You are on page 1of 18

INTERNATIONAL JOURNAL OF CLIMATOLOGY

Int. J. Climatol. 26: 829–846 (2006)


Published online 23 December 2005 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/joc.1267

DIURNAL PATTERNS OF RAINFALL IN A TROPICAL ANDEAN VALLEY


OF SOUTHERN ECUADOR AS SEEN BY A VERTICALLY POINTING
K-BAND DOPPLER RADAR
JÖRG BENDIX,* RÜTGER ROLLENBECK and CHRISTOPH REUDENBACH
University of Marburg, Faculty of Geography, Laboratory for Climatology and Remote Sensing (LCRS), D-35032 Marburg, Germany
Received 30 May 2004
Revised 30 May 2004
Accepted 25 August 2005

ABSTRACT
The diurnal precipitation dynamics in an east-west-oriented valley that connects the Amazon lowlands and the inter-
Andean basin of southern Ecuador (Rio San Francisco valley) is investigated by means of a K-band rain-radar profiler
(located at the ECSF research station, latitude: 3° 58’S, longitude: 79° 4 W) and additional remotely sensed data. A pre-
dawn/dawn (5:30–6:30 LST) maximum of rainfall is found and a secondary peak is observed after noon (14:30–15:30
LST). Although the frequency distribution of rain rates reveals that a great portion of rainfall is of stratiform character,
vertical profiles of rain rate and droplet concentration points to the important contribution of embedded convection and/or
showers produced by local heating for the overall amount of rainfall. Specific differences in stratification and process
dynamics could be found for both peak times. The pre-dawn maximum can be related to mesoscale instabilities over the
Peruvian Amazon close to the south Ecuadorian border. Extended cold air drainage flow from the Andes and low-level
confluence due to the concavity of the Andean chain in this area leads to convective instability in the nocturnal Amazonian
boundary layer, which is extended to the study area by the predominant easterlies in the mid-troposphere. Rain clouds
with at least embedded shallow convection can overflow the bordering ridges of the San Francisco valley providing rains
of higher intensity at the ECSF research station. On the contrary, the afternoon convective precipitation can be caused by
locally induced thermal convection at the bordering slopes (up-slope breeze system) where the ECSF station profits from
precipitation off the edge of these local cells due to the narrow valley. Copyright  2005 Royal Meteorological Society.

KEY WORDS: Ecuador; Andes; radar profiler; rainfall

1. INTRODUCTION

Plant ecologists are especially interested in altitudinal gradients of precipitation because the vertical zonation
of tropical alpine vegetation (e.g. cloud forests) sometimes seems to mirror the amount of rainfall (e.g. Richter,
2003; Sklenář and Lægaard, 2003), keeping in mind that a considerable quantity of water can be provided
through cloud/fog water deposition to the vegetation (e.g. Ataroff, 1998). Diurnal patterns of precipitation
are especially important because nocturnal rainfall is normally less affected by evaporation losses (e.g. from
the leaf surface) than by precipitation during daylight. The diurnal course of rainfall in different altitudes
of tropical high mountain valleys and its mechanism of formation has been discussed since the late 1960s.
However, just a few local studies have been available from the tropical Andean valleys up to today, mostly
due to the sparse network of operational raingauge stations (e.g. Barry, 1992; Barry, 2001; de Angelis et al.,
2004a) that rarely provide data in a high temporal resolution of better than 1 h. Furthermore, the sensitivity
of the gauges is often not high enough to register trace precipitation without greater errors, and this, in fact,
can be the dominating weather situation in higher altitudes (refer to Groisman and Legates, 1994).

* Correspondence to: Jörg Bendix, University of Marburg, Faculty of Geography, Laboratory for Climatology and Remote Sensing
(LCRS), Deutschhausstr. 10, D-35032 Marburg, Germany; e-mail: bendix@staff.uni-marburg.de

Copyright  2005 Royal Meteorological Society


830 J. BENDIX, R. ROLLENBECK AND C. REUDENBACH

The early pioneering work of Weischet (1969) points to several specific types of diurnal patterns in the
tropical Andes. He stated in his summarising paper for the eastern cordillera of the tropical Andes that
precipitation increases with altitude between 2600 and 3300 m asl. These results were in contrast to the
situation of narrow mountain chains or isolated summits where rainfall decreases with altitude from a
maximum zone of precipitation that lies between 900 and 1400 m asl for humid tropical slopes. In the
presence of intra-montane highlands like the inter-Andean basins of Ecuador that are surrounded by mountain
chains (western and eastern cordillera), Weischet (1969) concluded that the increase of precipitation is due
to a secondary rainfall maximum (‘Randschwellen-Maximum’ between 2600 and 3300 m asl), which is
mainly caused by heating effects of the basin. Resulting convection promotes rainfall on the bordering slopes,
additionally triggered by the up-slope breeze system and/or barrage effects of the easterlies at the windward
side (eastern slopes of the east cordillera) which is generally more moist than the western slopes of the east
cordillera (for Ecuador refer to Vuille et al., 2000; Bendix and Rafiqpoor 2001).
A specific deviation from this general behaviour could be observed for steep and narrow valleys that
breach the eastern cordillera chains in an east–west direction and connect the Amazon lowlands with the
inter-Andean basins (Troll, 1968; Weischet, 1969). Rainfall sometimes strongly decreases in the lowest parts
of such valleys where the distance between the valley bottom and the bordering ridges amounts to a few
thousand meters. In extreme situations, local ‘desert islands’ can be the result. The occurrence of ‘dry valleys’
was explained as an effect of the valley-breeze system (Troll, 1968). The up-slope breeze during daylight
promotes cloud formation at the ridges, but the countering branch of the valley circulation causes divergence
and cloud clearance over the valley. However, observations showed that rainfall in these valleys also occurred
during daylight. Weischet (1969) derived the hypothesis that this precipitation originated in the upper cloud
layer of the bordering ridges, which is significantly reduced by evaporation losses on its way to the valley
bottom. Accounting for the diurnal course of precipitation, several authors stated that the valleys should also
be characterised by nocturnal rainfall. This assumption was made on the basis of the reversal of the valley
circulation at night with a confluence of the down-slope breeze and resulting convection over the valley bottom
(e.g. Trojer, 1959). However, Weischet (1969) concluded that the nocturnal down-slope system cannot be the
driving factor of productive precipitation on a local scale. On the one hand, the nocturnal atmosphere in a
valley with a weak inclination to the foothills is characterised by a stable atmospheric stratification that does
not promote the formation of precipitation. On the other hand, the nocturnal cold air drainage flow in a valley
with great inclination to the foreland is directly channelised to a down-valley wind that suppresses confluence
at the valley axis. At least, no convincing dynamic explanation could be given for the occurrence of nocturnal
precipitation.
More recent research could show that precipitation in tropical high mountains is due to a complex interaction
of atmospheric processes and external factors on several scales. Lin (2003) and Kirshbaum and Durran
(2004) generally classify orographic rainfall and mechanisms of convective initiation in several categories
mainly because of atmospheric stability, where convective orographic rainfall needs buoyant instability in
orographic clouds: (1) The primary mechanism for orographic precipitation is the smooth stable ascent over
mountains (stable ascent hypothesis). Rainfall originates from relatively shallow orographic cap clouds.
(2) The seeder–feeder mechanism can intensify rainfall in multiple upstream layer-situations (Kirshbaum
and Durran, 2004), where small raindrops of a near-surface cloud (feeder) formed by orographic lifting are
washed out by raindrops from a higher seeder cloud (Lin, 2003). (3) The potential instability of the layer
that will be lifted to saturation over the mountains can cause embedded shallow convection with moderate
rainfall intensity in orographic clouds. This type of shallow cellular convection is mostly initiated by a vertical
displacement of air parcels, e.g. because of gravity waves, and promoted by a sufficiently long resident time
of the parcels in a cap cloud of sufficient thickness and the absence of wind shear. The embedded cells
are mostly less than 2 km higher than the top of the cap cloud. (4) During latent instability, air parcels are
lifted to the level of free convection, which enables the development of deep convective storms in orographic
clouds.
Several external factors can modify the formation of orographic precipitation and the initiation of
convection, which are especially local and larger-scale weather phenomena in interaction with the topographic
situation (Lin, 2003; Kirshbaum and Durran, 2004):
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
DIURNAL PATTERNS OF RAINFALL IN A TROPICAL ANDEAN VALLEY 831

On the local scale, (1) sensible heating of mountain slopes can cause local up-slope circulations in a
quiescent atmosphere releasing (thermally induced) instability. Cumuli and thunderstorms over peak areas
are the results that induce orographic precipitation. (2) Orographic rainfall of convective character is also
possible due to pure but intensive orographic lifting (orographically induced release of instability). Rainfall
is situated over the peak areas in case of strong blocking by a steep topography, or on the leeward side in
case of weak blocking. (3) Lee-side enhancement of convection with precipitation is possible in the case of
isolated mountains.
On larger scales, the approach/influence of a weather system to/on mountains can modify the circulation
and the convective system and trigger heavy precipitation. (4) Lin et al. (2001) stressed that heavy orographic
precipitation formation is frequently associated with a very moist low-level jet (LLJ) impinging the mountains,
mesoscale low-level confluence and a larger-scale upper-tropospheric-deep short-wave trough that helps
enhance the upward motion of the LLJ. (5) Houze (2004) points to the influence of Mesoscale Convective
Systems/Complexes (MCS/MCC) on the formation of orographic precipitation, especially in combination with
LLJ activity on the slopes of high mountains as, e.g. the Andes of South America.
The aim of the present paper is to examine the diurnal patterns of rainfall in an Andean valley that breaches
the eastern cordillera of southern Ecuador. The work was performed in the scope of a joint ecological research
project (Beck and Müller-Hohenstein, 2001) and is mainly based on the data of a vertically pointing Doppler
rain-radar profiler.

2. STUDY AREA AND DATA

The wider study area of the ecological research group is located in southern Ecuador and comprises an
altitudinal range between 800 and 3600 m asl (Figure 1). A characteristic feature in this area is the Andean
altitudinal depression where the eastern and western cordillera are not as clearly established as they are
in northern and central Ecuador. The main watershed between the drier valleys in the west and the moist
eastern slopes is the Cordillera Oriental de los Andes with altitudes up to ∼3600 m asl. The central study
area covers the Reserva Biósfera de San Francisco where the basis for research is the Estación Cientı́fica de
San Francisco (ECSF) station which is located in the valley of the Rio San Francisco (latitude: 3° 58 18 S,
longitude: 79° 4 45 W, altitude: 1860 m asl), in the upper reaches of the Rio Zamora. Figure 1 shows a
longitudinal section of the Rio San Francisco valley and a cross section at the ECSF station. The main
watershed between the valley and the basin of the provincial capitol of Loja in the west is marked by a
pass-top altitude of ∼2900 m asl (El Tiro). The valley declines to ∼1600 m asl down to the confluence with
the Rio Zamora at a distance of ∼14 km. At the research station, the valley has a ridge-to-ridge width of
about 7 km and the depth from the valley bottom to the highest summit of the southern ridge (Cerro de
Consuelo) is about 1.3 km.
Specific meteorological equipment is installed in the study area of the joint research project. The main
instrument of the current study is a mobile K-band Doppler rain-radar profiler (Klugmann et al., 1996), the
Micro Rain-Radar MRR (version 3). Eight automatic weather stations including tipping-bucket rain gauges are
operated by one sub-project along an altitudinal transect and provide hourly data on standard meteorological
elements (Richter, 2003). Three principal stations of the national Ecuadorian weather service Instituto Nacional
de Meteorologiá e Hidrologı́a, (INAMHI) complete the observations in the area. An X-band local area weather
radar (LAWR) is installed at the Cerro de Consuelo, which provides raw counts of radar reflectivity for the
entire study area every 5 min with a spatial resolution of 500 m and a range of 60 km (Jensen, 2002; Bendix
et al., 2004). A HRPT satellite-receiving station that is installed at the INAMHI headquarters in Quito supplies
NOAA-AVHRR images of the study area (Bendix et al., 2003b).
MRR-3 data were acquired in the wider study area between December 2001 and April 2003, mostly with
1 min temporal and 269 m vertical resolution (30 height levels in the free atmosphere). Most of the time, the
MRR-3 was operated at ECSF where 175 daily data sets could be taken, which are the basis of the current
study. Data for an additional 20 days are available from El Tiro, and 102 measurements have been performed
at various remote locations for the calibration of the X-band radar (LAWR).
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
832 J. BENDIX, R. ROLLENBECK AND C. REUDENBACH

80iW 78i 76i C longitudinal section D


2900 El Tiro

2700
0i Quito

2500

altitude [m asl]
2300

2i Guayaquil 2100

1900

1700
ECSF
4 iS LOJA
ZAMORA 1500
1 2 3 4 5 6 7 8 9 10 11 12 13 14
CHINCHIPE 100 km
distance [km]

A cross section B
A D 3300
3100 Cerro de Consuelo
2900
Sabanilla

altitude [m asl]
ECSF 2700
2500
2300
2100
El Tiro
1900 ECSF
C B MRR-3 Rio San Fransisco
1700
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6 6.5 7
distance [km]

Figure 1. Location and topographic features of the study area

3. THE MRR SYSTEM

The vertically pointing FM-CW-Doppler (FM-CW = frequency modulated continuous wave technology)-
MRR operates in the K-band (24 Ghz, 12.5-mm wavelength) and transmits a FM-CW signal along a vertical
path of the free atmosphere over the antenna. Technical set-up and general signal processing is described
in detail in Klugmann et al. (1996), Löffler-Mang et al. (1999) and Peters et al. (2002). The MRR consists
of a radar module, a specific signal processing card and a parabolic antenna. The radar module is directly
attached to the feed of an offset parabolic reflector antenna with 50-cm efficient aperture diameter. The mean
transmit power is 50 mW with a beamwidth of 3° . The low weight of a few kilograms for the whole system
also allows mobile application in a remote and inaccessible mountainous area. Data storage and processing
is performed on a standard notebook by using the RS232 interface for communication.
The basic parameter of a vertically pointing Doppler radar with FM-CW technology is a vertical profile
of the velocity spectrum of falling hydrometeors (Löffler-Mang et al., 1999), which is derived from Doppler
shift and radar reflectivity. Derived parameters are the average drop spectrum N(d) (m−3 ·mm−1 ), the mean
fall velocity (m·s−1 ), rain rate (mm·h−1 ), liquid water content (g·m−3 ) and apparent radar reflectivity (dBZ),
as observed by a conventional scanning C-band weather radar (Peters et al., 2002, for an example refer to
Figure 2). The vertical profile of all derived parameters consists of 30 levels with a user-controlled nominal
resolution between 10 and 269 m and the corresponding range (maximum: 30.269 m to ∼8 km). The MRR-3
at the ECSF station was completely operated in the maximum range mode with a vertical resolution of 269 m.
The advantage of the MRR system is a higher sensitivity at low rain rates, which are clearly underestimated
by rain gauges and the ability to provide vertical profiles of rainfall.
There are some known limitations of the MRR technology that could worsen the quantitative rain
measurement. Accounting for the implemented function of Gunn and Kinzer (1949) between particle diameter
and terminal fall velocity, only droplets with diameters of 0.25–6 mm can be detected (Löffler-Mang et al.,
1999). However, serious errors would only occur for very high-intensive rain events that are rare in the study
area. A second source of error is due to droplet shape and the kind of hydrometeor because data processing
presumes spherical water drops. The deviation of larger oblate drops (Ø > 1 mm) from the spherical shape
can lead to errors of up to 10%. Greater errors occur for other hydrometeors such as snowflakes and hail.
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
DIURNAL PATTERNS OF RAINFALL IN A TROPICAL ANDEAN VALLEY 833

8070
7532 radar reflectivity [dBZ]
<10.05
6994
11.93
6456 13.80
5918 15.68
5380 17.55

[m above ground]
19.43
4842 21.30
4304 23.18
3766 25.06
26.93
3228 28.81
2690 30.68
2152 32.56
34.43
1614
36.31
1076 38.18
538 ‡40.06
0 -1
rain rate [mm h ]
7532
<0.55
6994
2.71
6456 4.87
5918 7.03
5380 9.18
[m above ground]

11.34
4842
13.50
4304 15.66
3766 17.82
3228 19.98
22.14
2690
24.29
2152 26.45
1614 28.61
1076 30.77
32.93
538 ‡35.09
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23

Figure 2. Radar reflectivity and rain rate at ECSF, 2 January 2003

All those problems are not essential for the study area because snow and hail as well as large droplets are
rare, at least in the lower levels of the radar beam. Bright bands in greater altitudes are a principal problem.
To exclude this from data evaluation, only the lowest 10 levels above the ECSF station (<4490 m asl) are
exploited because the 0 ° C layer in the free atmosphere over Ecuador is nearly constant between 4750 and
5050 m asl throughout the year (after radiosonde data from Guayaquil, Chu and Hastenrath, 1982). The major
source of error is the measured velocity, which represents the terminal velocity as shifted by the vertical wind
velocity (Peters et al., 2002). Because no information on vertical wind speed is available, the influence of
vertical wind and turbulence is neglected during signal processing.
The accuracy of rain rate retrieval also depends on microwave attenuation by raindrops in a precipitating
cloud. While this effect is of minor importance for the near-surface layers of the MRR (Löffler-Mang
et al., 1999), it increases with vertical path length through a rain area. This would lead to a continuous
underestimation of rain rates with height by using the uncorrected radar reflectivity. Matrosov (2005) could
show that a linear relation between rainfall rate and attenuation exists for the Ka -band. He used a linear
correction function for the determination of rain profiles from a Ka -band radar profiler and could show that
variability in attenuation contributes little to errors in rainfall rates aloft (∼10%). A similar linear correction
procedure for attenuation is applied during MRR calibration, where a coefficient of 0.035 dB km−1 ·h·mm−1
is applied. The coefficient was derived during a 3-year field campaign (Zingst, Germany; refer to Peters et al.,
2002), mainly by comparing MRR profiles and rain rates from a C-band weather radar of the German Weather
Service at different altitudinal levels.

4. AVERAGE STRUCTURE OF RAINFALL IN THE VALLEY OF THE RIO SAN FRANCISCO

4.1. Seasonal and diurnal patterns of rainfall in Ecuador and the wider study area
The seasonal patterns of precipitation vary in different regions of Ecuador. The coastal lowlands and the
inter-Andean basins of northern and central Ecuador are characterised by two maxima in February, March
and April, (FMA) and October, November and December, (OND) respectively (Bendix and Lauer, 1992).
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
834 J. BENDIX, R. ROLLENBECK AND C. REUDENBACH

However, the OND maximum is not established in the drier coastal area of southern Ecuador. The Ecuadorian
Amazon region is characterised by humid conditions throughout the year, but a third maximum in JJ occurs
close to the foothills of the eastern slopes. The eastern slopes of northern and central Ecuador (∼900–3000 m
asl) are an interesting transition zone between the highland and the Amazonian type, with a distinct maximum
during June and July and lower rainfall during FMA/OND because convective activity over the Andes and
the Amazon during the latter phases requires compensating divergence over the eastern slopes (Bendix and
Lauer, 1992).
The diurnal course of rainfall in Ecuador also reveals regional differences that are partly comparable to
the situation as described by Weischet (1969). The station of Quito (inner-Andean basin) shows a distinct
rainfall maximum in the afternoon with peak precipitation at 16:00 LST. Precipitation in the coastal lowlands
(station Babahoyo) peaks between 0:00 and 2:00 LST (Zimmerschied, 1958). Data from Richter (2003) reveal
a different behaviour for the wider study area around the ECSF station. The phase of maximum precipitation
in Vilcabamba (1972 m asl), which is located in the lee of the main Cordillera Oriental del Los Andes, starts
in the late afternoon (16:00 LST) and lasts until night (23:00 LST) for the rainy season (December to May).
Additionally, a phase of weak rainfall is also observed during the early morning hours (2:00–6:00 LST).
A station in the Páramo de Cajanuma (3401 m asl), which is situated on the ridge of the main Cordillera
Oriental del Los Andes, is characterised by maximum precipitation in the late afternoon (16:00–20:00 LST),
but precipitation is frequently present throughout the day, especially in the rainy season between January and
June. On the windward side of the main cordillera, the rainy season is shifted to the austral winter (April
to July). The station El Libano (1986 m asl) southwest of the provincial capitol Zamora reveals a distinct
rainfall maximum between 12:00 and 16:00 LST, but scattered precipitation is also observed during the night
from 0:00 to 6:00 LST.

4.2. Diurnal patterns of rainfall at ECSF


Figure 3(a) presents the average rain amount and rain frequency at 30-min intervals for the whole data set
(175 days) at the ECSF station. Average rainfall is generally low throughout the day. However, two clear
maxima can be recognised with a main peak around sunrise (5:30–6:30 LST) and a secondary maximum
can be recognised in the early afternoon hours from 14:30 to 15:30 LST. A third period of above average

(a) 0.20 0.2


0.18 average
relative frequency
0.16 frequency 0.15
0.14
[mm]

0.12
0.1
0.10
0.08
0.06 0.05
0.04
0.02 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
LST [h]

(b) 8 0.3
mean SD frequency mean SD rainfall mean maximum rainfall
7 0.25
relative frequency

6
5 0.2
[mm]

4 0.15
3 0.1
2
1 0.05
0 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
LST [h]
sunrise sunset

Figure 3. (a) Average precipitation and relative rain frequency for 30-min intervals and (b) variability of rain amount and rain frequency
in the near-surface layer at ECSF

Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
DIURNAL PATTERNS OF RAINFALL IN A TROPICAL ANDEAN VALLEY 835

precipitation is observed at 23:00–23:30 LST. As expected, the diurnal course of rain frequency coincides well
with the temporal patterns of average rain amount with slightly enhanced values for the times of maximum
precipitation. On average, approximately 10% of all minutes (3 per 30 min interval) is rainy. Despite the low
average amounts, a distinct variability of precipitation dynamics in the data set is indicated by Figure 3(b).
The mean maxima of rainfall hit the peak times of the average sums, but notable deviations are detected for
the evening and early night situations. The relatively low rainfall frequency and the high maximum at e.g.
17:30–18:00 or 20:30–21:00 LST points to the occasional occurrence of stronger convective events, triggered
by the up-slope breeze system. The precipitation peak around sunrise reveals the greatest variability in rainfall
and rain frequency being characterised by different stratiform and convective mechanisms of rainfall formation.
The times of maximum rain rate (Figure 4) generally agree with the diurnal course of rain amount as
presented in Figure 3. The average rain rate over the day points to a high frequency of stratiform precipitation
that is, nevertheless, significantly greater in amount than the rain rate of a typical drizzle from continental
shallow stratus (<0.1 mm·h−1 , Vali et al., 1995). The result is supported by the investigation of Tokay and
Short (1996), which could show that ∼50% of the rainy period, which is dominated by stratiform dynamics,
reveals rain rates below 1 mm·h−1 .
In the marine tropics, Tokay et al. (1999) found an average stratiform rain rate of 1.86 mm·h−1 by using
1 min resolution disdrometer observations. This fits the average situation of the study area although it is located
in complex tropical topography. For convective events, Tokay et al. (1999) reported rain rates of 10.5 mm·h−1
on an average. Additionally, they distinguished further types of precipitation, which are also proven to be
significant for rainfall formation in complex terrain by Kirshbaum and Durran (2004). Mixed stratiform-
convective precipitation is characterised by average rain rates of about 2.4 mm·h−1 , which agrees quite well
with the early morning situation at ECSF. Kirshbaum and Durran (2004) could show that embedded shallow
convection provides a great variance in average rain rate ranging from 1.74 mm·h−1 in a more stratiform cap
cloud and 5.36 mm·h−1 in a more convective case. The mean maximum rain rates of Figure 4(a) point out that
convective or mixed rain events must also be part of rainfall dynamics in the Rio San Francisco valley. The
highest rain rates in the data set were measured on 6 January 2003, 8:39 LST (36.7 mm·h−1 ) and 16 January
2003, 12:04 LST (37.4 mm·h−1 ) and mark convective situations with a rain amount of ∼1.2 mm per min.

(a) 3.5 18
average rain rate

mean maximum rain rate [mm h-1]


16
3 mean maximum rain rate
14
mean rain rate [mm h-1]

2.5 12
10
2
8
1.5 6
4
1
2
0.5 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
LST [h]

(b) 0.6 1
0.5 06 LST 0.8
cumulative
frequency

0.4 15 LST
0.6
0.3 day
0.2 day 0.4
0.1 0.2
0 0
<01 1<2 2<3 3<4 4<5 5<6 6<7 7<8 8<9 9<10 >10
rain rate [mm h-1]

Figure 4. (a) Average and mean maximum rain rate for 30-min intervals (only minutes with rain) and (b) frequency distribution of rain
rates in the near-surface layer at ECSF

Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
836 J. BENDIX, R. ROLLENBECK AND C. REUDENBACH

The frequency distributions of the rain rates reveal that approximately 60% of all minutes with rainfall have
clear stratiform character (Figure 4(b)). In comparison to the daily mean, both periods of peak rainfall tend
to have slightly higher rain rates. The afternoon peak (15:00 LST) shows a dominant position at moderate
levels between 2 and 4 mm·h−1 . The pre-dawn/dawn peak (5:30–6:30 LST) reveals a surplus of greater rain
rates, especially in the segment of >10 mm·h−1 . Rain rate is also related to the average droplet spectrum
(Figure 5).
Generally, the average 24-h spectrum is relatively broad in comparison to warm tropical maritime
spectra of convective precipitation as published by Tokay and Short (1996) and Tokay et al. (1999). It
resembles more the average warm stratiform spectrum of Tokay et al. (1999) but reveals significantly
higher droplet concentrations for smaller droplets than could be expected for orographic rainfall in high
mountains. Droplet concentration looks more like data that are measured for mid latitude cold advective
clouds (Diem and Strantz, 1971; Strantz, 1971). However, the spectrum is not constant throughout the
day. While the 14:30–15:30 LST spectrum still resembles the daily average, the mean 5:30–6:30 LST
spectrum for the time of maximum precipitation comprises significantly more droplets, especially at greater
diameters.

4.3. Vertical profiles


The secondary rainfall maximum of the eastern slopes of the east cordillera in Ecuador seems to be located
between 3200 and 4000 m asl, as indicated by the vegetation and some sparsely distributed rain gauge stations
(Lauer, 1981). The situation in the free atmosphere over the ECSF building agrees quite well with this general
finding (Figure 6a). The rainfall maximum in the vertical profile is observed for the fifth level at 1347 m
above the ECSF station (=3207 m asl) with ∼5 mm·day−1 . The rainfall amount between this level and ground
decreases to 35% or to ∼3.2 mm·day−1 . A slight decrease is also obvious for the altitudinal levels above
3200 m asl. The comparison of the temporal maxima at 6:00 and 15:00 LST reveals interesting differences
in the vertical stratification of rainfall. The shape of the vertical profile during the afternoon maximum is
quite similar to the daily mean as is shown in Figure 6(a), but it shows a slightly higher decrease in rainfall
of 39% from the maximum zone at ∼3200 m asl to the ECSF station. This decrease, however, is clearly less
pronounced (16%) during the pre-dawn/dawn maximum where a significant reduction of rainfall amount is
observed at least in the lowest two levels.

1000
day
5:30 - 6:00 LST
100 14:30 - 15:00 LST
concentration [m-3]

10

0.1

0.01

0.001
0.13 0.21 0.29 0.38 0.49 0.6 0.74 0.91 1.12 1.4 1.83 2.79
radius [mm]

Figure 5. Average droplet spectra in the near-surface layer at ECSF

Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
DIURNAL PATTERNS OF RAINFALL IN A TROPICAL ANDEAN VALLEY 837

(a) (b)
24h 2693 5:30-6:30 14:30-
2693
LST 15:00 LST
2424 2424

2155 2155

[m above ECSF station] 1885 1885

1616 1616

1347 1347

1077 1077

∆R =12.2 mm
∆R = 5.2 mm
∆R =283 mm

= -16%

= -39%
= -35%
808 808

539 539

269 269

0 200 400 600 800 1000 1200 60 50 40 30 20 10 0 10 20 30 40 50 60


ECSF station [mm] [mm] [mm]
1860 m asl

Figure 6. (a) Rainfall for the whole data set (175 days) and (b) separated for dawn (6:00–630 LST) and afternoon (15:00–15:30 LST)
maximum in the free atmosphere over the ECSF station (m asl = 1860 m + y-axis)

Cifelli et al. (2000) concluded on the basis of 400 vertical profiles of tropical maritime rainfall that the
shape of the vertical reflectivity or rainfall profile is a good indication for the type of precipitation. For pure
stratiform rainfall, they found a minimum vertical variation due to the equilibrium between depletion (e.g.
evaporation) and growth (e.g. coalescence), which is not similar to the ECSF profiles. The daily and afternoon
profiles of Figure 6(a) resemble more situations that are characterised by mixed precipitation (stratiform with
embedded shallow convection or transition from convective to stratiform rain). However, the more complex
structured profile at 6:00 LST also shows shape features that were observed during convective situations by
Cifelli et al. (2000).
Evidence of mixed precipitation is also given by the shape of droplet concentrations in different altitudinal
levels (Figure 7).
The curves of the mean spectra at the height level of maximum precipitation (1347 m above ECSF) show
a distinct depression (especially 6:00 LST) at ∼0.5-mm diameter. Rao et al. (2001) found that this shape is
characteristic of transition periods between convection and uniform stratiform rain. A slight decrease of mean
droplet diameter from the maximum altitudinal belt (1347 m above ECSF, mean Ø0.54 mm) to the valley
bottom (mean Ø0.50 mm) and the greater similarity of the shape of droplet concentration at both levels point
to evaporation losses of droplets on their way to the valley bottom at noon. This is different for the dawn
maximum where the mean droplet diameter is nearly equal (∼0.57 mm) at both levels, especially for greater
droplets. Possible reasons could be reduced evaporation losses due to lower temperatures, a greater vertical
extension of the precipitating cloud downhill or the seeding of a nocturnal low stratus in the valley by a
higher precipitating cloud.
Because the pre-dawn/dawn maximum seems to be especially partly characterised by convective dynamics,
two case studies shall examine the origin of rainfall formation.

5. CASE STUDIES

5.1. Rainfall dynamics on 6 February 2003


An interesting period of pulsing nocturnal and dawn precipitation was found on 6 February 2003 (Figure 8).
Rain started at midnight and lasted throughout the morning with a quasi-periodic change of intensities. Peak
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
838 J. BENDIX, R. ROLLENBECK AND C. REUDENBACH

1000
539
100 1347

concentration [m-3]
10

0.1
5:30 - 6:00 LST
0.01
0.13 0.21 0.29 0.38 0.49 0.6 0.74 0.91 1.12 1.4 1.83 2.79
radius [mm]

1000
539
100 1347
concentration [m-3]

10

0.1

0.01
14:30 - 15:00 LST
0.001
0.13 0.21 0.29 0.38 0.49 0.6 0.74 0.91 1.12 1.4 1.83 2.79
radius [mm]

Figure 7. Mean droplet spectra for the maximum rainfall level and the near ground level at 6:00 and 15:00 LST

rain rates <10 mm·h−1 were measured between 4:20 and 4:50 LST at 3200 m asl, indicating convective
activity after a longer period of mostly stratiform precipitation. The oscillation of rain rate points to returning
showers that are based on embedded cellular convection. It is obvious that precipitation is formed at
higher levels because rainfall continuously decreases during its fall to the valley bottom by evaporation
losses. Weak stratiform rain continued until dawn, but after 6:36 LST a slight intensification of rain rates
occurred. Around noon (not shown here) strong rainfall was measured with rain rates >30 mm·h−1 (11:42
LST).
The entire dynamics points to larger-scale atmospheric instability. This is confirmed by inspect-
ing satellite images (Figure 9). A Mesoscale Convective Complex developed over the Peruvian Ama-
zon close to the southeastern borderline of Ecuador at 4:15 LST. This indicates atmospheric instabil-
ity, which has an effect on the eastern slopes of the east cordillera where initial cellular convection
can be detected in the area of the ECSF which is embedded in an extended stratiform cap cloud
layer covering the Amazonian slopes of the east cordillera. The 4:45 LST image shows that these ini-
tial cells are merged by forming a Mesoscale Convective System/Complex (MCS/MCC; for an expla-
nation of MCS/MCC refer to Maddox, 1980; Houze, 1993, 2004) over southeastern Ecuador, which
is the origin of stronger rain rates at ECSF that are located at the northwestern boundary of this
system.
The MCS/MCC collapsed during the following hour, leaving instability in the region of the ECSF. Locally
embedded cells were still discovered in the dawn/postdawn images, which were responsible for the new
increase in rain rate between 6:36 and 6:48 LST at ECSF.
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
DIURNAL PATTERNS OF RAINFALL IN A TROPICAL ANDEAN VALLEY 839

12
269 m above ECSF
10 808
8 1347

[mm h-1] 6
4
2
0
4:06 4:12 4:18 4:24 4:30 4:36 4:42
LST

12

10 269 m above ECSF


808
8 1347
[mm h-1]

0
6:06 6:12 6:18 6:24 6:30 6:36 6:42 6:48
LST

Figure 8. Vertical profiles of rain rate for early morning and dawn precipitation events at ECSF, 6 February 2003

4:15 LST 4:45 LST 6:15 LST

ECSF

Figure 9. GOES IR images from 6 February 2003. Dark grey shades represent warm, bright-shaded cold clouds

5.2. Rainfall dynamics on 22 July 2002


A strong pre-dawn precipitation event was observed on 22 July 2002 which lasted from 5:30 to 6:00 LST
(Figure 10). Light rainfall with rain rates of up to 2 mm·h−1 were observed during the morning hours until
9:00 LST. After a rain-free noon period, a strong afternoon rain event started at 17:30 LST, reaching peak
values of >20 mm·h−1 . The different stratification of rain rates is of interest. While the afternoon maximum
shows decreasing values with decreasing altitude due to evaporation losses, the maximum rain rate in the pre-
dawn event is observed for the third altitudinal level at 2600 m asl. Rain rate decreases above and below this
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
840 J. BENDIX, R. ROLLENBECK AND C. REUDENBACH

20
269
808
15 1347
[mm h-1]
10

0
5:36 5:42 5:48 5:54 6:00 6:06
LST

20 269
808
1347
15
[mm h-1]

10

0
17:30 17:36 17:42 17:48 17:54 18:00
LST

Figure 10. Vertical profiles of rain rate for pre-dawn and afternoon precipitation maxima at ECSF, 22 July 2002

terrain height. A weather cam image for this period (not shown here) reveals a low cloud base approximately
at the altitude of the third altitudinal level. It is obvious that droplets still grow (e.g. because of coalescence)
until the third level, probably also because of the seeding of the typical pre-dawn low stratus in the valley
by higher cloud.
Evidence is given by comparing the droplet concentrations for the peak times of rainfall (Figure 11). The
pre-dawn maximum is characterised by relatively small droplets with a maximum radius of 1.4 mm, whereas
the maximum droplet size during the afternoon peak is 1.8 mm. The 5:45 LST shape at 3200 m asl is
significantly changed at the third level where a secondary peak of droplets with a radius of 0.6 mm is formed
by the growth of smaller droplets while the concentration of droplets >1-mm radius slightly decreases. By
passing the third and second height level, the secondary peak in droplet concentration is clearly reduced by
growth, which leads to a greater amount of droplets >0.74-mm radius in the near-surface layer at the ECSF.
The situation is completely different for the afternoon peak where droplet concentration is at its maximum at
3200 m asl. Droplet amount decreases during the fall of the droplets by evaporation losses over all diameter
classes. However, a certain amount of small droplets is aggregated to greater droplets on their way to the
near-surface layer, which contributes to a strong decrease of small droplets.
NOAA-AVHRR cloud top heights can help explain the formation of the pre-dawn and afternoon rainfall
maxima (Figure 12). A mesoscale instability zone (∼150 km) in the southern part of Ecuador seems to be
responsible for the formation of the pre-dawn rain event. This instability is still seen in the 7:19 LST overpass
reaching from the Peruvian Amazon to the study area where the ECSF is again located at the northwestern
edge of a zone with cloud top heights >8 km, clearly indicating strong convection. This system decays after
sunrise, but a wide band of lower stratiform cap clouds is formed along the eastern cordillera. Small patches
of embedded shallow convection (cloud top height >6 km) can still be detected close to the ECSF area (9:51
LST), which are responsible for the weak precipitation during the morning hours.
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
DIURNAL PATTERNS OF RAINFALL IN A TROPICAL ANDEAN VALLEY 841

10000
269
1000 808

concentration [m-3]
1347
100

10

1 22nd July 2002


5:45 LST
0.1
0.13 0.21 0.29 0.38 0.49 0.6 0.74 0.91 1.12 1.4 1.83 2.79
radius [mm]

100000
269
10000
concentration [m-3]

808
1000 1347

100

10

1 22nd July 2002


17:42 LST
0.1
0.13 0.21 0.29 0.38 0.49 0.6 0.74 0.91 1.12 1.4 1.83 2.79
radius [mm]

Figure 11. Droplet concentration for the (a) pre-dawn and (b) afternoon event at several altitudes

NOAA - AVHRR cloud top height [m]


0 0
1000 1000
2000
2000
3000
ECSF 4000 3000
5000 4000
6000 5000
7000
6000
8000

7:19 LST 9:51 LST 14:30 LST

Figure 12. Cloud top heights (m asl) in southern Ecuador on 22 July 2002 as derived from NOAA-AVHRR data. The radar circle of
the local area weather radar (LAWR) has a diameter of 120 km

After noon, the instability collapses and the stratiform cloud layer is replaced by scattered clouds that
indicate the beginning of thermal convection (image 14:30 LST).
Scans of the LAWR can help clarify rainfall dynamics (Figure 13). Extended rainfields were developed
during the pre-dawn maximum with highest intensities southeast of the Rio San Francisco valley. These
spatial patterns of high rainfall are associated with barrage effects of the southeasterly winds at the fore-lying
but lower cordillera chains (e.g. C. de Paradones, C. de Naguipa). The lower parts of the area (basin of
Zamora) with altitudes of <1900 m asl remain free of rain. The ECSF station is already located in the lee
of the cordillera chains with high rain rates (e.g. C. de Consuelo) and receives its rainfall from the spatial
extension of the windward convective cells that overflow ridges with the easterly–southeasterly winds. The
spatial rainfall patterns during the afternoon period are less homogeneous and structured on a smaller scale.
The rain area is smaller and less orientated towards the windward side of the main cordillera chains. Various
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
842 J. BENDIX, R. ROLLENBECK AND C. REUDENBACH

ECSF

Loja

5:45 LST 5:50 LST

rainfall
-

17:40 LST 17:45 LST +

Figure 13. Local area weather radar (LAWR) observations on 22 July 2002 (raw counts). Bright grey levels indicate low dBZ values/rain
rates and dark shades indicate high dBZ values/rain rates

small convective cells can be detected, which point to the occurrence of local thermal convection on the
bordering slopes of the valleys. However, the ECSF itself does not lie in the centre of such cells but receives
its rainfall from the edge of these local rain cells.

6. DISCUSSION AND CONCLUSIONS

The study has shown that rainfall in the valley of the Rio San Francisco is characterised by a primary
pre-dawn/dawn and a secondary afternoon maximum. Stronger pre-dawn precipitation is related to mesoscale
instability, whereas afternoon rainfall also shows the influence of local thermal convection. However, it should
be stressed that a great part of rainfall at the ECSF station is of stratiform character but the main contribution
to rainfall amount seems to be provided by stronger events. The interesting question is the source of pre-
dawn/dawn instability during these events, which seems to be caused by larger-scale atmospheric dynamics.
Rickenbach (2004) found a secondary maximum of rainfall in southwestern Amazonia after midnight,
which was explained by large-scale squall lines from the coast, moving inland, which locally modulate the
diurnal variation of cloud formation and rain. Siqueira and Machado (2004) could show that subtropical
frontal systems from the southeast of South America can also influence convection at low latitudes, mostly in
austral summer. Garreaud and Wallace (1997) investigated the diurnal march of convective clouds along the
subtropical eastern Andean slopes (10–22 ° S) by means of the relatively low-resoluting ISCCP B3 data set
(30 km, 3 h). They found the well-known peaks of precipitation over land in the afternoon (∼13:00–17:00
LST Ecuador), which was mostly due to orographic uplift. Over the lowlands, this maximum gradually decays
during the night partly because of long-lived anvil debris. However, a secondary weak pre-dawn maximum
occurs at ∼6:00 LST in altitudes of about 2000 m. The authors point out that this maximum is presumably
related to enhanced low-level convergence during night time as a result of the nocturnal circulation between
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
DIURNAL PATTERNS OF RAINFALL IN A TROPICAL ANDEAN VALLEY 843

the Andes and the Amazon region. Mapes et al. (2003a) confirm the late afternoon maximum on the basis
of GOES-GPI data (spatial resolution 0.1° , temporal 3 h) but found also mesoscale specifics (e.g. nocturnal
precipitation in the valley of Rio Magdalena). A modelling study with the fifth-generation Pennsylvania State
University NCAR Mesoscale Model (MM5) suggests that gravity waves radiating eastwards from a stable-
stratified Andean mixed layer might also promote nocturnal precipitation (Mapes et al., 2003b). Overall, the
authors found that rainfall is frequently related to concavities in coastline or terrain lines. This was also proven
for nocturnal convection and precipitation in the Gulf of Guayaquil during El Niño situations (Bendix, 2000).
A diurnal analysis that is mainly based on TRMM-PR data reveals comparable results for the eastern Andean
slopes (de Angelis et al., 2004a). The authors found a nocturnal rainband (the so-called Andes-occurring
system, AOS) that forms between 1:00 and 3:00 LST on the eastern slopes of the Andes (including Ecuador)
and tends to travel towards the Amazon until 10:00 LST. They assumed that precipitation might be associated
with the convergence of cold katabatic flow from the Andes and the Amazon warm air pool, where katabatic
flows induce instability by acting as a cold front. Terra (2004) also stressed the importance of low-level
orographic gravity waves (cold air drainage flow) for cloud and rainfall formation.
However, looking at local details for the eastern Andean slopes of Ecuador during El Niño and La Niña
events on a day-to-day basis, a closed rainband structure cannot always be confirmed (Bendix et al., 2003a).
The analysis of rainfields that is retrieved from GOES data by means of the Enhanced Convective Stratiform
Technique (ECST, Reudenbach et al., 2001, Reudenbach and Bendix, 2003) and TRMM-PR data showed
that two distinct areas of nocturnal rainfall exist (Bendix et al., 2003b). One centre is located in northern
Ecuador on the border to Columbia and the other centre covers the eastern slopes of southern Ecuador on the
border to Peru. The eastern slopes of the central Ecuadorian Andes receive only small amounts of precipitation
(1:00–7:00 LST) during cold phases or are free of rain during warm events. Generally, nocturnal precipitation
especially in the south Ecuadorian centre is significantly stronger during cold events.
It is striking that both centres of nocturnal rainfall extend up to ∼3000 m asl and are connected with a
concavity in the Andean chain and are related to greater river valleys (as e.g. Rio Marañon in the southern
part), which can act as effective drainage systems for cold air pools from the highlands. Both initial nocturnal
convective systems described in the current study (Figures 9 and 12) are formed in the same region as the
instabilities observed especially during cold events, when cold air drainage flow should be increased. Visual
inspection of AVHRR images from other situations with pre-dawn rainfall at ECSF reveals a similar mesoscale
situation. The most probable mechanism is in accordance with the assumptions of de Angelis et al. (2004a,b).
A confluence of cold air drainage flow from the eastern Andean valleys (e.g. the Rio Marañon) produces
extended atmospheric instability and deep convection after midnight in the Amazonian forelands southeast
of the Ecuadorian border. The instability is then shifted westwards by the mid- and upper-tropospheric
easterlies, thus extending convective activity to the eastern Andean slopes, which is additionally strengthened
by barrage effects and forced convection. This leads to deep or embedded shallow convective rainfall, also
at the windward sides of the individual cordillera chains, which peaks at pre-dawn/dawn and is frequently
extended to the first hours after sunrise with decreased rain rates. The upper San Francisco valley and the
ECSF station can profit from rainfields overflowing from the ridge of the sheltering cordillera.
The situation for strong convective afternoon precipitation is rather characterised by isolated thermal cells,
which produce rainfall on the bordering slopes of the San Francisco valley. The ECSF station profits from
this rainfall at the edge of the cells (also stratiform region) because the valley is relatively narrow.
Unfortunately, no radiosonde data are available from the valley and the relevant Amazon foreland area,
which could help improve understanding of the release of instability, especially during the precipitation peak
around sunrise. However, data of meteorological stations for the case study of 21/22 July 2002 (Figure 14)
points to an influence of the nocturnal cold air drainage flow and the local heating during daylight on rainfall
formation as described in this paper. Wind direction in the valley (ECSF met station) reveals the typical
diurnal course from an up-valley breeze during daylight (northeasterly flow) to a nocturnal katabatic flow
system (southwesterly direction). While the local heating of the slopes over the day causes temperatures
>24 ° C, the nocturnal cold air drainage flow contributes to significantly lower values <12 ° C. Obviously,
the clear diurnal cycle is not established at pass altitudes (station El Tiro, for location refer to Figure 1).
The prevailing wind direction is southeast all over the period. The radar movie of Figure 13 reveals that the
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
844 J. BENDIX, R. ROLLENBECK AND C. REUDENBACH

315 ECSF
270 El Tiro
225
180
135
90
45
0
25 ECSF El Tiro
20
15
10
5
0

12:00 18:00 0:00 6:00 12:00 18:00


LT

Figure 14. Wind direction, air temperature (2 m) and rainfall for two stations in the study area (ECSF met station at 1960 m asl, El
Tiro met station at 2825 m asl) for the 21/22 July 2002

pre-dawn rain cells track southeast (also the direction to the Mesoscale Convective System in Figure 12) to
northwest. It is striking that the wind direction in the valley turns for a short time to synoptic conditions (SE)
during the precipitation peak before 6:00 LT, which points to the mesoscale influence in combination with
the katabatic drainage flow system.
Poveda et al. (2005) have observed nocturnal rainfall maxima at the eastern flanks of the Andes and
found that they are mainly associated with the life cycle of Mesoscale Convective Systems by an upscale
development of storm size through the night, supported by diurnally driven breezes and gravity waves.
Evidence for mesoscale dynamics is given by a lack of seasonality in the diurnal pattern as it should be
expected, e.g. as a result of the passage of the ITCZ. Houze (2004) stressed that LLJs at slopes of high
mountain systems can help maintain low-level buoyancy and lead to an intensification of MCS/MCC, which
are e.g. initiated by low-level cold pool dynamics. LLJs east of the Andes are frequent in all seasons and
are well known to be responsible for transporting large quantities of water into central South America (e.g.
Bendix and Lauer, 1992; Liebmann et al., 2004). Finally, the study of Velasco and Fritsch (1987) has proven
that MCCs frequently occur in the Amazon foreland southeast of Loja as likewise illustrated by the presented
case studies. The secondary afternoon/early evening peak of rainfall is most probably triggered by local
heating of the valley slopes (Figure 14) over the day, releasing instability. Poveda et al. (2005) argued that
convection is significantly enhanced by the entrance of low-level, moisture-laden winds from the Amazon
which ascend because of orographic lifting and reach condensation level in the afternoon or around sunset. de
la Torre et al. (2004) stressed in this context that anabatic winds are not sufficient enough, but specific moist
enthalpy convergence near the ground and instability conditions are necessary. For an Andean valley, they
could show that specific moist enthalpy convergence is three times higher on days with deep convection and
heavy precipitation. Figure 14 illustrates that Amazon air masses can enter the San Francisco valley because
of the up-valley breeze from 7:00 to 18:00 LT. Specific humidity at the ECSF meteorological station (3:00
LT: 9.9 g·kg−1 , 10:00 LT: 12.3 g·kg−1 ) for 22 July 2002 points to an enhanced moisture supply from the
Amazon by the up-valley breeze system.
Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
DIURNAL PATTERNS OF RAINFALL IN A TROPICAL ANDEAN VALLEY 845

In summary, future analysis of a comprehensive LAWR data set by means of image tracking and mesoscale
modelling should provide deeper insight into rain formation processes in the study area around the ECSF
station.

ACKNOWLEDGEMENTS

The current work was performed within the framework of the DFG Research Group FOR 402 ‘Functionality
in a tropical mountain forest: diversity, dynamic processes and utilisation. Potentials under ecosystem
perspectives’ and was generously funded by the German Research Council DFG (BE 1780/5-1, 5-2). It
is also part of the LCRS activities within the IPWG (WMO Int. Precipitation Working Group). The authors
thank Jane Kölzer and Melanie Schulz for operating the MRR in Ecuador and NCI (Loja) for logistical
support. We are indebted to G. Peters and B. Fischer (Meteorol. Inst., Univ. of Hamburg) for providing a
tool for revised MRR data calibration. Thanks also go to Ricardo Ponce and Enrique Palacios (INAMHI)
for operating the HRPT-receiving station and logistical support. We thank M. Richter and P. Emck (Inst. of
Geogr., Univ Erlangen-Nürnberg) for providing the data of the meteorological stations.

REFERENCES
Ataroff M 1998. Importance of Cloud-Water in Venezuelan Andean Cloud Forest Water Dynamics. Proceedings of the 1st International
Conference on Fog and Fog Collection, Vancouver, Canada, 19–24 July 1998, 25–28 (Vancouver: Conference on Fog and Fog
Collection).
Barry RG. 1992. Mountain climatology and past and potential future climatic changes in mountain regions: A review. Mountain Research
and Development 12: 71–86.
Barry RG. 2001. Mountain Weather and Climate. Routledge: London.
Beck E, Müller-Hohenstein K. 2001. Analysis of undisturbed and disturbed tropical mountain forest ecosystems in Southern Ecuador.
Die Erde 132: 1–8.
Bendix J. 2000. Precipitation dynamics in Ecuador and Northern Peru during the 1991/92 El Niño – A remote sensing perspective.
International Journal of Remote Sensing 21: 533–548.
Bendix J, Lauer W. 1992. The rainy seasons in Ecuador and their climatic interpretation (in German). Erdkunde 46: 118–134.
Bendix J, Rafiqpoor MD. 2001. Thermal conditions of soils in the Páramo of Papallacta (Ecuador) at the upper tree line. Erdkunde 55:
257–276.
Bendix J, Reudenbach C, Rollenbeck R 2003a. The Marburg Satellite Station. Proceedings 2002 Met. Sat. Users’ Conference. Dublin,
2–6 September 2002, EUMETSAT, 139–146.
Bendix J, Gämmerler S, Reudenbach C, Bendix A. 2003b. A case study on rainfall dynamics during El Niño/La Niña 1997/99 in
Ecuador and surrounding areas as inferred from GOES-8 and TRMM-PR observations. Erdkunde 57: 81–93.
Bendix J, Rollenbeck R, Palacios WE. 2004. Cloud detection in the Tropics – a suitable tool for climate-ecological studies in the high
mountains of Ecuador. International Journal of Remote Sensing 25: 4521–4540.
Chu PS, Hastenrath S. 1982. Atlas of Upper-Air Circulation Over Tropical South America. Department of Meteorology: Madison.
University of Madison-Wisconsin.
Cifelli R, Williams CR, Rajopadhyaya DK, Avery SK, Gage KS, May PT. 2000. Drop-size distribution characteristics in tropical
Mesoscale Convective Systems. Journal of Applied Meteorology 39: 760–777.
de la Torre A, Daniel V, Tailleux R, Teitelbaum H. 2004. A deep convection event above the Tunuyán valley near the Andes mountains.
Monthly Weather Review 132: 2259–2267.
de Angelis CF, McGregor GR, Kidd C. 2004a. A 3 year climatology of rainfall characteristics over tropical and subtropical South
America based on tropical rainfall measuring mission precipitation radar data. International Journal of Climatology 24: 385–399.
de Angelis CF, McGregor GR, Kidd C. 2004b. Diurnal cycle of rainfall over the Brazilian Amazon. Climate Research 26: 139–149.
Diem M, Strantz R. 1971. Types of raindrop spectra II: Dependence on rain intensity (in German). Meteorologische Rundschau 24:
23–26.
Garreaud RD, Wallace JM. 1997. The diurnal march of convective cloudiness over the Americas. Monthly Weather Review 125:
3157–3171.
Groisman PY, Legates DR. 1994. The accuracy of United States precipitation data. Bulletin of the American Meteorological Society 75:
215–227.
Gunn R, Kinzer GD. 1949. The terminal velocity of fall for water droplets in stagnant air. Journal of Meteorology 6: 243–248.
Houze RA Jr. 1993. Cloud Dynamics. Academic Press. San Diego, San Francisco, New York, Boston, London, Sydney, Tokyo.
Houze RA Jr. 2004. Mesoscale convective systems. Reviews of Geophysics 42: RG4003, DOI:10.1029/2004RG000150.
Jensen NE. 2002. X-Band local area weather radar – preliminary calibration results. Water Science and Technology 45: 135–138.
Kirshbaum DJ, Durran DR. 2004. Factors governing cellular convection in orographic precipitation. Journal of the Atmospheric Sciences
61: 682–698.
Klugmann D, Heinsohn K, Kirtzel H-J. 1996. A low cost 24 GHz FM-CW-Doppler radar rain profiler. Contributions to Atmospheric
Physics 69: 247–253.
Lauer W. 1981. Ecoclimatological conditions of the Paramo belt in the tropical high mountains. Mountain Research and Development
1: 209–221.

Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)
846 J. BENDIX, R. ROLLENBECK AND C. REUDENBACH

Liebmann B, Kliadis GN, Vera CS, Saulo AC, Carvalho LMV. 2004. Subseasonal variations of rainfall in South America in the vicinity
of the low-level jet east of the Andes and comparison to those in the South Atlantic Convergence Zone. Journal of Climatology 17:
3829–3842.
Lin YL, Chiao S, Wang TA, Kaplan ML, Weglarz RP. 2001. Some common ingredients for heavy orographic rainfall. Weather and
Forecasting 16: 633–660.
Lin YL. 2003. The dynamics of orographic precipitation, Preprints of The Harold D. Orville Symposium, 26 April 2003. Institute for
Atmospheric Sciences: Rapid City, SD, 68–89.
Löffler-Mang M, Kunz M, Schmid W. 1999. On the performance of a low-cost K-band Doppler radar for quantitative rain measurements.
Journal of Atmospheric and Oceanic Technology 16: 379–387.
Maddox RA. 1980. Mesoscale convective complexes. Bulletin of the American Meteorological Society 61: 1374–1387.
Mapes BE, Warner TT, Xu N, Negri AJ. 2003a. Diurnal pattern of rainfall in northwestern South America. Part I: Observation and
context. Monthly Weather Review 131: 799–812.
Mapes BE, Warner TT, Xu N, Negri AJ. 2003b. Diurnal pattern of rainfall in northwestern South America. Part III: Diurnal gravity
waves and nocturnal convection offshore. Monthly Weather Review 131: 830–844.
Matrosov SY. 2005. Attenuation-based estimates of rainfall rates aloft with vertically pointing Ka -band radars. Journal of Atmospheric
and Oceanic Technology 22: 43–54.
Peters G, Fischer B, Andersson T. 2002. Rain observation with a vertically looking Micro Rain Radar (MRR). Boreal Environmental
Research 7: 353–362.
Poveda G, Mesa OJ, Salazar LF, Arias PA, Moreno HA, Vieira SC, Agudelo PA, Toro VG, Alvarez JF. 2005. The diurnal cycle of
precipitation in the tropical Andes of Columbia. Monthly Weather Review 133: 228–240.
Rao TN, Rao DN, Mohan K. 2001. Classification of tropical precipitating systems and associated Z-R relationships. Journal of
Geophysical Research 106: 17699–17711.
Reudenbach C, Heinemann G, Heuel E, Bendix J, Winiger M. 2001. Investigation of summertime convective rainfall in Western Europe
based on a synergy of remote sensing data and numerical models. Meteorology and Atmospheric Physics 76: 23–41.
Reudenbach C, Bendix J 2003. Satellite Based Rainfall Retrieval with Meteosat, GOES and MSG in the Mid-Latitudes and the Tropics.
Proceedings of the 1st Workshop of IPWG, 23–27 Sept. 2002, Madrid. EUM P34, EUMETSAT, 135–144.
Richter M. 2003. Using plant functional types and soil temperatures for eco-climatic interpretation in southern Ecuador. Erdkunde 57:
161–181.
Rickenbach TM. 2004. Nocturnal cloud systems and the diurnal variation of clouds and rainfall in southwestern Amazonia. Monthly
Weather Review 132: 1201–1219.
Siqueira JR, Machado LAT. 2004. Influence of the frontal systems on the day-to-day convection variability over South America. Journal
of Climate 17: 1754–1766.
Sklenář P, Lægaard S. 2003. Rain-shadow in the high Andes of Ecuador evidenced by Páramo vegetation. Arctic Antarctic and Alpine
Research 35: 8–17.
Strantz R. 1971. Types of raindrop spectra (in German). Meteorologische Rundschau 24: 19–23.
Terra R. 2004. PBL stratiform cloud inhomogeneities thermally induced by the orography: a parameterization for climate models.
Journal of the Atmospheric Sciences 61: 644–663.
Trojer H. 1959. Fundamentos para una zonificación meteorológico del trópico y especialmente de Colombia. Boletı́n de Informaciones
de Centro Nacional. de Investigationes Café 10: 289–373.
Troll C. 1968. The Cordilleras of the Tropical Americas. Aspects of climatic, phytogeographical and agrarian ecology. Colloquium
Geographicum 9: 15–56.
Tokay A, Short DA. 1996. Evidence from tropical raindrop spectra of the origin of rain from stratiform versus convective clouds.
Journal of Applied Meteorology 35: 359–371.
Tokay A, Short DA, Williams CR, Ecklund WL, Gage KS. 1999. Tropical rainfall associated with convective and stratiform clouds:
Intercomparison of disdrometer and profiler measurements. Journal of Applied Meteorology 38: 302–320.
Vali G, Kelly RD, Pazmany A, McIntosh RE. 1995. Airborne radar and in-situ observations of a shallow stratus with drizzle. Atmospheric
Research 38: 361–380.
Velasco I, Fritsch M. 1987. Mesoscale convective complexes in the Americas. Journal of Geophysical Research 92: 9591–9613.
Vuille M, Bradley RS, Keimig F. 2000. Climate variability in the Andes of Ecuador and its relation to tropical Pacific and Atlantic sea
surface temperature anomalies. Journal of Climate 13: 2520–2535.
Weischet W. 1969. Climatological principles of the vertical distribution of rainfall in tropical mountains (in German). Die Erde 100:
287–306.
Zimmerschied W. 1958. Preliminary announcements on the rainfall situation of Ecuador (in German). Meteorologische Rundschau 11:
156–162.

Copyright  2005 Royal Meteorological Society Int. J. Climatol. 26: 829–846 (2006)

You might also like