiance respectively. In mountainous regions with low water vapor content, in temperatures are higher under

the effect of direct solar exposure than under shaded surfaces.

Plant metabolism is affected by daily temperature swing and not by mean temperatures. If our goal is

agricultural production, then we must focus on temporal and spatial analysis of maximum and minimum

temperatures. In mountainous regions plants are subject to much higher daily temperature swing than in lower

lands. Maximum temperature can be managed by plants depending on their type of metabolism (C3, C4 and CAM).

Accordingly, plants regulate their internal temperature, but for most of them, 40°C is the threshold when protein

(enzymes) de-naturalization starts.

Crops areas are more affected by low temperature threshold than high temperature threshold in mountains.

The presence of water bodies in mountains is important not only as a water source for irrigation, but at high

altitudes, they modify the daily temperature swing by: (i) increasing the atmospheric water vapor content; and (ii)

capturing energy across the water body profile.

12

Page 13: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Figure 6: Thermo regulator effect of Lake Titicaca in: (a) minimum temperature, (b) temperature swing, and (c)

maximum temperature.

Because of its transparency, sunbeams penetrate deeper in water than in soils. As a consequence, solar

energy is stored in a larger body of mass in water than its equivalent area in soil. Large water bodies absorb large

13

Page 14: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

amounts of energy, which in case of mountainous regions, is released to the atmosphere during night avoiding

extreme minimum temperatures and diminishing radiative frost occurrences. To demonstrate the thermo regulator

effect, annual maximum and minimum temperatures from 16 weather stations around Lake Titicaca (Figure 2) were

analyzed (SENAMHI-MEM, 2003). To diminish the altitudinal effect, all temperatures were standardized to the lake

level increasing its temperature by using the dry adiabatic lapse rate (+9.8 °C km). Distances from each weather

station to the lake were estimated using a geographical information system (GIS). Figure 6 show the scatterplots

and the coefficient of determination between minimum temperatures, temperature swing and maximum

temperatures versus distance. Minimum temperatures decreased linearly with distance to the lake, whereas

temperature swing and maximum temperature increase lineally. Distance from Lake Titicaca explains 73.9%, 72.9%

and 54.7% of the spatial variability of each variable respectively up to a distance of 120 km. However, no more than

5 km around Lake Titicaca can be used for agricultural purposes (François et al., 1999).

In the case of shallow water bodies, the volume of mass where the solar energy is stored could be relatively

small; however, before temperatures drop below 0°C, the change of state of water from liquid to solid presents a

buffer effect in decreasing minimum temperatures due to the energy released during the process (333 J g-1 of water).

15.3.2.4 Favorable use of mountainous temperature knowledge

Knowledge of these facts gives us the possibility to think of solutions for the food production challenge.

Some agriculture management practices make use of the thermo regulator effect and the buffer effect provided by

water. Raised field systems (Figure 7) in a series of elevated soil platforms (up to 1.2 m high and 2-20 m wide)

surrounded by canals (1.6 – 4.5 m wide) flooded wiht water have been utilized in the Lake Titicaca area since

centuries ago (Kolata and Ortloff, 1989; Sánchez de Lozada et al., 1998). These constructions are named

‘Camellones’ in Spanish, ‘Waru-warus’ in Quechua or ‘Suka kollo’ in Aymaras (native languages). These are used to

take advantage of the thermo-regulative effect of Lake Titicaca. This technique has been used to cultivate potato

among other native roots and tubers between 3800 and 4000 meters a.s.l. (De la Torre and Burga, 1986). The crop

temperature is always 1 to 2°C higher on the platforms of the raised fields compared to crop temperatures in the

plains, due not only to the thermo regulative effect of water but also to the higher relative humidity in the air around

the canopy (Lhomme and Vacher, 2002; Sanchez de Lozada et al., 1998). Another similar technology is named

‘Cocha’, which is a natural or artificial land depressions from 0.5 to 5 meter depth and up to 3000 m2 in size (De la

Torre and Burga, 1986). This land depression is used for storing water for crop irrigation but mainly to decrease the

effects of minimum temperatures effects on crops.

Another possibility for diminishing the effect of low temperature in mountainous regions is to increase plant

density. Weak plants will die first, delivering available water from their necrotic tissues, creating the desirable water

change-of-state buffer effect. This practice creates an inter-plant competition for soil nutrients, but increases the

possibility of plant survival, which increases food security. Together with these practices, it is common to mix

commercial varieties that are usually frost susceptible but take up nutrient efficiently, with native frost-resistant

varieties with low productivities. Under good seasonal climate conditions, commercial varieties will uptake most of

14

Page 15: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

the soil nutrients and because they grow fast, they will intercept most of the sunlight. On the other hand, in cold

seasonal climate the commercial varieties will not survive, but native varieties will assure crop yields. Intercropping

is another possibility that in those areas where good seasonal climatic condition exist.

Figure 7: Raised field system (‘Camellón’) outline. Measures taken from Sánchez de Lozada et al. (1998), and

Lhomme and Vacher (2003).

Knowing the differences in the thermal capacity, conductivity and diffusivity between soils and rocks

(Clauser and Huenges, 1995; Sass et al., 1971; Vosteen and Schellschmitdt, 2003), allows farmers to increase the

capacity to store more energy and to deliver it during night. This delivered energy creates a microclimate around the

plant that is capable to diminish the frost effects. Nobel et al., (1992) reported the effect of rocks on soil temperature

and soil water potential, and it relationship with rooting patterns in a desert region. In mountainous areas where

rocks are available, these are placed at the side of the commercial plants. The thermo regulator effects of the rocks

can also be used in stone terraces known as ‘Andenes’ built by the Incas. On these bench terraces plants grow

more vigorously close to the rock wall compared to the ones planted on the opposite side. However, all kind of

terraces with and without stone walls create favorable microclimates close to walls by increasing the area where

incoming solar radiation is stored, delivering the energy during night.

15

Page 16: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Rocks are also used to build break wind walls avoiding excessive evapotranspiration by breaking the leaf

boundary limit. Small areas surrounded by short rock barriers are used to protect sensible crops from wind and low

temperatures (Schreier et al., 2002). Highly valuable commercial crops are usually planted in these small farmyards.

A major strategy against frost events is using topoclimatology, which can be achieve at two different spatial

scales: (i) regional scale by mapping frost risk, and (ii) local and field scale by estimating potential cold air

accumulation.

Regional scale: To help to plan a large-scale frost protection campaign, it is necessary to use a method

that combines long historical records of weather station data with spatially extended data, usually with poor temporal

coverage. A simple method to perform this is obtaining algorithms relating point data (measured at weather station)

with the grid cell data (extracted from satellite or radar images) at the same location. Next the algorithms can be

applied to the entire grid in the images, finally obtaining the minimum temperature maps. With the resulting maps, a

reclassification can be performed based on temperature thresholds in order to find the areas affected by different

intensities frost events (François et al., 1999).

Local and field scale: An important issue to have on mind is that air is considered a fluid. Movement of

high density cold air close to ground affects crops in the form of frosts. Because it is considered a fluid, topographic

depressions are susceptible to be filled by cold air moving over complex orography2. After a frost event it is common

to find crop parcels highly affected by the low temperatures, whereas meters away, unaffected parcels with the

same crops and varieties continue growing. In most of the cases, these highly affected parcels are surrounded by

rock, mud or shrub barriers up to one meter in high. These walls transform the parcel in a swimming pool where cold

air is captured thus exposing the crops to low temperatures for longer periods. Under these circumstances, a cold-

air drainage placed in the lower boundary will lead the cold air moving downslope without negative consequences.

Lindkvist et al. (2000) described the use of land attributes (plane and profile curvature) calculated from

Digital Elevation Models (DEM) to define zones highly susceptible to frost events. In their study performed in the

Scandinavian mountain range ‘the Scandes’, they found very low frost risk in convex terrain and exposed upper

slopes, whereas high frost risk was found at broad concave areas, mainly a part of large ‘U-shaped’ valley bottoms.

‘V-shaped’ concavities were highly affected by frost events due to the high degree of wind shelter and accumulation

of cold air. In the case of the mountainous plateaus surrounded by elevations, air is colder in the open flat areas

than in the hillslopes since dense cold air flows downwards and is accumulated in the lower areas (Collier et al.,

1989).

Mapping frost-risk zones based on potential cold air accumulation will support decision makers from

watershed level to farm level about where to plant. Availability of high resolution DEMs permits to found hotspots

where plants are potentially overexposed to cold air for longer periods. Figure 8 shows an example of terrain

analysis based on a DEM, used for detecting fluxes and accumulation of cold air. A more refined methodology is

presented by Chung et al., (2006). Use of differential GPSs opens the opportunity to combine this methodology with

precision agriculture practices.

2 Orography is the study of the physical geography of mountains and mountain rages.

16

Page 17: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

34373433343534503450

34393450345034503475

34503450347534753475

34503475347535003500

34753500350035003525

AltitudeFlow direction

####

Figure 8: Flow direction determined by grid cell analysis. (a) DEM, (b) DEM’s zoom, and (c) numerical analysis of a

grid cell subset indicating divergence and convergence of cold air. The number of flow direction arrows at

each cell is inversely proportional to the frost risk potential in the cell.

15.3.3. Rainfall

Rainfall has one of the highest spatio-temporal variability of all climatic variables. This is especially true in

mountainous regions, but there is usually a lack of reliable data to cover large heterogeneous areas. In this section

we are describing the main physical processes involved in rainfall formation and distribution in mountains with

emphasis on convective and orographic rainfall. These have especial characteristics that change from local to

regional scales. The rule-of-thumb relationship between rainfall and altitude will also be discussed together with

some ideas of how to construct a weather station network in mountains for increasing it representativeness.

15.3.3.1 Rainfall processes in mountainous regions

17

Page 18: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

There are three main processes responsible for the development of rainfall in mountainous areas (i)

advection, (ii) convection; and (iii) orographic lifting. The first two are common to non-mountainous areas. Advective

rainfall is related to weather fronts moving across large regions due to natural air mass movement around the planet

through the general atmospheric circulation process. These fronts occur, when at least, two air masses face each

other and then produce air lifting, water condensation and rainfall. Convective rainfall, which usually occurs during

summer, is formed by air lifting after an air mass is warmed when it is close to a warm land surface. Convective

rainfall events are locally formed and their formation has a high spatial variability. The land surface warming is more

efficient in flat than in complex topography due to terrain rugosity, which increases the surface area where incoming

solar radiation is stored. Hence, energy storage capacity on flat areas is lower than those presenting rugosity

terrains. The excess energy is used to warm the low atmosphere, increasing air temperature, decreasing air density

and initiating faster the air lifting process in comparison to rough areas. Convective rainfall in plain mountainous

regions is formed at large scale like in the Andean High Plateau (Horel et al., 1989, Garreaud, 1999) and at local

scales (Romero et al., 2007a). Large scale convection in plain mountainous areas can affect the general

atmospheric circulation patterns as in the case in the denominated Bolivian High. This austral summer warm-cored

thermal anticyclone is due to the intense heating of the Andean High Plateau as a result of the incoming solar

radiation with high atmospheric transmissivity area (Baigorria et al., 2004). This high altitude anticyclone modifies

rainfall regimens from the southern Amazon basin to the Central and Southern Andes (Garreaud, 1999). Local scale

convective rainfall events in mountainous regions are highly erosive and combined with steep slopes and erodible

soils, can accelerate the soil erosion rate (Baigorria and Romero, 2007; Romero et al., 2007a and 2007b). These

types of events have large inter-annual variability and, depending on the geographical location, are related to ENSO

phases.

Orographic rainfall events are generated when an air mass is forced to lift when it faces elevation. The

typical example we have is the case of a Pacific Island: a paradise beach in the middle of the ocean with a high

volcano in the middle and a big cloud surrounding the peak. The lee side of the volcano is ignored by many which is

usually dry due to the Foehn effect we will describe in section 15.3.4. In the case of orographic rainfall, there are two

possibilities for their formation: (i) a single peak surrounded by a plain area (usually from volcanic origin) such as the

Kilimanjaro and most of the mountains in East Africa; and (ii) a mountain chain like the Himalayas in Asia, and the

Andes in South America, formed by tectonic plates. The big difference between them is that in the first case, the air

mass has the possibility of partially moving around the mountain, whereas in the case of mountain chains, the air

mass is forced to cross the mountain. In the last case, a partial air mass, which can not cross the mountain chain

and bounce back, converge to the upcoming air mass producing air lifting and rainfall on the mountain windward

side parallel to the mountain chain.

Air masses crossing mountain chains not only generate cloud formation and rainfall in the mountain, but

also generate mountain waves parallel to the lee side of the chain. When an air mass cross a mountain chain, air is

forced to rise in order to pass such obstacles, becoming cooler and denser than the surrounding air, and under the

influence of gravity, sink on the lee side of the barrier. The air then overshoots, and oscillates around its equilibrium

level, forming mountain waves (Barry and Chorley, 1998; Whiteman, 2000). During the lifting process, water vapor

18

Page 19: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

condenses at the crests of the waves, thus forming clouds across and downwind from the mountain barrier. The

amplitude of the waves depends on the initial displacement of the flow above its equilibrium position on the

windward side of the mountain and is directly proportional to the height of the barrier (Queney, 1948). Wavelength is

proportional to wind velocity and increases in value when air stability decreases.

It is important to notice that rain formation is different than rainfall distribution. Rainfall formation processes

are well studied; however, while water drops fall to the land surface they are affected by wind fluxes, which are

especially turbulent in mountainous areas. Rainfall distribution is affected not only by terrain height (Hevesi et al.,

1992), but also by proximity to the moisture sources, terrain relief, and the aspect relative to the direction of the

approaching wind (Blocken et al., 2005; Daly et al., 1994; Marquinez et al., 2003; Mellor, 1996; Whiteman, 2000).

Hence, maps showing rainfall distribution in mountainous regions are of limited value when detailed information

about a particular site is needed, such as in the example of modeling processes such as crop growth or soil erosion.

Therefore, it is necessary to include additional information such as orography and atmospheric circulation patterns at

the time of interpolation in a more physical and dynamic manner (Baigorria, 2005).

15.3.3.2 Rainfall, altitude and orography

Many studies, especially those related to rainfall interpolation, place emphasis in the hypothetical

relationship between rainfall and altitude. These kinds of proxies are often used as an alternative to solve the lack of

information. Generalization of rainfall values, indexes and/or coefficients to large areas without take into account

climatic controls, orography and other factors could have large effect on the end results. The lack of insight in the

use of these proxies does not mean that we can ignore them: ‘The absence of evidence is not the evidence of

absence’ (Sagan, 1997). Nevertheless, environmental sciences tend to use relationships observed over large areas

(small scales3) and implement them in small areas (large scales) at the watershed or even farm level. Especially in

complex mountainous terrain this may result in major errors.

All this does not mean that empiric approaches do not work or that process-based models perform better

than the empiric ones. It rather means that before applying any method, it is necessary to calibrate, validate and if

possible perform an uncertainty analysis of the models as well as the effect of the data resolution. From a scientific

point of view, process-based models are more appreciated than the empiric ones. However, process-based models

are complex to develop, usually require detailed inputs and have inherent problems in their operation related to their

complexity. Although they contribute to the cumulative process of scientific understanding, they do not always

perform better than simple empiric procedures. Perhaps the best approach is to disaggregate the empiric analysis in

components with a biophysical significance (Baigorria, 2005).

Mesoscale analyzes: When we analyze normalized difference vegetation index (NDVI) images (Sellers et

al., 1994) around mountainous chains like the Andes and the Himalayas, relationships between altitude and rainfall

are not directly apparent. Comparing two opposite hillsides on the same mountain, one in the windward and the

3 From the geographical point of view.

19

Page 20: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

other in the lee side, at the same altitude and distance from the peak, larger NDVI values in the windward direction

are detected in comparison to the ones in the lee. Larger amounts of vegetation are identified in satellite images

because of its greenish color, whereas more arid (brownish color) areas are located on the other hillside.

Soil development is due to interaction of the well known soil forming factors (Jenny, 1941). Presence of

different rainfall regimes at both mountain hill-sides influences the moisture regimes of the soil, and then it affects

the soil formation rate. Those parent materials under extremely dry areas with scarcity of water and vegetation may

inhibit soil formation. On the other hand, parent material under moist conditions in warm climates favor to

redistribution of soluble materials which results a well-defined soil profile. As an example, the World Soil Resources

Map (USDA-NRCS, 2005; scale 1:130’000,000) shows highly weathered soils (Ultisols and Oxisols) in the windward

Eastern Andes of Peru; however, the lee western hillside presents young soils as Entisols and in some areas rocky

lands. An analogue case occurring in the Himalayas, where incipient to well developed soils like Inceptisols and

Ultisols, are found in the windward southern hillside, whereas a vast area of young Gelisols in the lee northern

hillside. Both, the Andean and Himalayan windward areas are subjected to wet winds coming from the Atlantic and

Indic oceans respectively. As these mountain chains acts as a natural barriers, orographic rainfall is limited only to

the windward side of the mountain.

According to Barry and Chorley (1998), the relationship between altitude and rainfall is given only in certain

altitudal ranges of the same mountainous hillside. They reported differences in the altitude where the maximum

precipitation is found and how the relationship changes in different regions. In tropical and subtropical zones for

example, maximum precipitation is found below the summit, after this point, precipitation decrease with altitude. As a

conclusion, simple linear relationships between rainfall and altitude perhaps could be valid in the same hillsides

facing the wind, but only at mesoscale level (Baigorria, 2005).

Regional and local analyzes: However, being in the same hillside of a mountain chain does not mean that

the altitude-rainfall relationship can be applied at higher resolutions. Table 2 shows rainfall data collected during one

month in a nine-automatic weather station network in a northern Andean highland watershed in Peru. Distances

between weather stations ranged from 0.8 to 10.3 km, and as expected, the rainfall spatial variability is not explained

by altitude. At daily scale, for example, weather stations at the same altitude (Paulino and Calvario) recorded 21.8

and 0 mm respectively. In another event, the lowest weather station (Manzanas) recorded 14.4 mm whereas the

highest (La Toma) recorded 3.8 mm. For the closest weather stations (Chagmapampa and Usnio) located at 800 m

the maximum differences of 7.1 mm were registered. From these data, no relationship can be established between

amount of rainfall and altitude at high resolutions. However, spatial rainfall variability can be explained by the

complexity of terrain in mountainous areas (Romero et al., 2007a; Whiteman, 2000).

Table 2. Total rainfall received between December 4th.1998 and January 4th 1999 at 9 different weather stations in La Encañada watershed, northern Peru (Adapted from Romero, 2005).

20

Page 21: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Distance between weather stations (km)

ID Weather station Altitude (m a.s.l.)

Rainfall(mm)

B C D E F G H I

A La Toma 3590 203 3.7 2.8 7.5 5.4 4.8 5.7 7.2 6.9B San José 3550 27 6.3 10.3 7.2 6.5 8.2 9.5 6.9C Quinuamayo 3500 298 5.0 4.2 3.9 3.7 5.2 6.9D Sogorón 3400 99 3.6 4.2 2.1 1.7 6.8E Chagmapampa 3300 99 0.8 1.7 2.3 3.4F Usnio 3260 117 2.1 3.0 3.3G Paulino 3250 29 1.5 5.1H Calvario 3250 166 5.2I Manzanas 3020 107

Orographic details are important because the steeper the underlying terrain, the higher the precipitation rate

when air is forced directly up the slope (Whiteman, 2000). Models based on the estimation of orographically forced

vertical motions and advection to simulate orographic precipitation (Barros and Lettenmaier, 1993; Pandey et al.,

2000; Sinclair, 1994; Smith, 2003) better describe this kind of processes on a small scale. Furthermore, airflow

acceleration over the crest of barrier with steep and narrow upwind faces may displace the precipitation maximum to

the lee side of the crest (Daly et al., 1994). However, from the rain formation point of view, orography must not be

seen as a direct effect of changes in the terrain height, but as the changes in the cloud’s path. For instance, air

transversally crossing a deep narrow pass, will not follow the orography falling slope down for finally climb on the

opposite side producing water condensation and rainfall. Otherwise, the air mass will ‘jump’ from side to side,

dissipating the orography effect. However, air crossing along the narrow pass will follow up or down the path

following the orography. Figure 9 shows the same watershed affected by different wind directions, and how the

interaction between orography and wind direction affect the wind flux over the area, and by doing so the air lifting

and orographic rainfall formation.

It is important to notice the difference between wind direction at the surface and at cloud levels. Wind

direction measured at weather station level is strongly influenced by roughness and complexity of the terrain.

Whiteman (2000) makes a recompilation of the most important changes in wind fluxes due to mountains. In our

case, for orographic rainfall, we are more focused on the wind direction at cloud altitude, which can be obtained from

reanalysis data (Kalnay et al., 1996), or from sequence of hourly geo-stationary (GOES) images (Baigorria, 2005).

21

Page 22: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Figure 9: Digital elevation model (DEM) and four simulated digital mountain wave models (DMWM) corresponding

to main wind directions. North (0°), East (90°), South (180°), and West (270°). Adapted from Baigorria,

2005.

15.3.3.3 Monitoring rainfall in mountainous regions

In general, according to Linsley et al. (1977), the sample errors, in relation to rainfall amount, tends to

increase when the average rainfall over an area is increased into the area, and it tends to decrease when the

network density, rainfall duration and area size are increased. According to the World Meteorological Organization

(1970) for hydrometeorological purposes, in the Tropical Mediterranean Mountain regions, one weather station is

recommended for an area between 100 to 250 km2. However, according to data shown in Table 2, the values

recommended were not enough for this complex terrain. Moreover, in other applications as dynamical crop modeling

(Jones et al., 2003) or soil erosion modeling (Nearing et al., 1989), the amount of rainfall is not only important but

also its distribution in time and space. The accuracy of rainfall data and the representativeness of the weather

station network are extremely important due to the impact of the model inputs over the output obtained.

In order to gather rainfall data across the area of interest, the ideal methodology entails on-site collection of

near-surface meteorological data. However, as the size of the target area increases, this approach becomes

22

Page 23: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

prohibitively expensive (Thornton et al., 1997). The use methodologies based on satellite and radar images facilitate

the understanding of spatial variability of the rainfall, however development of algorithm taking into account the

terrain orography make them hardly operative in mountainous regions so far.

There are other alternative ways to estimate the spatial distribution of precipitation: (i) geostatistical

techniques; and (ii) atmospheric modeling. During the last decades, geostatistics have developed different

interpolation techniques based on the spatial correlation between observations and later using correlations with

different terrain attributes (Hevesi, 1992; Kyriakidis et al., 2001; Marquinez et al., 2003). However, in mountainous

regions with a lack of both: weather stations and spatial representativeness, errors in the application of these

techniques are frequent due to the inability of the point data to capture the high variability of the rainfall.

The use of atmospheric models based on physical and dynamic processes are an alternative. General

circulation models (GCMs) operating at large grids scales are limited to resolve small-scale distribution of orographic

precipitation (Baigorria, 2005). Mesoscale models as well as models including orographically induced dynamics

(Barros and Lettenmaier, 1993; Sinclair, 1994; Smith, 2003) need to be initialized from a large-scale numerical

model, radiosonde data, radar data and/or surface observation. These models require substantial amounts of input

data, and even in developed countries where meteorological networks exits, the applications are not detailed

enough to support decision making for watershed and farms.

All the methods to estimate the spatial variability of rainfall are calibrated and validated based on the raw

data measured at weather station, and for mountainous regions it is necessary to increase the representativeness of

the data by planning the network according to certain principles. If the interaction between wind direction and the

mountain hillside aspect are the main source of spatial variation, then transects parallel to the wind direction are

recommended. Weather stations should follow wind direction and not necessarily altitudes.

If we are interest to measure the potential rainfall in the windward hillside, we should find the range where

the condensation level is located during the season we are interested on. Then, the weather station must be located

at this point.

Weather stations must be located in areas of high interannual variability, like in the boundaries of

climatological cyclones and anticyclones. This is because of small changes in the position of the core or its intensity

change the winds direction in the region, carrying different air masses from different regions which finally will

produce orographic precipitation.

And finally, if it is possible and when resources are available, increase the weather station density across

the region. With the new advances in technology, relative cheap portable data loggers that measure specific

variables are now available. A good idea is to support the main weather station network with a secondary network by

using these easy to install equipments. This extra information, which can be temporal and floating, will help us to

understand the rainfall spatial variability in our study areas and give us the capability to resolve the non-random

rainfall occurrence.

23

Page 24: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

15.3.3.4 Favorable use of mountainous rainfall knowledge

In places such as the Xinjiang province (China), where the optimal temperatures poorly overlap the optimal

rainfall regime, a deep irrigation is performed after harvesting at the end of the cropping season. Then, during

temperatures below zero the soil water freezes and is thus stored until the next cropping season. At the beginning of

the next cropping season, when temperatures rise, rainfall events are rare, but the water soil content is at optimum

condition for planting and plants can growth until rainfall events arrive. When they do, rainfed agriculture is usually

performed. Rainfall water is then stored for irrigating at the end of the cropping season to increase the water soil

content for the next cropping season. Some frost resistant roots and tubers that can be planted using non-sexual

seeds are planted at the end of the cropping season, before soils get frozen. Seeds stay in a frozen state for several

months together with the water and soil during all the low temperature season. For the next cropping season these

roots and tubers will begin to grow as soon as the stored soil water will be melted.

Figure 10: Frequency analysis of rainfall according to the wind direction. Watersheds of Chitán and San Gabriel

(Carchi, Ecuador).

Many traditional weather forecast indicators (based on meteorological, astronomical and biological

observations) used by farmers around the world include observation of wind and cloud direction (Baigorria, 2006;

Valdivia et al., 2002). These observations are directly related to the seasonal changes of the general atmospheric

circulation due to the energy balance. In case of mountainous regions, these changes cause the different air masses

to rise and to generate orographic rainfall in different hillsides at different times of the year, producing different

climatic regimes around the mountain. As shown in Figure 10, heaviest rainfall events are generated when wind

blows from the Northeastern, which occurs mostly during winter and spring as shown in Figure 11. This make the

Southwestern facing hillslopes more vulnerable to soil erosion but also creates a different ecosystem by affecting

rainfall regimes and soil forming processes (section 15.4). In different degrees, the remaining hillslopes in this area

are mostly affected by non-erosive rainfall events with comparative lower amount of water occurring mostly during

24

Page 25: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

summer and fall (Figure 11). This last combination tends to creates drier ecosystems with relative slow soil forming

process in comparison to the previous one (Baigorria, 2005). Farmers at San José de Llanga province of Aroma

(Bolivia; 17°22’ S latitude, 67°50’ W longitude, and 3750 m altitude) in the Andean High Plateau, know that winds

coming from the north are related to dry years, while wind from the east are related to rainy years with good yields.

These farmers are climatologically located in the boundaries of the Bolivian High (Horel et al., 1989, Garreaud,

1999), and the wind they observed are the ones that corresponded to the 700 hPa, the altitudinal range they live in.

What they observe is the relative position of the Bolivian High (Valdivia et al., 2002). In doing so, they forecast

incoming air masses from the dry northern Andes or from the wet Amazon basin respectively. Of course this kind of

empiric observation transmitted from generation to generation take time, however, application of the principle of

knowing where our study area is located according to some important climatic control will give us a short cut in the

process.

Figure 11: Seasonal frequency analysis of wind direction. (a) Summer, (b) Fall, (c) Winter, and (d) Spring.

15.3.4. Atmospheric humidity and Foehn effect

When unsaturated air ascend (descend), it cools (warms) at the thermodynamically determined rate of 9.8°C

by 1 km altitude, which is called dry adiabatic lapse rate (γd). This dynamic adiabatic cooling (warming) is caused by

the increase (decrease) in the volume of the air mass due to pressure changes in relation to the altitude. Once the

air becomes sufficiently cold while ascending it can no longer maintain water as a vapor, the vapor condenses

forming cloud droplets. The altitude at which the change of water state occurs is called the condensation level. The

process releases the latent heat of condensation which is defined as the heat gained by the air when water vapor

25

Page 26: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

changes into a liquid (2500 J g-1 of water), which adds to the buoyancy of the air, causing it to rise more rapidly.

From the condensation level, because the air is now saturated (air temperature equal dew point), it cools at a slower

rate, and this is referred to as the saturated adiabatic lapse rate (γs). If the air mass is still rising, air temperature as

well as dew point decreases together following γs, which as a difference with γd, is not constant and vary according to

the pressure and air temperatures. During all the upward movement, rainfall is produced if the water drop weight is

greater than the wind ascending forces. When the mountain summit is crossed, and according to the atmospheric

stability, the air mass begins to go down-slope on the lee side of the mountain by gravitational forces. Then, the air

mass begins to increase its temperature according to the γd; however the dew point remains without change

because there is no moisture gain during the process. This loss off moisture while ascending, followed by the

adiabatic compression and warming as air descend the slope on the leeward side of a mountain range is designated

as Föhn or Foehn effect (Figure 12).

The Foehn effect dissects both sides of the mountain in two different temperature and moisture regimens:

the leeward side is warmer and dryer than the windward side which additionally receives the effects of the

orographic rainfall. A typical example takes place in the Central Andes, where rainfall forest is found at the windward

east hillside while desert are at the leeward west hillside. Table 3 shows two transversal transects across the Andes

where the differences in precipitation can be observed. Of course this area is also influenced by the South Pacific

Anticyclone and the cold Humboldt Current which modify the air temperature (Baigorria et al., 2004).

Figure 12: Hypothetical example of Foehn effect.

At smaller scales dry and warm air masses on the leeward side of mountains increase the crop water

requirements because the vapor pressure deficit increases and in doing so, increase the crop evapotranspiration.

Again crop zoning is the best approach, and it depends on the intensity of the Foehn effect. The higher the

orographic barrier, the higher the adiabatic compression. Then moisture and temperature regimen differences at

both sides depend on the altitude of the condensation level relative to the summit. If the condensation level is not

reached, moisture and temperatures at both sides will remain unchanged. In the case water is necessary but the

condensation level is poorly reached forming only fog (clouds at ground level), water can be obtained by capturing

26

Page 27: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

the fog using condensation surfaces. These surfaces act as filters for moisture, allowing water vapor to condense at

the surface or to aggregate small weightless water drops that can not fall over the ground. These condensation

surfaces are usually dense nets placed perpendicular to the wind path as parallel walls in order to filter as much

moisture as possible. In the desert coasts of Chile and Peru, under the presence of cold water currents, fog’s water

is naturally captured by rocks located in low hill slopes facing the wind. This captured water maintains natural forest

locally called ‘Lomas’ which can include certain salt tolerant trees and shrubs.

Table 3: Two transversal transects across the Andes showing changes of rainfall patterns as the air masses are crossing from west to east.

Weather station Latitude(South)

Longitude(West)

Altitude (m a.s.l.)

Rainfall(mm)

San Ramón – Junín 11°08’ 75°20’ 800 2050Huancayo 12°02’ 75°19’ 3308 765Cosmos 12°09’ 75°34’ 4575 1047A. von Humboldt 12°05’ 76°56’ 238 16

El Porvenir 6°35’ 76°19’ 230 1041San Ramón - San Martín 5°56’ 76°05’ 184 2158Bambamarca 6°40’ 78°31’ 2536 737Chiclayo 6°47’ 79°50’ 31 37

15.4 Application of the spatio-temporal climate variability knowledge in mountainous regions

In this section we are presenting two case studies in mountainous regions where knowledge of the spatio-

temporal climate variability is linked to knowledge provided by other scientific studies. The first case study allows us

to investigate the role of climate in the formation of derived-volcanic ash soils. The second case study investigates

the spatial variability of soil erosion and runoff in a mountainous watershed, process that seriously affects

sustainable agriculture in those areas.

15.4.1 Influence of climate on soil formation: A case study

Many studies showed the importance of altitude zoning when explaining the differences between rate of

pedogenesis proceeds (Zehetner et al., 2003; Vacca et al., 2003). Other studies have confirmed the general

importance of rainfall as a primary factor regulating soil development pathways in volcanic materials (Parfitt et al.,

1985; Nizayimana et al., 1997; Ugolini and Dahlgren, 2002), but few have analyzed the climatic parameters and their

relationship with soil characteristics (Chatwick et al., 2003).

The five main soil-forming factors of parent material, topography, climate, organisms, and time (Jenny,

1941) influence the type of process (physical, chemical, and biological), the duration, and the rate of soil

development at a given location (van Breemen and Buurman, 1998). Parent material and topography define the

initial state where climate and organisms initiate the physical, chemical, and biological soil forming process resulting

27

Page 28: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

over time in the current pedon. In this conceptual model of soil formation, climate is an important driving factor

because it influences weathering rates both directly and indirectly through the type and quantity of plants and

organisms affecting soil organic matter dynamics (Baldwin et al., 1938; Zehetner, 2003; Ugolini and Dahlgren,

2002).

Digital soil mapping use reproducible, quantitative methods that make extensive use of auxiliary information

to spatially predict soil properties (Scull et al., 2003; McBratney et al., 2003). Several methods are rapidly being

developed ranging from geostatistical, geographical information systems (GIS), topographic analysis and remote

sensing (Bell et al., 2000; Mueller and Pierce, 2003; Scull et al., 2003; McBratney et al., 2003; Nanni and Demattê,

2006). Nevertheless, high resolution input data remain a serious constraint for many cases, and this is why little

attention has been directed to climate variability as a driving factor behind soil variability (McKenzie and Ryan, 1999;

King et al., 1999; Ryan et al., 2000; Guo et al., 2006). Process-based interpolated maps (Baigorria et al., 2001;

Baigorria, 2005) provide a new basis for digital soil mapping using climatic variability. Climate as a soil-forming factor

can now be studied at higher resolution and may enable values of soil properties at a particular site to be predicted

by disaggregating soil mapping units and incorporating secondary climate information.

In this case study we are going to explore the predictability of the spatial variation of soil organic matter

(SOM) in a mountainous region as a function of topography and climate variables. The study area was located in the

Ecuadorian Andes between 0°42’ N, 78°30’ W and 0°32’ N, 77°30’ W in the province of Carchi (Figure 13). The area

includes the Chitán and San Gabriel watersheds ranging between 2700 and 3840 m a.s.l. The study area is highly

suitable for a first exploration of the relations between soil properties and short-distance variability because of the

importance of climatic variability in soil genesis.

Figure 13: Location map of Carchi – Ecuador, including altitudes and soil units (adapted from Baigorria, 2005). Soil

units: Duriudoll (Cf, Ck), Andic Argiudoll (Hf), Typic Dystrandept (Dp, Dm), Typic Hudrandept (Dv).

28

Page 29: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Four automatic weather stations recording maximum and minimum temperatures, rainfall, and incoming

solar radiation were installed in the study area and operated for three years. Over the same period, sequences of

infrared images from the GOES satellite were used to derive the main wind direction at the altitude of clouds

formation using daily time steps (based on cloud movement every 15 minutes). All these daily data were used to

produce daily maps of maximum and minimum temperatures, rainfall and incoming solar radiation. From them,

monthly and yearly aggregations yielded high resolution maps (Figure 14) used as a predictors of SOM as a

following step (Figure 15).

Figure 14: Interpolated climate maps: (a) annual average of maximum temperature, (b) annual average of minimum

temperature, (c) annual average of incoming solar radiation; and (d) total annual rainfall (adapted from

Baigorria, 2005).

From the study area, 190-georeferenced soil samples were taken across all the soil units. Values were

extracted for topographic and climatic layer at the soil sample points, forming a database containing SOM,

topographic and climatic data. The database was stratified on the basis of the five main soil units. For each soil unit,

one linear regression was performed to predict SOM as a predictor altitude. In addition, two sets of stepwise multiple

regression models were generated with these databases. The first regression analysis was between SOM and

topographic variables, whereas the second was between SOM and topographic and climatic variables. A stepwise

procedure was carried out for sequentially entering and/or removing independent variables one at a time into a

regression equation in an order that improves the regression equation’s predictive ability. This method is particularly

useful for screening data sets in which consist of many independent variables from where it is possible to identify a

smaller subset of variables that determine the value of a dependent variable. Results in Table 4 shows the extra

29

Page 30: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

value gained for using topographic and climatic variables as predictors for SOM at each soil unit, whereas Figure 15

shows the final SOM map. Other soil characteristics were also analyzed as a function of topography and climate,

which can be found in more detail in Baigorria (2005) and Baigorria et al., (2007).

Table 4: Changes in the coefficient of determination according to the independent variables used to predict the spatial distribution of SOM

Soil Altitude Topographic‡ Topographicunit and Climatic‡

Cf – Hf 0.047ns 0.288** 0.606**Ck – Cf 0.005ns 0.454** 0.898**

Dp 0.283** 0.283** 0.936**Dv – Dm 0.004ns 0.004ns 0.772**

Hf 0.046ns 0.091* 0.939**‡ Adjusted R2

*, ** = Significant at the 0.05 and 0.01 probability levelsns = non significant

Figure 15: Map of soil organic matter obtained using stratified multiple regression models using as

predictors climatic and topographic variables (adapted from Baigorria, 2005).

30

Page 31: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

From the results we conclude that there is an added value in explaining soil the variability within soil

units by using climatic variables as predictors in the volcanic-ash derived soils. The use of these secondary

data supports the refinement of the usually applied altitude-temperature or altitude-rainfall relationships.

Rainfall and minimum temperatures were the climatic variables frequently used to explain spatial variation of

soil characteristics. However, incoming solar radiation and maximum temperature became important

variables in soil units where steep slopes were found. Weights given to each climatic variable changed

across soil units, possibly because different physical, chemical and biological processes involved. Seasonal/

monthly disaggregated information of climate significantly improved the spatial predictability of SOM.

Monthly climatic predictors must be seen as the upper and/or lower bounds where physical, chemical and

biological processes are promoted or restricted. Finally, one should realize that this mountainous area was

an ideal case where climate is likely to govern soil variation at short distances. In other regions, one may

have other soil forming factors as well; however, in mountainous areas, orography represents an advantage

for this kind of applications.

15.4.2 Soil erosion in mountains: A case study

Soil erosion caused by water has been a problem ever since land was first cultivated (Morgan, 1995).

This is especially true in mountainous regions where steep slopes add extra potential energy to the process. It

is known that soil loss is related to rainfall erosivity, soil erodibility, slope of the land and the nature of the plant

cover. Erosion occurs partly through the detaching power of raindrops striking the soil surface and partly

through the contribution or rain to runoff, being rainfall intensity considered the most important characteristic

that influences particle detachment and splash (Hillel, 1998; Morgan, 1995). In mountainous regions of the

humid tropics, erosion is expected to increase with the increment of slope steepness and slope length as a

result of respective increases in velocity and volume of surface runoff. If we add the population pressure living

in those areas, it forces people to farm more marginal lands, often unwisely, so increasing the risk of soil

erosion (Kessler and Stroosnijder, 2005).

Rainfall data are of interest for land use planning because rainfall characteristics such as duration,

frequency and intensity affect the soil erosion process (Schwab, et al., 1993; Whiteman, 2000).

Rainfall can be characterised in many ways, varying from total precipitation in a year, season or other

period, to daily rainfall or totals per rainfall event (Hoogmoed, 1999). However, often a shortage of

water for farming is not the consequence of low annual rainfall but of poor seasonal distribution

(Sivakumar and Wallace, 1991). The response of soil to rainfall in terms of soil loss can be variable.

Dramatic erosion processes can be observed during a rainy season, when heavy but not extreme

precipitation intensities coincide with infrequent high soil moisture conditions in the watershed

(Romero et al., 2007a).

The next study case is related to soil erosion assessment in the Andes, from small plots to watershed

level scale (Romero, 2005). The analysis of rainfall and soil characteristics, as well as other factors that affect

31

Page 32: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

erosion, are important and it constituted the basic requirement for erosion quantification and qualification.

Generally speaking, these kinds of studies are carried out in one location at the time, because the installation

of runoff plots are often expensive, time consuming and do not always represents the regional spatial

variability. Applying erosion models may become important for the analysis of hillslope and watershed

processes and their interactions, and for the development and assessment of watershed management

scenarios (He, 2003).

The study area was located in La Encañada watershed in the northern Andes (Figure 16), which

received between 500 and 1000 mm per year during 1995 to 2000. Three weather stations installed in the

area were used: La Toma, Usnio and Manzanas (Table 2). During the evaluated years, the Climate Prediction

Center (CPC) (National Oceanic and Atmospheric Administration - NOAA, 2006) established the conditions in

the Tropical Pacific area as Neutral years for 1995 and 1996, with a weak incidence of La Niña at the end of

1995; a strong El Niño, from the end of 1997 to the beginning of 1998, and a weak and a strong episodes of

La Niña during the end of 1998 to the beginning of 1999, and from the end of 1999 to the beginning of 2000,

respectively. The main results showed that almost 80% of rainfall events had an average intensity lower than

2.5 mm h-1, and 4% with and average intensity higher than 7.5 mm h-1, and a maximum value of 156.3 mm h-1.

During El Niño year the high intensity events increased, compared to La Niña and Neutral years (Table 5).

However, La Niña was characterised by a large total rainfall, but with low rainfall intensities. Detailed

information can be obtained in Romero et al. (2006a).

The soil parent material mainly consists of limestone, sandstone, siltstone and shale, and the

dominant soils in the area are classified as Entisols (Fluvents), Inceptisols (Ochrepts and Umbrepts) and

Mollisols (Aquolls and Ustolls) in the U.S. Taxonomic Classification System (INRENA, 2002; Figure 16). The

most common soil texture was sandy loam and the organic matter content medium to high (over 2%).

According to this characteristics, measured interrill (Ki) and rill (Kr) erodibility values were low in the area

(Romero et al., 2007b), being the most erodible soils those with the greatest amount of silt and very fine sands

and the most resistant the clayey soils.

32

Page 33: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Table 5. Frequency analysis of rainfall intensity classes for Neutral/El Niño/La Niña years at different locations in La Encañada watershed, north Peru.

Neutral years (1995 and 1996)

< 2.5 mm h-1 2.5 – 7.5 mm h-1 > 7.5 mm h-1

La TomaNo. of events 163 63 23% of total 65.5 25.3 9.2UsnioNo. of events 130 78 16% of total 58.1 34.8 7.1ManzanasNo. of events 179 63 15% of total 69.6 24.5 5.8

El Niño year (1997)

< 2.5 mm h-1 2.5 – 7.5 mm h-1 > 7.5 mm h-1

La TomaNo. of events 151 55 7% of total 70.9 25.8 3.3UsnioNo. of events 99 42 2% of total 69.2 29.4 1.4ManzanasNo. of events 72 19 20% of total 64.8 17.1 18.0

La Niña years (1998 and 1999)

< 2.5 mm h-1 2.5 – 7.5 mm h-1 > 7.5 mm h-1

La TomaNo. of events 355 37 0% of total 90.6 9.4 0UsnioNo. of events 289 21 2% of total 92.6 6.7 0.6ManzanasNo. of events 370 27 5% of total 92.0 6.7 1.2

33

Page 34: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

(a) (b)

(c) (d)

Climaticzones

Figure 16: Spatial information of La Encañada watershed, northern Peru: (a) location map, (b) soil map, (c) climate

map; and (d) slope map (adapted from Baigorria and Romero, 2007).

34

Page 35: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Once the climatic and soil data were collected and analysed all around the watershed, ‘flying erosion plots

systems’ (Romero, 2005; Romero and Stroosnijder, 2002; Stroosnijder, 2003) were installed at four locations within

the watershed to evaluate runoff and soil loss from natural rainfall events. Flying plots consist of a set of portable low

cost materials to delineate and monitor a series of ‘Wischmeier’ type of plots. With flying plots only a few

measurements are used to validate prediction technology, and they can be re-installed several times during one

rainy season for instance to measure the effect of different land cover, soil types, climates and topographies. Then,

the slope angle and soil management were evaluated at each flying plot. With all the collected data; the hillside

version of the Water Erosion Prediction Project model (WEPP; Flanagan and Nearing, 1995) was validated and

calibrated for the study area. The soil loss process at watershed level was assessed by using a tool that integrates

GIS with WEPP (Baigorria and Romero, 2007). Using this interface, we generated runoff and soil loss maps under

different land use scenarios consistent to the current land use (Figure 17). Generation of these maps facilitated the

visualization of the erosion process at spatial and temporal scales according to the actual land use of the watershed.

Areas at risk of runoff and soil loss can be identified from the maps. Although the map does not give the soil loss at

watershed level, it can be used to identify hotspots, thus helping decision makers to formulate recommendations for

soil and water conservation.

Figure 17: Simulated soil loss and runoff maps of La Encañada watershed (adapted from Baigorria and Romero,

2007).

Water coming from rainfall or ice-melting can have different effects on mountain environments. It can be

responsible of the human life and vegetation cover if it is well distributed in a watershed. Soil acts as a “sponge”

which is able to store water for a limited time and for different uses (consumption, evaporation, transpiration,

35

Page 36: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

percolation). But sometimes, extreme rainfall events are poorly distributed in time and space, as they can be

concentrated over few days in a year in one specific location; therefore, the excess of water can be negative for

living species as well as for non-living resources, e.g. soils. If rainfall intensity, for example, exceeds the infiltration

capacity of the soil, runoff will be produced and water and nutrients can be lost. There may be a need to reduce

erosion to control the loss of nutrients and water from agricultural lands to prevent not only pollution of water bodies

but also assure agriculture sustainability for farmers living in the highlands.

As shown previously in this section, soil conservation design follows a sequence of events. Using tools like

modeling and GIS to identify areas susceptible to be eroded in a watershed, erosion can be reduced to an

acceptable rate by recommending some conservation strategies (Morgan, 1995). For example, farming slopes that

are too steep accelerate the erosion process, which is a common practice in developing countries. Under this

condition, a rapid establishment of vegetation is recommended before the first highly erosive rainfall event hit the

area. The challenge is to get this vegetation cover productive in the sort term, to meet the immediate needs of

farmers and the people living the affected area.

Mechanical field practices are used to control the movement of water and wind over the soil surface

(Morgan, 1995). For example: the use of bench terraces consists of a series of alternating shelves and risers and

are employed where steep slopes, up to 30°, need to be cultivated. The riser is vulnerable to erosion and is

protected by a vegetation cover and sometimes protected with stone or concrete. These structures appear to be

reasonably satisfactory as a conservation measure under a wide range of conditions, although they may require high

labor input for construction and maintenance. Case-studies of the effectiveness of bench terraces have been

reported in the loess areas of China (Fang et al. 1981), in clay loam soils of Jamaica under bananas (Sheng, 1981)

and in the Uluguru Mountains in Tanzania (Temple, 1972). In the Andean highlands of Peru and Bolivia, bench

terraces have been built by farmers since the Incas, in order to catch more water from rainfall and avoid runoff with

great success. An alternative conservation practice is the so-called slow-forming terraces which are more common

in slopes up to 14° (Baigorria et al., 2002). They are made by soil embankments across the slope supported by a

wall of stones (slope > 14°) or protected by grasses or bushes (slope < 14°). Its main function is to intercept soil

particles and organic matter carried out by runoff. Other type of terraces are the diversion and retention terraces,

design to intercept runoff and preserve water by storing it on the hillside, respectively, although they are useful for

slope inclination up to 7°.

Contouring rows is another recommended mechanical practice that prevents erosion and improves water

retention in the soil. Its effectiveness varies with the length and steepness of the slope, but usually it cannot be used

in slopes higher than 8°. These contouring row needs to have a slight grade across the slope. In other case, rainfall

may be accumulated around the crops, saturating the soil and so promoting fungus diseases such as potato late

blight. Finally, waterways are another conservation system that collects runoff at a non-erosive velocity to a suitable

disposal point.

36

Page 37: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

15.5 Risk associated to mountain climatic resources

In the context of Environmental Vulnerability (EV) analysis, the environment is understood as the

combination and interaction of natural and human systems; whereas vulnerability is defined as the degree to which a

system is sensitive to and unable to cope with adverse impacts of external stresses and shocks. An EV analysis

should thus include the identification of the external stress factors, how the system responds and interacts to reduce

its exposure, and a search for options to enhance its adaptive capacity to the stress factor (R. Quiroz [CIP], personal

communication).

This section presents an EV analysis in a watershed located in the Northern Andean Highlands of Peru, as a

case study to illustrate the process that could be used in different mountainous regions. The EV analysis was based

on agricultural intensification as the stress factor, and the magnitude by which it affected the interannual potato and

cereal producing areas was assessed. The vulnerability of current potato-cereal systems in this Andean watershed

was assessed based on how the system responds to management scenarios. The vulnerability indicators included:

yield and yield variability, soil erosion, runoff, and nitrogen leaching.

Several processes were modeled in order to generate the geo-spatial data required to assess the

environmental vulnerability of potato, wheat and barley growing areas in the watershed. SUBSTOR-Potato (Ritchie

et al., 1995) and CERES-Cereal (Ritchie et al., 1998) models from the DSSAT Crop System family models (Jones et

al., 2003) were used to simulate yields and nitrogen leaching for potato, and wheat and barley respectively. The

WEPP model (Flanagan and Nearing, 1995) was used to simulate soil erosion and runoff. The DSSAT and WEPP

models were previously calibrated for the study area by Bowen et al. (1999) and Romero (2005) respectively. Using

the same approach as described in section 15.4.2 (Baigorria and Romero, 2007) maps of all the mentioned

variables were produced under two different management scenarios: (i) the base scenario which mimic the current

conditions (rainfed and no external inputs; Sanchez, 2005), and (ii) the intensification scenario considering a N-

fertilization of 200 kg.ha-1 and enough irrigation to supplement rainfall in order to reach field capacity. The purpose of

this scenario was to evaluate the expected environmental cost of agricultural intensification using external inputs in

this mountainous region.

Optimal planting dates for each crop varied according to the location within the watershed and the crop. The

methodology to obtain the optimal planting dates used in this case study can be found at Baigorria (2006). After

simulating both scenarios for each grid cell and crop in the study area, yearly grids of yields, soil erosion and N-

leaching were obtained. The average and standard deviation for each grid cell across the 20 years of historical

record was performed, obtaining as a result, average and standard deviation grids of crop yields, soil erosion (Figure

17a) and N-leaching. To asses the EV analysis, averaged grids from the base scenario were subtracted from the

intensification scenario. Thus, grids of changes (increases/decreases) on yields (Figure 18a), soil erosion and N-

leaching for each crop due to differences in agricultural management were obtained. Finally, soil erosion- and N-

leaching- change grids were divided by it corresponding yield change grid. Thus, grids of soil erosion change per

unit of crop yield change (Figure 18b), and grids of N-leaching change per unit of crop yield change (Figure 18c) for

each crop were obtained.

37

Page 38: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Due to the high spatial variability not only in climate and soils but also in management of mountainous

regions, it is very unrealistic to find resulting EV maps showing large areas with the same results. Because making

decisions in these highly fractionated areas is difficult, the most important in these regions is to use this spatial

information to find the hotspots were degradation occurs. In mountainous regions, soil erosion and pollution (among

other environmental vulnerabilities) can be favorable under certain combinations between climate, topography, soils

and management. As shown in Figure 19, some areas increases productivity by increasing the external inputs;

however in some other areas this will not occur. Depending on the location, some increases in productivity will

increase or decrease soil erosion and or pollution by NO-3. Hence, thresholds can be established in order to locate

these hotspots and to spatially focus practices to preserve the environment within certain limits of degradation or to

turn around the process. Crop zoning in areas with high spatial variation such as mountains, must be focus not only

on increasing productivity but also taking account environmental conditions. In highly vulnerable regions like these,

risks associated to small sources can easily propagate to larger areas. A runoff hotspot under extreme climate

events can be transformed into a mud slide. Intensification of external inputs like N-fertilizers in specific hotspots can

contaminate large water bodies used not only for agriculture purposes. Risk analysis in mountainous regions must

take into account as much resolution as possible due to the potential major implications that can be generated from

few small sources.

15.6 Conclusions

Mountainous regions are highly sensitive to climate events and management practices need to understand

the importance of climate in order to assure food security for the farmers that occupy mountain regions.

The best way to apply the climate knowledge in mountainous regions is to know the general atmospheric

circulation patterns in relation to the topographic conditions and to the mountain hillside orientation. Many factors

influence climatic and weather conditions. An analysis of the advantages and disadvantages of using topographic

and climatic variables is essential if we hope to cope with the problems facing these areas. In this chapter we

presented examples and approaches on how to determine climatically induced hazards, and how to adapt

management practices that help to moderate climatic factors. Some of the examples provided can be used as

recipes whereas others only give ideas of how to use the agrometeorological knowledge more effectively in

mountainous areas to adjust to climate, improve food production and minimize degradation. Mountains are complex

environments and there is no single solution, but if we can improve our predictability of climatic variables we are

better able to improve the livelihood of farmers established in mountains.

38

Page 39: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Potato WheatBarley

(a)

(b)

(c)

-10 – -5-5 – 0

00 – 55 – 10

10 – 15

-2,500 – -1,000-1,000 – -100

-100 – -10-10 – 1010 – 100

100 – 1,0001,000 – 10,000

-5,000 – -500-500 – -50-50 – -5-5 – 0

00 – 5

5 – 5050 – 500

500 – 5,000

tn.ha-1

(tn.ha-1).(tn.ha-1)-1

mg NO3.l-1.(tn.ha-1)-1

Cultivated pasturesNative pasturesBare soilRiver

Figure 18: Environmental vulnerability maps for Potato, Barley and Wheat in the La Encañada watershed under optimal planting dates for each crop and location. (a) Yield difference between scenarios, (b) soil erosion change per unit of yield increase/decrease; and (c) NO3 pollution change per unit of yield increase/decrease (Adapted from Sánchez, 2005).

39

Page 40: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Acknowledgments:

The authors thanks the International Potato Center (CIP), the Peruvian National Service of Meteorology and

Hydrology (SENAMHI), the USAID/SM-CRSP through Grant # 291488, The National Oceanic and Atmospheric

Administration (NOAA) through Grant # NA96GP0239, the International Foundation of Science (IFS) through Grant

# C3083-1, and the global change System for Analysis, Research and Training (START).

References

Baigorria, G.A., 2005. Climate interpolation for land resource and land use studies in mountainous regions.

PhD. thesis dissertation. Soil Science and Geology Laboratory, Wageningen University and Research

Centre, Wageningen, Netherlands. 168 p.

Baigorria, G.A. 2006. Assessing the use of seasonal-climate forecasts to support farmers in the Andean

Highlands. In: M.V.K. Sivakumar and J.W. Hansen (Eds). Climate Prediction and Agriculture:

Advances and Challenges’. Springer-Verlag GmbH & Co. KG, Heidelberg, Germany.

Baigorria, G.A., W.T. Bowen, and J. Stoorvogel. 2001. Estimating the spatial variability of weather in mountain

environments. In: International Potato Center (Ed). Scientist and farmer: partners in research for the

21st century. CIP Program Report 1999-2000.

Baigorria, G.A., and C.C. Romero. 2007. Assessment of erosion hotspots in a watershed: integrating the

WEPP model and GIS in a case study in the Peruvian Andes. Environmental Modelling & Software, In

press. doi: 10.1016/j.envsoft.2006.06.012.

Baigorria, G.A., C.C. Romero, and M. Olivares. 2002. La Encañada and Tambomayo watersheds, Cajamarca,

Peru. Disk 7 of 9. [CD-Rom]. In: Schreier, H., S. Brown, R. Bestbier, and R. Quiroz (Eds.). Himalayan

Andean Collaborative Watershed Project. University of British Columbia/IDRC-CRDI, BC.

Baigorria, G.A., C.C. Romero, J. Stoorvogel, and A. Veldkamp. 2007. Climate variability explaining regional

soil variability in volcanic ash soils in the Ecuadorian Andes. Soil Science Society of America Journal,

XX: XXXX-XXXX.

Baigorria, G.A., E.B. Villegas, I. Trebejo, J.F. Carlos, and R. Quiroz. 2004. Atmospheric transmissivity:

distribution and empirical estimation around the central Andes. International Journal of Climatology,

24: 1121-1136.

Baldwin, M., C.E. Kellogg, and J. Thorp. 1938. Soil classification. Yearbook of agriculture, USDA, Printing

Office, Washington, DC: 979-1001.

Barros, A.P., and D.P. Lettenmaier. 1993. Dynamic modeling of the spatial distribution of precipitation in

remote mountainous areas. Monthly Weather Review, 121: 1195-1214.

Barry, R.G., and R.J. Chorley. 1998. Atmosphere, Weather and Climate. Routledge, New York, NY.

40

Page 41: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Baudo, M.M., L.A. Meza-Zepeda, E. Tapio-Palva, and P. Heino. 1996. Induction of homologous low

temperature and ABA-responsive genes in frost resistant (Solanum commersonii) and frost-sensitive

(Solanum tuberosum cv. Bintje) potato species. Plant Molecular Biology, 30: 331-336.

Bentley, J.W., R. Tripp, and R. Delgado de la Flor. 2001. Liberalization of Peru’s formal seed sector.

Agriculture and Human Values, 18: 319-331.

Bell, J.C., D.F. Grigal, and P.C. Bates. 2000. A soil-terrain model for estimating spatial patterns of soil organic

carbon. p. 295-310. In J.P. Wilson and J.C. Gallant (ed.) Terrain analysis principles & applications.

John Willey & Sons INC.

Blocken, B., J. Carmeliet, and J. Poesen. 2005. Numerical simulation of the wind-driven rainfall distribution

over small-scale topography in space and time. Journal of Hydrology, 315: 252-273.

Borlaug, N.E., and C. Dowswell. 2000. Agriculture in the 21st century: Vision for research and development.

AgBioWorld. [On line] Available at: www.agbioworld.org/biotech-info/topics/borlaug/agriculture.html.

Verified: 9th July 2006.

Bowen, W.T., H. Cabrera, V. Barrera, and G.A. Baigorria. 1999. Simulating the response of potato to applied

nitrogen. p. 381-386. In: CIP Program Report 1997-1998. CIP.

Cardi, T., E. D’Ambrosio, D. Consoli, K.J. Puite, and K.S. Ramulu. 1993. Production of somatic hybrids

between frost-tolerant Solanum commersonni and S. tuberosum: characterizacion of hybrid plants.

Theoretical and Applied Genetics, 87: 193-200.

Chatwick, O.A., R.T. Gavenda, E.F. Kelly, K. Ziegler, G.C. Olson, W.C. Elliot, and D.M. Hendricks. 2003. The

impact of climate on the biogeochemical functioning of volcanic soils. Chemical Geology, 202:

195-223.

Chen, P.M., M.J. Burke, and P.H. Li. 1976. The frost hardiness of several solanum species in relation to the

freezing of water, melting point depression, and tissue water content. Botanical Gazzete, 137:

313-317.

Chung, U., H.H. Seo, K.H. Hwang, B.S. Hwang, J. Choi, J.T. Lee, and J.I. Yun. 2006. Minimum temperature

mapping over complex terrain by estimating cold air accumulation potential. Agricultural and Forest

Meteorology, 137: 15-24.

Clauser, C., and E. Huenges. 1995. Thermal conductivity of rocks and minerals. In: AGU Reference Shelf 3

Rock Physics and Phase Relations. A Handbook of Physical Constants. Pp. 105-125.

Climate Prediction Center (CPC) - National Oceanic and Atmospheric Administration (NOAA). 2006, ‘Cold and

warm episodes by season’. [Online]. Available at:

http://www.cpc.noaa.gov/products/analysis_monitoring/ensostuff/ensoyears.shtml. Verified 21st

February (2006).

Collier, P., A.M.E. Runacres, and J. McClatchey. 1989. Mapping very low surface temperature in the Scottish

Highlands using NOAA AVHRR data. International Journal of Remote Sensing, 10: 1519-1529.

Crissman, C.C., J.M. Antle, and S.M. Capalbo (eds.). 1998. Economic, Environmental and Hearlth Tradeoffs in

Agriculture: Pesticides and the Sustainability of Andean Potato Production. CIP and Kluwer Academic

Publishers. 281 p.

41

Page 42: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Daly, C., R.P. Neilson, and D.L. Phillips. 1994. A digital topographic model for distributing precipitation over

mountainous terrain. Journal of Applied Meteorology, 33: 140-158.

De la Torre, C., and M. Burga. 1986. Andenes y Camellones en el Perú Andino. CONCYTEC, Lima, Peru.

Dosio, G.A.A., L.A.N. Aguirrezábal, F.H. Andrade, and V.R. Pereyra. 2000. Solar radiation intercepted during

seed filling and oil production in two sunflower hybrids. Crop Science, 40: 1637-1644.

Dubayah, R., and P.M. Rich. 1995. Topographic solar radiation models for GIS. International Journal of

Geographical Information Science, 9: 405-419.

Erez, A., and J.A. Flore. 1986. The quantitative effect of solar radiation on ‘Redhaven’ peach fruit skin color.

HortScience, 21: 1424-1426.

Fang, Z. S., P. H. Zhou, Q. D. Liu, L.T. Ren, and H. X. Zhang. 1981. Terraces in the loess plateau of China. pp

481-513. In: R.C.P. Morgan (ed.), Soil conservation: problems and prospects. Chichester, Wiley.

Fellman, J.D., A. Getis, and J. Getis. 1992. Introduction to Geography. Chicago: WCB, p. 66.

Figueroa, S., P. Satyamurti, and P.L. Silva Dias. 1995. Simulation of the summer circulation over the South

American region with an eta coordinate model. Journal of Atmospheric Science, 52: 1573-1584.

Flanagan, D.C., Nearing, M.A. 1995. USDA-Water Erosion Prediction Project (WEPP). WEPP user summary.

NSERL Report No. 10. USDA-ARS National Soil Erosion Research Laboratory, West Lafayette, IN.

Forbes, G.A., M.G. Chacón, M.V. Taipe, and R.J. Hijmans. 2001. Simulating potato late blight in the highland

tropics. pp. 87-93. In: International Potato Center (Ed). Scientist and farmer: partners in research for

the 21st century. CIP Program Report 1999-2000.

François, C., R. Bosseno, J.J. Vacher, and B. Seguin. 1999. Frost risk mapping derived from satellite and

surface data over the Bolivian Altiplano. Agricultural and Forest Meteorology, 95: 113-137.

Garreaud, R.D., 1999. Multiscale analysis of the summertime precipitation over the Central Andes. Monthly

Weather Review, 127: 901-921.

Guo, Y., P. Gong, R. Amundson, and Q. Yu. 2006. Analysis of factors controlling soil carbon in the

conterminous United States. Soil Science Society of America Journal, 70: 601-612.

He, C. 2003. Integration of geographic information systems and simulation model for watershed management.

Environmental Modelling & Software 18: 809-813.

Hevesi, J.A. 1992. Precipitation estimation in mountainous terrain using multivariate geostatistics. Part I:

Structural analysis. Journal of Applied Meteorology, 31: 661-676.

Hillel, D., 1998. Environmental Soil Physics. Academic Press. USA.

Hillel, D., 2006. The Natural History of the Bible: An Environmental Exploration of the Hebrew Scriptures.

Columbia University Press, NY.

Hoogmoed, W.B. 1999. Tillage for soil and water conservation in the semi-arid tropics. PhD thesis dissertation,

Wageningen University, The Netherlands. 184 p.

Horel, J., A. Nahamann, and J. Geisler. 1989. An investigation of the annual cycle of the convective activity

over the tropical Americas. Journal of Climate, 2: 1388-1403.

42

Page 43: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Huddleston, B., E. Ataman, P. de Salvo, M. Zanetti, M. Bloise, J. Bell, G. Franceschini, and L.F. d’Ostiani.

2003. Towards a GIS-based analysis of mountain environments and populations. Environment and

Natural Resources Working Paper No. 10, FAO, Rome, Italy.

INRENA, 2002. Estudio integrado de caracterización de recursos naturales renovables en microcuencas

altoandinas para el alivio a la pobreza en la sierra. Microcuenca La Encañada, Cajamarca. In:

Baigorria, G.A., C.C. Romero and M. Olivares (eds). La Encañada & Tambomayo watersheds, Peru.

University of British Columbia/IDRC-CRDI, BC.

Jaime-Garcia, R., R.M. Davis, and W.E. Fry. Effect of solar radiation on survival of Phytophthora infestans

sporangia under glass and plastic greenhouse conditions. Phytopathology, 89:S36. Publication no.

P-1999-0251-AMA.

Jenny, H., 1941. Factors of soil formation: a system of quantitative pedology. McGraw-Hill Book Company,

NY.

Jones, J.W., G. Hoogenboom, C.H. Porter, K.J. Boote, W.D. Batchelor, L.A. Hunt, P.W. Wilkens, U. Singh,

A.J. Gijsman, and J.T. Ritchie. 2003. The DSSAT cropping system model. European Journal of

Agronomy, 18: 235-265.

Kalnay, E., M. Kanamitsu, R. Kistler, W. Collins, D. Deaven, L. Gandin, M. Iredell, S. Saha, G. White, J.

Woollen, Y. Zhu, A. Leetmaa, B. Reynolds, M. Chelliah, W. Wbisuzaki, W. Higgins, J. Janowiak, K.C.

Mo, C. Ropelewski, J. Wang, R. Jenne, and D. Joseph. 1996. The NCEP/NCAR 40 year reanalysis

project. Bulletin of the American Meteorological Society, 77: 437-471.

Kessler, C.A., and L. Stroosnijder. 2005. Land degradation assessment by farmers in Bolivian mountain

valleys. Land Degradation & Development, 17: 235-248.

King, D., H. Bourennane, M. Isambert, and J.J. Macaire. 1999. Relationship of the presence of a

noncalcareous clay-loam horizon to DEM attributes in a genthly sloping area. Geoderma, 89: 95-111.

Kliewer, W.M. 1977. Influence of temperature, solar radiation and nitrogen on coloration and composition of

Emperor Grapes. American Journal of Enology and Viticulture, 28: 96-103.

Kliewer, W.M., L.A. Lider, and H.B. Shultz. 1967. Influence of artificial shading of vineyards on the

concentration of sugar and organic acid in Grapes. American Journal of Enology and Viticulture, 18:

78-86.

Kolata, A.L., and C. Ortloff. 1989. Thermal analysis of Tiwanaku raised fields in the Lake Titicaca basin of

Bolivia. Journal of Archaeological Science, 16: 233-263.

Kyriakidis, P.C., J. Kim, and N.L. Miller. 2001. Geostatistical mapping of precipitation from rain gauge data

using atmospheric and terrain characteristics. Journal of Applied Meteorology, 40: 1855-1877.

Layne, D.R., Z. Jiang, and J.W. Rushing. 2001. Tree fruit reflective film improves red skin coloration and

advances maturity in peach. Hort Technology, 11: 234-242.

Lee, R., 1978. Forest Microclimatology. Columbia University Press, NY. 276 p.

Lhomme, J.J., and J.J. Vacher. 2002. Modelling nocturnal heat dynamics and frost mitigation in Andean raised

field systems. Agricultural and Forest Meteorology, 112: 179-193.

43

Page 44: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Lindkvist, L., T. Gustavsson, and J. Bogren. 2000. A frost assessment method for mountainous areas.

Agricultural and Forest Meteorology, 102: 61-67.

Linsley, R., 1977. Hidrología para ingenieros. (In Spanish) Second Edition. Ediciones Castilla, Barcelona,

Spain.

Marquinez, J., J. Lastra, and P. Garcia. 2003. Estimation models for precipitation in mountainous regions: the

use of GIS and multivariate analysis. Journal of Hydrology, 270: 1-11.

McBratney, A.B., M.L. Mendonça-Santos, and B. Minasny. 2003. On digital soil mapping. Geoderma, 117:

3-52.

McKenzie, N.J., and P.J. Ryan. 1999. Spatial prediction of soil properties using environmental correlation.

Geoderma, 89: 67-94.

Mellor, D. 1996. The modified turning bands (MTB) model for space-time rainfall. I. Model definition and

properties. Journal of Hydrology, 175: 113-127.

Merzlyak, M.N., A.E. Solovchenko, and O.B. Chivkunova. 2002. Patterns of pigment changes in apple fruits

during adaptation to high sunlight and sunscald development. Plant Physiology and Biochemistry, 40:

679-684.

Mizubuti, E.S.G., D.E. Aylor, and W.E. Fry. 2000. Survival of Phytophthora infestans sporangia exposed to

solar radiation. Phytopathology, 90: 78-84.

Moreno, U. 1985. Environmental effects on growth and development of potato plants. pp 481-501. In: P.H. Li,

(ed) Potato Physiology. Academic Press, NY.

Morgan, R.P.C. 1995. Soil Erosion and Conservation. 198 p. Longman Group, Malaysia.

Mujica, A. 1994. Andean Grains and Legume. pp. 131-148. In: Bermejo, J.E.H. and J. Leon (eds.). Neglected

Crops: 1492 from a Different Perspective. Plant Production and Protection Series No. 26. FAO, Rome,

Italy.

Mueller, T.G., and F.J. Pierce. 2003. Soil carbon maps: Enhancing spatial estimates with simple terrain

attributes at multiple scales. Soil Science Society of America Journal, 67: 258-267.

Nanni, M.R., and J.A.M. Demattê. 2006. Spectral reflectance methodology in comparison to traditional soil

analysis. Soil Science Society of America Journal, 70: 393-407.

Nizeyimana, E. 1997. A toposequence of soils derived from volcanic materials in Rwanda: morphological,

chemical and physical properties. Soil Science, 162: 350-360.

Nearing, M.A., G.R. Foster, L.J. Lane, and S.C. Finker. 1989. A process-based soil erosion model for USDA –

Water Erosion Prediction Project Technology (WEPP). Transactions of the ASAE, 32(5): 1587-1593.

Nobel, P.S., P.M. Miller, and E.A. Graham. 1992. Influence of rocks on soil temperature, soil water potential,

and rooting patterns for desert succulents. Oecologia, 92: 90-96.

Pandey, G.R., D.R. Cayan, M.D. Dettinger, and K.P. Georgakakos. 2000. A hybrid orographic plus statistical

model for downscaling daily precipitation in northern California. Journal of Hydrometeorology, 1:

491-506.

Parfitt, R.L., and A.D. Wilson. 1985.Estimation of allophone and halloysite in three sequences of volcanic soils,

New Zealand. p. 1-8. In E. Fernandez Caldas and D.H. Yaalon (eds.) Volcanic soils: weathering and

44

Page 45: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

landscape relationships of soils on tepha and basalt. Catena Suppl. 7. Catena Verlag, Cremlingen,

Germany.

Peixoto, J.P., and A.G. Oort. 1992. Physics of Climate. American Institute of Physics, New York, NY. 520 p.

Piskolczi, M., C. Varga, and J. Racskó. 2004. A review of the meteorological causes of sunburn injury on the

surface of apple fruit (Malus domestica Borkh). Journal of Fruit and Ornamental Plant Research, 12:

245-252.

Queney, P. 1948. The problem of airflow over mountains: A summary of theoretical studies. Bulletin of the

American Meteorological Society, 29: 16-26.

Quiroz, C.F., and R.A. Cárdenas. 1997. Maca (Lepidium meyenii Walp.). pp. 173-198. In: Hermann, M. and J.

Heller (eds.). Andean roots and tubers: Ahipa, arracacha, maca and yacon. Promoting the

conservation and use of underutilized and neglected crops. 21. Institute of Plants Genetics and Crop

Plant Research. Gatersleben/ International Plant Genetic Resources Institute. Rome, Italy.

Quiroz, R., C. León-Velarde, R. Valdivia, P. Zorogastúa, G.A. Baigorria, C. Barreda, J. Reinoso, M. Holle, and

H. Li-Pun. 2003. Making a difference to Andean livelihoods through an integrated research approach.

pp. 111-122. In: Harwood, R.R., and A.H. Kassam (Eds). Research Towards Integrated Natural

Resources Management: examples of research problems, approaches and partnerships in action in

the CGIAR. Rome, Italy.

Reay, P.F., and J.E. Lancaster. 2001. Accumulation of anthocyanins and quercetin glycosides in Gala and

Royal Gala apple fruit skin with UV-B-Visible irradiation: modifying effects of fruit maturity, fruit side,

and temperature. Scientia Horticulturae, 90: 57-68.

Ritchie, J.T., T.S. Griffin, and B.S. Johnson. 1995. SUBSTOR: functional model of potato growth, development

and yield. p. 401-435. In: P. Kabat, B. Marshall, B.J. van den Broek, A. Vos, and H. van Keulen (Eds).

Modelling and Parameterization of the Soil-Plant-Atmosphere System. Wageningen Pers,

Wageningen, The Netherlands.

Ritchie, J.T., U. Singh, D.C. Godwin, and W.T. Bowen. 1998. Cereal growth, development and yield. p. 79-98.

In: G.Y. Tsuji, G. Hoogenboom, and P.K. Thornton (Eds). Understanding Options for Agricultural

Production. Kluwer Academic Publishers, Dordrecht, the Netherlands.

Romero, C.C., 2005. A multi-scale approach for erosion assessment in the Andes. Tropical Resource

Management Papers No. 61. Wageningen University and Research Centre, Department of

Environmental Science, Erosion and Soil & Water Conservation Group. Wageningen, The

Netherlands. 147 p.

Romero, C.C., G.A. Baigorria, and L. Stroosnijder. 2007a. Changes of erosive rainfall for El Niño and La Niña

years in the northern Andean Highlands of Peru: The case of La Encañada watershed. Climatic

Change, In press.

Romero, C.C., L. Stroosnijder, and G.A. Baigorria. 2007b. Interrill and rill erodibility in the northern Andean

Highlands. Catena, In press. doi: 10.1016/j.catena.2006.07.005.

Romero, C.C., and L. Stroosnijder. 2002. A multi-scale approach for erosion impact assessment for

ecoregional research in the Andes. Paper IV-0-2: Proceedings SAAD-III, Lima, Peru.

45

Page 46: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Romero, F., C. Nieto, and R. Guarca. 1989. Estudio morfológico y agronómico de oca (Oxalis tuberosa Mol.),

melloco (Ullucus tuberosus Loz.) y mashua (Tropaeolum tuberosum R.&P.) en Moyocancha,

Chimborazo. In Spanish. CONUEP-ESPOCH, Riobamba, Ecuador.

Rotem, J., and H.J. Aust. 1991. The effect of ultraviolet and solar radiation and temperature on survival of

fungal propagules. Journal of Phytopathology, 133: 76-84.

Rotem, J., B. Woding, and D.E. Aylor. 1985. The role of solar radiation, especially ultraviolet, in the mortality of

fungal spores. Phytopathology, 75: 510-514.

Ryan, P.J., N.J. McKenzie, D. O’Connell, A.N. Loughhead, P.M. Leppert, D. Jacquier, and L. Ashton. 2000.

Integrating forest soils information across scales: spatial prediction of soil properties under Australian

forests. Forest Ecology and Management, 138: 139-157.

Sagan, C. 1997. The demon-haunted world: Science as a candle in the dark. Random House, Inc. US.

Sánchez de Lozada, D., P. Baveye, and S. Riha. 1998. Heat and moisture dynamics in raised field systems of

the lake Titicaca region (Bolivia). Agricultural and Forest Meteorology, 92: 251-265.

Sánchez, M.L. 2005. Modelamiento geoespacial para el análisis de la vulnerabilidad ambiental bajo diferentes

escenarios de manejo agrícola: Caso La Encañada. MSc thesis. In Spanish, with abstract in English.

Universidad Nacional Agraria La Molina, Lima, Peru.

Sass, J,H., A.H. Lachenbrusch, and J. Munroe. 1971. Thermal conductivity of rocks from measurements on

fragments and its application to heat-flow determination. Journal of Geophysical Research, 76:

3391-3401.

Scull, P., J. Franklin, O.A. Chadwick, and D. McArthur. 2003. Predictive soil mapping: a review. Progress in

Physical Geography, 27: 171-197.

Sheng, T. C., 1981. The need for soil conservation structures for steep cultivated slopes in the humid tropics.

pp 357-72. In: R. Lal and E.W. Russel (eds.), Tropical agriculture hydrology. Chichester, Wiley.

Schreier, H., S. Brown, R. Bestbier, and R. Quiroz. 2002. Himalayan Andean Watershed Comparison. In [9 CD

ROMs]. Himalayan Andean Collaborative Watershed Project. University of British Columbia/IDRC-

CRDI, BC.

Schawb, G., D. Fangmeier, W. Elliot and R. Frevert. 1993. Soil and water conservation engineering. John

Wiley & Sons, Inc.

Sellers, P.J., C.J. Tucker, G.J. Collatz, S.O. Los, C.O. Justice, D.A. Dazlich, and D.A. Randall. 1994. A global

1 degree by 1 degree NDVI data set for climate studies. Part 2: The generation of global fields of

terrestrial biophysical parameters from the NDVI. International Journal of Remote Sensing, 15:

3519-3545.

SENAMHI-MEM (Servicio Nacional de Meteorología e Hidrología del Perú and Ministerio de Energía y Minas).

2003. Peruvian Atlas of Solar Energy. (In Spanish). Project PER/98/G31: Electrification of rural zones

based on photovoltaic energy. Lima, Peru. 115 p.

Sinclair, M.R. 1994. A diagnostic model for estimating orographic precipitation. Journal of Applied

Meteorology, 33: 1163-1175.

46

Page 47: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Sivakumar, M.V.K. and Wallace, J.S. 1991. ‘Soil water balance in the Sudano-Sahelian Zone: need, relevance

and objectives of the workshop’, In: Sivakumar, M.V.K., Wallace, J.S., Renard, C. and Giroux, C.

(eds.), Soil Water Balance in the Sudano-Sahelian Zone.

Smiths, R.B. 2003. A linear upslope-time-delay model for orographic precipitation. Journal of Hydrology, 282:

2-9.

Stroosnijder, L. 2003. Measurement of erosion: Is it possible? Catena, 64:162-173.

Tang, Z., and J. Fang. 2006. Temperature variation along the northern and southern slopes of Mt. Taibai,

China. Agricultural and Forest Meteorology, 139: 200-207.

Temple, P. H. 1972. Measurements of runoff and soil erosion at an erosion plot scale with particular reference

to Tanzania. Geografiska Annaler 54-A: 203-20.

Thornton, P.E., S.W. Running, and M.A. White. 1997. Generating surfaces of daily meteorological variables

over large regions of complex terrain. Journal of Hydrology, 190: 214-251.

Toudert, F.A., and H. Mayer. 2007. Thermal comfort in an east-west oriented street canyon in Freiburg

(Germany) under hot summer conditions. Theoretical and Applied Climatology, 87: 223-237.

Ugolini, F. C., and R.A. Dahlgren. 2002 Soil development in volcanic ash. Global Environmental Research, 6:

69-81.

United Nations (UN). 2003. General Assembly 58th Session A/58/134.

United States Department of Agriculture, Natural Resources Conservation Service (USDA-NRCS). 2005.

Global Soil Regions. Soil Survey Division World Soil Resources. [On line] Available at:

http://soils.usda.gov/use/worldsoils/mapindex/order.html. Verified: 18th July 2006.

U.S. Geological Survey’s EROS Data Center. 1996. GTOPO30. [On line] Available at:

http://edc.usgs.gov/products/elevation/gtopo30/README.html#h1. Verified: 4th August 2006.

Vacca, A., P. Adamo, M. Pigna, and P. Violante. 2003. Genesis of tephra-derived soils from the Roccamonfina

Volcano, South Central Italy. Soil Science Society of America Journal, 67: 198-207.

Valdivia, C., R. Quiroz, P. Zorogastúa, and G.A. Baigorria. 2002. Climate variability, Andean livelihood

strategies, development and adaptation in the Andean region. In: Abstracts: 2002 Spring meeting

AGU/EUG/GS/MSA. Washington DC.

Van Breemen, N., and P. Buurman. 1998. Soil formation. Kluwer Academic Publishers, Dordrecht, The

Netherlands.

Van Swaaij, A.C., H. Nijdam, E. Jacobsen, and W.J. Feenstra. 1987. Increased frost tolerance and amino acid

content in leaves, tubers and leaf callus of regenerated hydroxyproline resistant potato clones.

Euphytica, 36: 369-380.

Vosteen, H.D., and R. Schellschmidt. 2003. Influence of temperature on thermal conductivity, thermal capacity

and thermal diffusivity for different types of rock. Physics and Chemistry of the Earth, 28: 499-509.

Wallis, J.G., H. Wang, and D.J. Guerra. 1997. Expression of a synthetic antifreeze protein in potato reduces

electrolyte release at freezing temperatures. Plant Molecular Biology, 35: 323-330.

Wang, Q., Y. Kakubari, M. Kubota, and J. Tenhunen. 2007. Variation on PAR to global solar radiation ration

along altitude gradient in Naeba Mountain. Theoretical and Applied Climatology, 87: 239-253.

47

Page 48: Applications of Climatic Resources in Mountainous Areas · However, mountain chains and peaks cause the air to lift and then condensation processes in the lower troposphere occurs,

Whiteman, C.D. 2000. Mountain Meteorology: Fundamentals and Applications. Oxford University Press. 355 p.

World Meteorological Organization (WMO). 1970. Guide to Hydrometeorological Practices, 2nd edition. WMO

Technical Papers No 168. 82, pp. III-8-III-11, Geneva, Switzerland.

Zehetner, F., W.P. Miller, and L.T. West. 2003. Pedogenesis of volcanic ash soils in Andean Ecuador. Soil

Science Society of America Journal, 67: 1797-1809.

48