Using a gridded global data set to characterize regional

advertisement
1
Using a gridded global data set to characterize regional hydroclimate in central Chile
2
3
E.M.C. Demaria1, E. Maurer2*, J. Sheffield3, E. Bustos1, D. Poblete1, S. Vicuña1, F. Meza1
4
5
1
6
Chile
7
2
Civil Engineering Department, Santa Clara University, Santa Clara, CA, USA
8
3
Department of Civil and Environmental Engineering, Princeton University, Princeton, NJ, USA
Centro Interdisciplinario de Cambio Global, Pontificia Universidad Católica de Chile, Santiago,
9
10
*Corresponding author, emaurer@engr.scu.edu, 408-554-2178.
11
12
Proposed submission to J. Hydrometeorology
13
14
1
15
Abstract
16
Central Chile is facing dramatic projections of climate change, with a consensus for declining
17
precipitation, negatively affecting hydropower generation and irrigated agriculture. Rising from
18
sea level to 6,000 meters within a distance of 200 kilometers, precipitation characterization is
19
difficult due to a lack of long-term observations, especially at higher elevations. For
20
understanding current mean and extreme conditions and recent hydroclimatological change, as
21
well as to provide a baseline for downscaling climate model projections, a temporally and
22
spatially complete data set of daily meteorology is essential. We use a gridded global daily
23
meteorological data set at 0.25 degree resolution for the period 1948-2008, and adjust it using
24
monthly precipitation observations interpolated to the same grid using a cokriging method with
25
elevation as covariate. For validation, we compare daily statistics of the adjusted gridded
26
precipitation to station observations. For further validation we drive a hydrology model with the
27
gridded 0.25-degree meteorology and compare stream flow statistics with observed flow. We
28
validate the high elevation precipitation by comparing the simulated snow extent to MODIS
29
images. Results show that the daily meteorology with the adjusted precipitation can accurately
30
capture the statistical properties of extreme events as well as the sequence of wet and dry events,
31
with hydrological model results displaying reasonable agreement for observed flow statistics and
32
snow extent. This demonstrates the successful use of a global gridded data product in a relatively
33
data-sparse region to capture hydroclimatological characteristics and extremes.
34
35
36
2
37
1. Introduction
38
Whether exploring teleconnections for enhancing flood and drought predictability or assessing
39
the potential impacts of climate change on water resources, understanding the response of the
40
land surface hydrology to perturbations in climate is essential. This has inspired the development
41
and assessment of many large scale hydrologic models for simulating land-atmosphere
42
interactions over regional and global scales [e.g., Lawford et al., 2004; Milly and Shmakin, 2002;
43
Nijssen et al., 2001a; Sheffield and Wood, 2007].
44
45
A prerequisite to regional hydroclimatological analyses is a comprehensive, multi-decadal,
46
spatially and temporally complete data set of observed meteorology, whether for historic
47
simulations or as a baseline for downscaling future climate projections. In response to this need,
48
data sets of daily gridded meteorological observations have been generated, both over
49
continental regions [e.g.,Cosgrove et al., 2003; Maurer et al., 2002] and globally [Adam and
50
Lettenmaier, 2003; Sheffield et al., 2006]. These have benefited from work at coarser time scales
51
[Chen et al., 2002; Daly et al., 1994; Mitchell and Jones, 2005; New et al., 2000; Willmott and
52
Matsuura, 2001], with many products combining multiple sources, such as station observations,
53
remotely sensed images, and model reanalyses.
54
55
While these large-scale gridded products provide opportunities for hydrological simulations for
56
land areas around the globe, they are inevitably limited in their accuracy where the underlying
57
density of available station observations is low, the station locations are inadequate to represent
58
complex topography, or where the gridded spatial resolution is too large for the region being
3
59
studied. Central Chile is an especially challenging environment for characterizing climate and
60
hydrology since the terrain exhibits dramatic elevation changes over short distances, and the
61
orographic effects produce high spatial heterogeneity in precipitation in particular. In general, the
62
observation station density in South America is inadequate for long-term hydroclimate
63
characterization [de Goncalves et al., 2006]. While some of South America is relatively well
64
represented by global observational datasets [Silva et al., 2007], regions west of the Andes are
65
much less so [Liebmann and Allured, 2005].
66
67
In this study, we utilize a new high-resolution global daily gridded dataset of temperature and
68
precipitation, adjust it with available local climatological information, and assess its utility for
69
representing river basin hydrology. Recognizing the value in simulating realistic extreme events,
70
we assess the new data product for its ability to produce reasonable daily streamflow statistics.
71
We evaluate the potential to reproduce climate and hydrology in a plausible manner, such that
72
historical statistics are reproduced.
73
74
The principal aim of this study is to produce a gridded representation of the climate and
75
hydrology of central Chile, and demonstrate a methodology for producing a reasonable set of
76
data products that can be used for future studies of regional hydrology or climate. Given these
77
regional results, we assess the potential to export the method to other relatively data-sparse
78
regions, where representative climatological average information is available but long-term daily
79
data are inadequate. The paper is organized as follows: Section 2 describes the study area. In
80
Section 3 we describe the data, the hydrological model, and the methodological approach.
4
81
Results of the adjusted data set validation and model simulations are discussed in Section 4.
82
Finally, the main conclusions of the study are presented in Section 5.
83
84
2. Region
85
The focus area of this study is the region in central Chile, encompassing the four major river
86
basins (from north to south, the Rapel, Mataquito, Maule, and Itata Rivers) between latitudes
87
35.25º S and 37.5º S (Figure 1). The climate is Mediterranean, with 80% of the precipitation
88
falling in the rainy season from May-August [Falvey and Garreaud, 2007]. The terrain is
89
dramatic, rising approximately 6000 meters within a horizontal distance of approximately 200
90
km, producing sharp gradients in climate [Falvey and Garreaud, 2009].
91
92
Mean precipitation is approximately 500 mm per year at the North end of the study domain, and
93
as much as 3000 mm per year in the high elevations at the Southern end of the domain. Although
94
climate information in the valley or mountain foothills is well represented by meteorological
95
stations it is evident from Figure 1 that the high elevation areas are under-represented by any of
96
the observation stations.
97
98
Our study region in Central Chile is especially important from a hydroclimatological standpoint,
99
as
it
contains
more
than
75%
of
the
Country’s
total
irrigated
agriculture
100
[www.censoagropecuario.cl/index2.html] and reservoir storage of any region in the country, and
101
provides water supply for some of Chile's largest cities. A changing climate is evident in recent
102
hydroclimate records [Rubio-Álvarez and McPhee, 2010], and future climate projections for the
5
103
region indicate the potential for very large impacts [Bradley et al., 2006]. The vulnerability of
104
Central Chile to projected climate change is high, with robust drying trends in General
105
Circulation Model (GCM) projections, and a high sensitivity to changing snow melt patterns
106
[Vicuna et al., 2011], who also discuss the challenges in characterizing climate in a Chilean
107
catchment with few precipitation observations, and none at high elevations.
108
109
3. Methods and data
110
3.1 Gridded data set development
111
We begin with a gridded global (land surface) forcing dataset of daily precipitation and
112
minimum and maximum temperatures at 0.25º spatial resolution (approximately 25 km),
113
prepared following Sheffield et al. [2006]. To summarize, the forcing dataset is based on the
114
NCEP–NCAR reanalysis [Kalnay et al., 1996] for 1948-2008, from which daily maximum and
115
minimum temperature and daily precipitation are obtained at approximately 2º spatial resolution.
116
Reanalysis temperatures are based on upper-air observations, though precipitation is a model
117
output and thus exhibits significant biases.
118
119
The reanalysis temperatures are interpolated to a 0.25º spatial resolution, using lapsing
120
temperatures of -6.5ºC/km based on the elevation difference between the large reanalysis spatial
121
scale and the elevation in each 0.25º grid cell. Precipitation is interpolated to 0.25º using a
122
product of the Tropical Rainfall Measuring Mission (TRMM) [Huffman et al., 2007] following
123
the methods outlined by Sheffield et al. [2006]. The daily statistics of precipitation (number of
124
rain days, wet/dry day transition probabilities) are corrected by resampling to match the
6
125
observationally-based monthly 0.5º rain days data from the Climate Research Unit [CRU,
126
Mitchell and Jones, 2005] and the daily statistics of the Global Precipitation Climatology Project
127
[Huffman et al., 2001] and the global forcing dataset of Nijssen et al. [2001a]. To ensure large-
128
scale correspondence between this data set and the CRU monthly data set, precipitation is scaled
129
so the monthly totals match the CRU monthly values at the CRU spatial scale. Maximum and
130
minimum temperatures are also scaled to match the CRU time series, using CRU monthly mean
131
temperature and diurnal temperature range.
132
133
While the incorporation of multiple sources of extensively reviewed data provides an invaluable
134
data product for global and continental scale analyses, as discussed by Mitchell and Jones [2005]
135
ultimately much of the local characterization is traceable to a common network of land surface
136
observations [Peterson et al., 1998], which is highly variable in station density for different
137
regions. For example, for the region of study shown in Figure 1, an average of 3-4 observation
138
stations are included in the CRU precipitation data product, and none are in high-elevation areas.
139
This results in a few low elevation meteorological stations in Chile on the western side of the
140
Andes, and the next observation station to the east is in a more arid area in Argentina. Thus, the
141
resulting precipitation fields in the gridded product for this region show a spatial gradient
142
opposite to that published by the Dirección General de Aguas [DGA, 1987]. Figure 2a shows the
143
spatial distribution of gridded global total annual precipitation that displays a notable decrease of
144
rainfall with elevation. Conversely the DGA precipitation map is able to capture the
145
climatological orographic enhancement of precipitation by the Andes (Figure 2b). The
146
precipitation lapse rates for the latitudinal bands -35.125º S and -36.125º S show a negative
7
147
gradient of precipitation with elevation in the global gridded data set whereas the DGA
148
precipitation shows a positive gradient for the period 1951-89 (Figures 2c and 2d, respectively).
149
150
Local data from the DGA of Chile, some monthly and some daily, were obtained to characterize
151
better the local climatology. While still biased toward low elevation areas, the stations (Figure 1)
152
do cover a wider range and include altitudes up to 2400 m. These stations were filtered to include
153
those that had at least 90% complete monthly records for a 25-year period, with the period
154
containing the most complete data coverage identified as 1983-2007. Since the climate of Central
155
Chile is also modulated by interannual variability linked to El Nino and the Pacific Decadal
156
Oscillation [Garreaud et al., 2009] a 25-year period was chosen to limit the influence of a single
157
predominant phase of low-frequency oscillations during the climatological period. The monthly
158
average precipitation for the selected DGA stations was interpolated onto the same 0.25º grid
159
using cokriging, with elevation being the covariate. This method of cokriging has been shown to
160
improve kriging interpolation to include orographic effects induced by complex terrain [Diodato
161
and Ceccarelli, 2005; Hevesi et al., 1992].
162
163
This process produced twelve monthly mean precipitation maps for the region. The same 1983-
164
2007 period was extracted from the daily gridded data set, and monthly average values were
165
calculated for each grid cell. Ratios (twelve, one for each month) of observed climatology
166
divided by the gridded data set average were then calculated for each grid cell. Daily values in
167
the gridded data set were adjusted to create a new set of daily precipitation data, Padj, which
168
matches the interpolated observations produced with cokriging, using a simple ratio:
8
Padj i, j,t   Pgrid i, j,t  
Pobs,mon i, j 
Pgrid,mon i, j 
(1)
169
where Pgrid is the original daily gridded 0.25º data at location (i,j), Pobs is the interpolated
 170
observed climatology, overbars indicate the 25-year mean, and the subscript “mon” indicates the
171
month from the climatology in which day t falls.
172
173
This same method was applied to a global dataset of daily meteorology in a data sparse region in
174
Central America, resulting in improved characterization of precipitation and land surface
175
hydrology [Maurer et al., 2009]. In addition, this new adjusted data set includes the full 1948-
176
2008 period, despite the fact that local observations are very sparse before 1980.
177
178
To validate the adjusted precipitation data set, we computed a set of statistical parameters widely
179
used to describe climate extremes [dos Santos et al., 2011; X. Zhang and Yang, 2004].
180
Additionally to evaluate the temporal characteristics of rainfall events we computed the
181
probability of occurrence of wet and dry days, and the transition probabilities between wet and
182
dry states [Wilks and Wilby, 1999]. Table 1 shows a description of the statistics used.
183
184
To evaluate if the adjusted precipitation data set was capturing the orographic gradient of
185
precipitation we compared model simulated Snow Water Equivalent (SWE) to the MODIS/Terra
186
Snow Cover data set, which is available at 0.05 degree resolution for 8-day periods starting from
187
the year 2000. MODIS snow cover data are based on a snow mapping algorithm that employs a
188
Normalized Difference Snow Index [Hall et al., 2006]. To estimate snow cover from the
189
meteorological data, the Variable Infiltration Capacity (VIC) model was employed (se model
9
190
description in Section 3.2). The accuracy of simulated snow cover relative to that of random
191
chance was measured with the Heidke Skill Score [HSS, Wilks, 2006]. A score equal to one
192
would indicate perfect agreement between VIC simulated snow cover and observations; a value
193
greater than zero indicates some predictive skill. Thus, the closer the HSS is to one, the less
194
likely it is that the agreement between observations and simulations has been obtained by
195
chance. The score is computed as:
196
HSS 
2ad  bc)
a  c c  d  a  bb  d
(2)
197
 198
where a and d are the numbers of hits (i.e., pairs of successful MODIS-VIC no snow and snow
199
estimates, respectively), b is the number of cases when VIC simulates snow but MODIS does not
200
measure it, and c is the number of events when snow is observed by MODIS but not simulated
201
by VIC.
202
203
3.2 Hydrologic Model Simulations
204
To assess the ability of the daily gridded meteorology developed in this study to capture daily
205
climate features across the watersheds, we simulate the hydrology of river basins in the region to
206
obtain streamflow and snow cover estimates. The hydrologic model used is the Variable
207
Infiltration Capacity (VIC) model [Cherkauer et al., 2003; Liang et al., 1994]. The VIC model is
208
a distributed, physically-based hydrologic model that balances both surface energy and water
209
budgets over a grid mesh. The VIC model uses a “mosaic” scheme that allows a statistical
210
representation of the sub-grid spatial variability in topography, infiltration and vegetation/land
10
211
cover, an important attribute when simulating hydrology in heterogeneous terrain. The resulting
212
runoff at each grid cell is routed through a defined river system using the algorithm developed by
213
Lohmann et al. [1996]. The VIC model has been successfully applied in many settings, from
214
global to river basin scale [e.g., Maurer et al., 2002; Nijssen et al., 2001b; Sheffield and Wood,
215
2007].
216
217
For this study, the model was run at a daily time step at a 0.25º resolution (approximately 630
218
km2 per grid cell for the study region). Elevation data for the basin routing were based on the 15-
219
arc-second Hydrosheds dataset [Lehner et al., 2006], derived from the Shuttle Radar Topography
220
Mission (SRTM) at 3 arc-second resolution. Land cover and soil hydraulic properties were based
221
on values from Sheffield and Wood [2007], though specified soil depths and VIC soil parameters
222
were modified during calibration. The river systems contributing to selected points were defined
223
at a 0.25º resolution, following the technique outlined by O’Donnell et al. [1999].
224
225
4. Results and Discussion
226
The adjusted data set was validated in several ways. First, daily statistics were compared
227
between the adjusted global daily data set and local observations, where available. Second,
228
hydrologic simulation outputs were compared to observations to investigate the plausibility of
229
using the new data set as an observational baseline for studying climate impacts on hydrology.
230
231
4.1 Gridded meteorological data development and assessment
11
232
The quality of daily gridded precipitation fields was improved using available monthly observed
233
precipitation. Rain gauge records from DGA were selected using two criteria: stations with
234
records of twenty-five years and with no more than 10% missing daily measurements. Based on
235
those two constraints the period 1983-2007 was identified as that with the largest number of
236
reporting stations. From the pool of 70 available stations, 40 stations met the two criteria (Figure
237
1). Except for the Itata river basin, which had two stations located at 1200 and 2400 meters
238
above sea level, most of the selected stations were located in the central part of the region at
239
elevations below 500 meters. Mean precipitation was computed for each month and for each
240
selected station, resulting in 12 mean values for the 25-year climatological period.
241
242
Cokriging was then applied to produce a set of 12 maps of climatological precipitation at 0.25º
243
spatial resolution. A scatter plot between observed and predicted average (1983-2007) monthly
244
precipitation for July, the middle of the rainy season, is shown in Figure 3. Cokriged monthly
245
totals match observations quite closely for the region with a bias equal to -0.8 % with respect to
246
the observed values and a relative RMSE of 0.50 %, which is expected since the stations were
247
included in the kriging process.
248
249
Figure 4 shows the adjusted gridded annual precipitation fields and the difference from the
250
original gridded observed data set for the period 1950-2006. It is evident that in the more humid
251
southern mountainous portion of the study area there has been a marked increase in precipitation
252
with the adjustment, incorporating the more detailed information embedded in the rain gauge
253
observations. Differences between original and adjusted gridded precipitation indicates the
12
254
existence of a band along the Andes where annual precipitation is greater in the adjusted
255
precipitation data set (Figure 4b).
256
257
To verify how the adjusted daily precipitation relates to observations, we compared daily rainfall
258
at selected 0.25º grid points with the day-by-day means of three rain gauge stations located in
259
approximately a 50 km diameter circle (Figure 5). Rain gauge stations were selected from the
260
pool of 40 stations used to perform the cokriging interpolation, hence they had a record of 25
261
years with not more than 10% missing values. Selected stations were located, when possible, not
262
more that 50% higher (maximum elevation difference was 150 meters except at Loc3 where it
263
was 500 meters) or lower elevations than that of the 0.25º grid cell. Four 0.25º grid points were
264
selected for the comparison. The locations of the four grid points are listed in Table 2. For these
265
four locations, we computed basic statistics, bias, RMSE and correlation coefficient for daily
266
observed (OBS) and daily adjusted gridded precipitation (ADJ) for Austral summer (DJF) and
267
Austral winter (JJA) for the period 1983-2007. Summary statistics are shown in Table 3. The
268
bias is defined as the sum of the differences between ADJ and OBS, and the RMSE is equal to
269
the root mean squared error between daily ADJ and OBS precipitation values.
270
271
Mean daily values are very close for the observed and adjusted datasets for both seasons, which
272
is expected given the adjustment process. The variability of daily precipitation within each
273
season, represented by the standard deviation, also compares relatively well, though the adjusted
274
gridded data show greater variability than the observations during the rainy winter season. A
275
high RMSE and low correlation values indicate that temporal sequencing differs between the two
276
data sets. This is not unexpected, since the daily precipitation in the original 0.25º gridded data
13
277
was derived from reanalysis, and as such it is a model output that does not incorporate station
278
observations [Kalnay et al., 1996], and is resampled on a daily basis to correct the daily statistics
279
of precipitation. Thus, while important characteristics of daily precipitation variability are
280
represented in the 0.25º gridded data, and monthly totals should bear resemblance to
281
observations (at least as represented by the underlying monthly data such as CRU),
282
correspondence with observed daily precipitation events is not anticipated.
283
284
With this limitation in mind, we focus on the statistics of daily hydroclimatology, rather than
285
event-based statistics. Especially given the rising interest in characterizing extreme events in the
286
context of a changing climate [IPCC, 2011], the ability of the adjusted daily gridded dataset to
287
characterize extreme statistics is important. We compute a set of statistical variables frequently
288
used to describe climate extremes, using the RClimDex software [X. Zhang and Yang, 2004;
289
Xuebin Zhang et al., 2005]. Additionally we compute the unconditional probability of
290
occurrence of wet and dry days and the corresponding transition probabilities. Figure 6 shows
291
boxplots of six statistical parameters listed in Table 1 for the four locations. Extreme
292
precipitation events (R95p) are well captured in the adjusted gridded data set at most locations.
293
The agreement between adjusted total annual precipitation (PRCPTOT) and observations is good
294
with an average bias of -9% from the observed station mean (not shown), although this is
295
constrained by design, as noted above. The Simple Daily Intensity Index (SDII), which is a
296
measure of the mean annual intensity of rainfall, also shows good agreement at the four
297
locations, indicating that the number of rainy days is well represented in the adjusted dataset.
298
The number of days with intensities larger than the 20 mm (R20mm) compares well between
299
observations and adjusted gridded precipitation, though the adjusted gridded data slightly
14
300
underestimate observations. The maximum consecutive number of dry days and wet days in a
301
year is lower for the adjusted gridded observations compared to observations suggesting the
302
durations of wet and dry events are shorter in the adjusted gridded data set.
303
304
Figure 7 shows that the probabilities of a day being wet or dry are comparable between both
305
datasets (panels 7a and 7b). Conversely the adjusted precipitation data shows an average
306
transition probability of a wet day followed by a wet day of 0.21 compared to 0.50 obtained for
307
the observations suggesting that the duration of storm events is shorter in the adjusted gridded
308
dataset. This could partially explain the underestimation of maximum consecutive wet days
309
(CWD) in the adjusted gridded precipitation as well.
310
311
The statistical quantities presented in Figures 6 and 7 were compared statistically using the
312
correlation coefficient and a two-sample unpaired Student’s t-test. These are summarized in
313
Table 4. Statistics linked to high intensity events (R99p and R95p and maximum 1-day
314
precipitation (RX1day) have statistically indistinguishable means for all four locations. The
315
mean annual precipitation and intensity parameters (Prcptot and SDII) show equal means for
316
three out of the four locations. Conversely the parameters, R5mm R20mm, albeit strongly
317
correlated, were found to have statistically different mean values. This phenomenon of a gridded
318
precipitation data set having lower extreme precipitation values than station observations was
319
also noted in the South American study of Silva et al. [2007] and is consistent with the effect of
320
spatial averaging, i.e., comparing the average of a 630 km2 0.25º grid cell to the smaller, more
321
discrete area represented by the three averaged stations [Yevjevich, 1972]. The statistics related
15
322
to duration of wet and dry spells showed statistically different population means at all 4
323
locations.
324
325
4.2 Hydrologic Model Validation of Adjusted Meteorology
326
To assess the representation in the new meteorological data set of basin-wide and high elevation
327
areas, the adjusted gridded data developed and assessed in the previous sections were then used
328
to drive the VIC hydrologic model. Since the precipitation was shown to be comparable to
329
observations (where available) in many important respects, another validation of the driving
330
meteorology would be the successful simulation of observed streamflow and snow cover.
331
Records of observed streamflow in the region tend to be incomplete or for short periods, and
332
since most of the rivers are affected by reservoirs and diversions the flows often do not reflect
333
natural streamflow as simulated by the VIC model. For this project, we focused on three sites,
334
shown in Figure 1, which have more complete records and were judged to be relatively free of
335
anthropogenic influences.
336
337
For the site on the Mataquito River, the VIC model was calibrated to monthly stream flows for
338
the period 1990-1999 using the Multi-Objective Complex Evolution (MOCOM-UA) algorithm
339
[Yapo et al., 1998]. The three optimization criteria used in this study were the Nash-Sutcliff
340
model efficiency [NSE, Nash and Sutcliffe, 1970] using both flow (NSE) and the logarithm of
341
flow (NSElog), and the bias, expressed as a percent of observed mean flow. This provides a
342
balance between criteria that penalize errors at high flows and others that are less sensitive to a
343
small number of large errors at high flows [Lettenmaier and Wood, 1993]. Figure 8 shows the
344
VIC simulation results for the calibration period and for the validation period of 2000-2007. The
16
345
flows for both periods generally meet the criteria for “satisfactory” calibration based on the
346
criteria of Moriasi et al. [2007], with a NSE > 0.50 and absolute bias < 25% (the third criterion
347
of Moriasi was not calculated for this experiment). While during the validation period several of
348
the maximum annual flow peaks are overestimated, resulting in a lower NSE score compared to
349
the calibration period, the reasonable peaks, low flows, and satisfactory calibration and
350
validation do serve to provide further validation of the driving meteorology as plausible.
351
352
Despite the highly variable precipitation across the study region, we applied the same VIC
353
calibrated parameters from the Mataquito basin to the entire domain and used the VIC model to
354
generate streamflow at the other two gage sites. This avoids the possibility of allowing extensive
355
calibration to hide meteorological data deficiencies. The simulated flows for the period 2000-
356
2007 for each site, and the associated statistics, are in Figures 9 and 10. The simulated flows on
357
average show little bias in both locations. The Claro River NSElog value is low, reflecting the
358
underestimation of low flows and overestimation of peak flows during the simulation period,
359
though the higher NSE value suggests the errors at the high flows are not as systematic. The
360
Loncomilla River displays a general overestimation by VIC of low flows, though both NSE and
361
NSElog are above the “satisfactory” threshold. While these are not demonstrations of the best
362
hydrologic model that could be developed for each basin, or the best that the VIC model could
363
produce (since no calibration was performed for two of the three basins), they do provide some
364
further validation that the driving meteorology appears plausible, and does not appear to show
365
any systematic biases.
366
17
367
Additionally, a comparison of four streamflow properties is shown in Figure 11 for the three
368
simulated basins. We calculate the center timing (CT), defined as the day when half the annual
369
(water year) flow volume has passed a given point [Stewart et al., 2005], where the water year
370
runs from April 1 through March 31. CT values lie within the -11 to 17 day window compared to
371
observed values, indicating the snow melting season is reasonably captured by the model (Figure
372
11a). The unpaired Student’s t-test indicates the distributions have equal means at a 5%
373
significance level. The water year volume and the 3-day peak flow are systematically
374
overestimated by VIC simulations, however their means are found to statistically equal with the
375
exception of the Rio Claro 3-day peak flow. Low flows are over and underestimated by VIC
376
simulations but only the Loncomilla River has means that are statistically different (Figure 11d).
377
378
Recognizing the high dependence of this region on snow melt and thus the importance of this
379
process being well represented, we validate the high elevation meteorology of the new data set
380
by comparing VIC simulated SWE to MODIS 8-day snow coverage for the six events between
381
2002 and 2007 (one image per year). The satellite images were selected to capture the snow
382
cover in mid to late August in each year, approximating the maximum snow accumulation in the
383
region. Following Maurer et al. [2003] a snow depth of 25.4 mm (1 inch) was used as threshold
384
to indicate the presence of snow on the ground. MODIS snow coverage was interpolated to a
385
0.25º grid using triangle-based cubic interpolation. VIC simulated SWE was averaged to match
386
the MODIS eight-day period. Strong similarities in the spatial extent are found between MODIS
387
and VIC simulated snow coverage for the period August 21-28, 2002 (Figure 12). The average
388
area covered by snow in the six years is 172,320 and 167,050 km2 in VIC simulations and
18
389
MODIS, respectively. This represents only a 3% error in the snow-covered area simulated by the
390
VIC model.
391
392
Table 5 is a contingency table of relative frequencies of snow/no snow in MODIS and VIC
393
simulated SWE. We include all of the pixels for the six selected periods (total 1170). The
394
number of pixels classified as snow or no snow is similar in VIC and MODIS with frequencies
395
of 0.58 and 0.29 for no snow and snow classification, respectively. Conversely the occurrence of
396
misclassified snow/no snow events is quite low, on the order of 6% indicating an excellent
397
agreement between both data sources. The Heidke Skill Score for this data is 0.72, showing that
398
the agreement between observed and VIC simulated snow cover is unlikely to be due to chance.
399
400
Finally, the successful validation of the streamflow and simulated snow cover with observations
401
also implicitly supports the gridded temperatures in the data set. Reasonable end of season snow
402
extent and well simulated timing of flows in snow-dominated streams indicates that the
403
temperatures are not likely to be greatly in error.
404
405
5. Conclusions
406
In this study an adjusted gridded daily precipitation data set is developed for Central Chile for
407
the period 1948-2008. Rain gauge data are used to correct the inaccuracies in the representation
408
of orographic distribution of precipitation existent in the available global gridded data set.
409
Adjusted gridded data are validated using station observations and hydrological model
410
simulations.
19
411
412
In data-sparse regions, a simple cokriging method that incorporates topographic elevation as
413
covariate can be successfully used to improve the spatial representation of gridded precipitation
414
in areas with complex terrain. A month-to-month adjustment can effectively remove biases in
415
precipitation values hailing from few or nonexistent rain gauge observations.
416
417
The adjusted gridded precipitation is able to capture precipitation enhancement due to orography
418
in the region with a good representation of annual totals and precipitation intensity. However the
419
duration of storm events is slightly shorter than observed perhaps as a result of comparing a 630
420
km2 grid cell to the smaller, more discrete, areal precipitation represented by three averaged rain
421
gauges. The statistics of extreme precipitation events are well captured by the adjusted gridded
422
data set, which encourages its use for climate change applications.
423
424
Streamflow simulations in three basins realistically capture high and low flows statistical
425
properties indicating that the driving meteorology in the adjusted gridded data set is well
426
represented. Simulated SWE closely resembles satellite observations which can be linked to a
427
good depiction of winter precipitation at higher elevations, despite the driving meteorological
428
dataset not including high elevation station observations. While not explicitly tested here,
429
successful simulation of snow cover and flow in snow-dominated streams indicates that
430
temperatures in the gridded data set are also reasonable, and do not require adjustment.
431
432
Based on our results, the adjusted daily gridded precipitation data set can be successfully used
433
for hydrologic simulations of climate variability and change in Central Chile. The methodology
20
434
presented in this paper can be implemented in numerous data-sparse basins located in
435
mountainous regions around the globe with one caveat. The sensitivity of the results to the
436
number of rain gauges used to obtain plausible adjusted values was not determined; therefore the
437
quality of the adjusted data set will be constrained by the density of the local observation
438
network.
439
440
441
Acknowledgements
442
This study was funded by CORFO-INNOVA grant 2009-5704 to the Centro Interdisciplinario de
443
Cambio Global at the Pontificia Universidad Católica de Chile. A Fulbright Visiting Scholars
444
Grant also provided partial support to the second author. The authors are grateful to Paul
445
Nienaber and Markus Schnorbus of the Pacific Climate Impacts Consortium, University of
446
Victoria, BC, Canada, and Katrina Bennett at the U. of Alaska, Fairbanks, for providing updated
447
and improved code for the MOCOM/VIC application.
448
21
449
References
450
Adam, J. C., and D. P. Lettenmaier (2003), Adjustment of global gridded precipitation for
451
452
systematic bias, J. Geophys Res., 108(D9), 1-14.
Bradley, R. S., M. Vuille, H. F. Diaz, and W. Vergara (2006), Threats to Water Supplies in the
453
454
Tropical Andes, Science, 312(5781), 1755-1756.
Chen, M., P. Xie, J. E. Janowiak, and P. A. Arkin (2002), Global Land Precipitation: A 50-yr
455
456
Monthly Analysis Based on Gauge Observations, J. Hydrometeorology, 3(3), 249-266.
Cherkauer, K. A., L. C. Bowling, and D. P. Lettenmaier (2003), Variable infiltration capacity
457
cold land process model updates, Global Plan. Change, 38, 151-159.
458
Cosgrove, B. A., D. Lohmann, K. E. Mitchell, P. R. Houser, E. F. Wood, J. C. Schaake, A.
459
Robock, C. Marshall, J. Sheffield, Q. Duan, L. Luo, R. W. Higgins, R. T. Pinker, J. D.
460
Tarpley, and J. Meng (2003), Real-time and retrospective forcing in the North American
461
Land Data Assimilation System (NLDAS) project, J. Geophys. Res., 108(D22), 8842.
462
Daly, C., R. P. Neilson, and D. L. Phillips (1994), A statistical-topographic model for mapping
463
climatological precipitation over mountainous terrain, Journal of Applied Meteorology, 33,
464
140-158.
465
de Goncalves, L. G. G., W. J. Shuttleworth, B. Nijssen, E. J. Burke, J. A. Marengo, S. C. Chou,
466
P. Houser, and D. L. Toll (2006), Evaluation of model-derived and remotely sensed
467
precipitation products for continental South America, J. Geophys. Res., 111(D16), D16113.
468
DGA (1987), Balance hidrico de Chile Rep., Ministerio de Obras Publicas, Direccion General de
469
Aguas, Santiago, Chile.
22
470
Diodato, N., and M. Ceccarelli (2005), Interpolation processes using multivariate geostatistics
471
for mapping of climatological precipitation mean in the Sannio Mountains (southern Italy),
472
Earth Surface Processes and Landforms, 30(3), 259-268.
473
dos Santos, C. A. C., C. M. U. Neale, T. V. R. Rao, and B. B. da Silva (2011), Trends in indices
474
for extremes in daily temperature and precipitation over Utah, USA, Int. J. Climatol., 31(12),
475
1813-1822.
476
Falvey, M., and R. Garreaud (2007), Wintertime Precipitation Episodes in Central Chile:
477
Associated Meteorological Conditions and Orographic Influences, J. Hydrometeorology,
478
8(2), 171-193.
479
Falvey, M., and R. D. Garreaud (2009), Regional cooling in a warming world: Recent
480
temperature trends in the southeast Pacific and along the west coast of subtropical South
481
America (1979–2006), J. Geophys. Res., 114(D4), D04102.
482
Garreaud, R. D., M. Vuille, R. Compagnucci, and J. Marengo (2009), Present-day South
483
American climate, Palaeogeography, Palaeoclimatology, Palaeoecology, 281(3–4), 180-195.
484
Hall, D. K., G. S. Riggs, and V. V. Salomonson (2006), Updated daily. MODIS/Terra Snow
485
Cover 8-Day L3 Global 0.05deg CMG V005, Digital media, edited, National Snow and Ice
486
Data Center, Boulder, Colorado USA.
487
Hevesi, J. A., J. D. Istok, and A. L. Flint (1992), Precipitation Estimation in Mountainous
488
Terrain Using Multivariate Geostatistics. Part I: Structural Analysis, Journal of Applied
489
Meteorology, 31(7), 661-676.
490
Huffman, G. J., R. F. Adler, M. M. Morrissey, D. T. Bolvin, S. Curtis, R. Joyce, B. McGavock,
491
and J. Susskind (2001), Global Precipitation at One-Degree Daily Resolution from
492
Multisatellite Observations, J. Hydrometeorology, 2(1), 36-50.
23
493
Huffman, G. J., D. T. Bolvin, E. J. Nelkin, D. B. Wolff, R. F. Adler, G. Gu, Y. Hong, K. P.
494
Bowman, and E. F. Stocker (2007), The TRMM Multisatellite Precipitation Analysis
495
(TMPA): Quasi-Global, Multiyear, Combined-Sensor Precipitation Estimates at Fine Scales,
496
J. Hydrometeorology, 8(1), 38-55.
497
IPCC (2011), Intergovernmental Panel on Climate Change Special Report on Managing the
498
Risks of Extreme Events and Disasters to Advance Climate Change Adaptation, Summary for
499
Policymakers, Cambridge University Press, Cambridge, United Kingdom and New York,
500
NY, USA.
501
Kalnay, E., M. Kanamitsu, R. Kistler, W. Collins, D. Deaven, L. Gandin, M. Iredell, S. Saha, G.
502
White, J. Woollen, Y. Zhu, A. Leetmaa, and B. Reynolds (1996), The NCEP/NCAR 40-year
503
reanalysis project, Bull. Am. Met. Soc., 77(3), 437-472.
504
Lawford, R. G., R. Stewart, J. Roads, H. J. Isemer, M. Manton, J. Marengo, T. Yasunari, S.
505
Benedict, T. Koike, and S. Williams (2004), Advancing Global-and Continental-Scale
506
Hydrometeorology: Contributions of GEWEX Hydrometeorology Panel, Bull. Am. Met. Soc.,
507
85(12), 1917-1930.
508
Lehner, B., K. Verdin, and A. Jarvis (2006), HydroSHEDS Technical Documentation Rep.,
509
510
[http://hydrosheds.cr.usgs.gov] pp, Washington, DC. .
Lettenmaier, D. P., and E. F. Wood (1993), Hydrologic Forecasting, in Handbook of Hydrology,
511
edited by D. R. Maidment, pp. 26.21-26.30, McGraw-Hill Inc., New York, NY, USA.
512
Liang, X., D. P. Lettenmaier, E. Wood, and S. J. Burges (1994), A simple hydrologically based
513
model of land surface water and energy fluxes for general circulation models, J. Geophys
514
Res., 99(D7), 14415-14428.
24
515
Liebmann, B., and D. Allured (2005), Daily Precipitation Grids for South America, Bull. Am.
516
517
Met. Soc., 86(11), 1567-1570.
Lohmann, D., R. Nolte-Holube, and E. Raschke (1996), A large-scale horizontal routing model
518
to be coupled to land surface parameterization schemes, Tellus, 48A, 708-721.
519
Maurer, E. P., J. C. Adam, and A. W. Wood (2009), Climate model based consensus on the
520
hydrologic impacts of climate change to the Rio Lempa basin of Central America, Hydrol.
521
Earth System Sci., 13(2), 183-194.
522
Maurer, E. P., J. D. Rhoads, R. O. Dubayah, and D. P. Lettenmaier (2003), Evaluation of the
523
snow-covered area data product from MODIS, Hydrol. Processes, 17(1), 59-71.
524
Maurer, E. P., A. W. Wood, J. C. Adam, D. P. Lettenmaier, and B. Nijssen (2002), A long-term
525
hydrologically-based data set of land surface fluxes and states for the conterminous United
526
States, J. Climate, 15(22), 3237-3251.
527
Milly, P. C. D., and A. B. Shmakin (2002), Global Modeling of Land Water and Energy
528
Balances. Part I: The Land Dynamics (LaD) Model, J. Hydrometeorology, 3(3), 283-299.
529
Mitchell, T. D., and P. D. Jones (2005), An improved method of constructing a database of
530
monthly climate observations and associated high-resolution grids, Int. J. Climatol., 25(6),
531
693-712.
532
Moriasi, D. N., J. G. Arnold, M. W. V. Liew, R. L. Bingner, R. D. Harmel, and T. L. Veith
533
(2007), Model evaluation guidelines for systematic quantification of accuracy in watershed
534
simulations, Trans. of ASABE, 50(3), 885-900.
535
Nash, J. E., and J. V. Sutcliffe (1970), River flow forecasting through conceptual models part I --
536
A discussion of principles, J. Hydrol., 10(3), 282–290.
25
537
New, M. G., M. Hulme, and P. D. Jones (2000), Representing twentieth-century space-time
538
climate variability. Part II: development of 1961-90 monthly grids of terrestrial surface
539
climate, J. Climate, 12, 829-856.
540
Nijssen, B., R. Schnur, and D. P. Lettenmaier (2001a), Global retrospective estimation of soil
541
542
moisture using the VIC land surface model, 1980-1993, J. Climate, 14(8), 1790-1808.
Nijssen, B., G. M. O'Donnell, D. P. Lettenmaier, D. Lohmann, and E. F. Wood (2001b),
543
Predicting the discharge of global rivers, J. Climate, 14, 1790-1808.
544
O'Donnell, G., B. Nijssen, and D. P. Lettenmaier (1999), A simple algorithm for generating
545
streamflow networks for grid-based, macroscale hydrological models, Hydrol. Processes,
546
13(8), 1269-1275.
547
Peterson, T. C., R. Vose, R. Schmoyer, and V. Razuvaëv (1998), Global historical climatology
548
network (GHCN) quality control of monthly temperature data, Int. J. Climatol., 18(11),
549
1169-1179.
550
Rubio-Álvarez, E., and J. McPhee (2010), Patterns of spatial and temporal variability in
551
streamflow records in south central Chile in the period 1952–2003, Water Resour. Res., 46,
552
W05514, doi:05510.01029/02009WR007982.
553
Sheffield, J., and E. F. Wood (2007), Characteristics of global and regional drought, 1950-2000:
554
Analysis of soil moisture data from off-line simulation of the terrestrial hydrologic cycle, J.
555
Geophys. Res., 112(D17), D17115.
556
Sheffield, J., G. Goteti, and E. F. Wood (2006), Development of a 50-yr high-resolution global
557
558
dataset of meteorological forcings for land surface modeling, J. Climate, 19(13), 3088-3111.
Silva, V. B. S., V. E. Kousky, W. Shi, and R. W. Higgins (2007), An Improved Gridded
559
Historical Daily Precipitation Analysis for Brazil, J. Hydrometeorology, 8(4), 847-861.
26
560
Stewart, I. T., D. R. Cayan, and M. D. Dettinger (2005), Changes toward earlier streamflow
561
562
timing across western North America, J. Climate, 18(8), 1136-1155.
Vicuna, S., R. D. Garreaud, and J. McPhee (2011), Climate change impacts on the hydrology of
563
a snowmelt driven basin in semiarid Chile, Climatic Change, 105, 469-488,
564
410.1007/s10584-10010-19888-10584.
565
Wilks, D. S. (2006), Statistical Methods in the Atmospheric Sciences, 2 ed., 627 pp., Academic
566
567
doi:
Press, New York, NY, USA.
Wilks, D. S., and R. L. Wilby (1999), The weather generation game: a review of stochastic
568
weather models, Prog. Phys. Geography, 23(3), 329-357.
569
Willmott, C. J., and K. Matsuura (2001), Terrestrial air temperature and precipitation: monthly
570
and annual time series (1950–1999) (Version 1.02) Rep., Center for Climatic Research,
571
University of Delaware, Newark, DE, USA.
572
Yapo, P. O., H. V. Gupta, and S. Sorooshian (1998), Multi-objective global optimization for
573
574
hydrologic models, J. Hydrol., 204, 83-97.
Yevjevich, V. (1972), Probability and Statistics in Hydrology, Water Resources Publications, Ft.
575
576
Collins, CO, USA.
Zhang, X., and F. Yang (2004), RClimDex (1.0) User Guide Rep., Climate Research Branch,
577
Environment Canada, Downsview, Ontario, Canada.
578
Zhang, X., E. Aguilar, S. Sensoy, H. Melkonyan, U. Tagiyeva, N. Ahmed, N. Kutaladze, F.
579
Rahimzadeh, A. Taghipour, T. H. Hantosh, P. Albert, M. Semawi, M. Karam Ali, M. H. Said
580
Al-Shabibi, Z. Al-Oulan, T. Zatari, I. Al Dean Khelet, S. Hamoud, R. Sagir, M. Demircan,
581
M. Eken, M. Adiguzel, L. Alexander, T. C. Peterson, and T. Wallis (2005), Trends in Middle
582
East climate extreme indices from 1950 to 2003, J. Geophys. Res., 110(D22), D22104.
27
583
584
Table 1 - List of statistical quantities and descriptions.
Name
Description
R95p
Annual total precipitation when rainfall > 95th percentile
R99p
Annual total precipitation when rainfall > 99th percentile
PRCPTOT
Annual total precipitation in wet days (with daily rainfall, RR >=1mm)
CWD
Consecutive wet days: largest number of consecutive wet days with RR >=1mm
CDD
Consecutive dry days: largest number of consecutive dry days with RR <=1mm
SDII
Simple daily intensity index: mean annual intensity for RR > = 1 mm
R5mm
Annual count of days with Precipitation >= 5 mm
R20mm
Annual count of days with Precipitation >= 20 mm
RX1d
Maximum 1 day precipitation in the year
RX5d
Maximum 5 days precipitation in the year
PW
Probability of Wet days
PD
Probability of Dry days
PWW/PDD
Probability of a wet /dry day followed by a wet/dry day
PWD/PDW
Probability of a wet/dry day following a dry/wet day
585
586
587
28
588
589
Table 2 - Location of adjusted gridded precipitation grid cells used in daily precipitation validation.
0.25º Grid Cell Abbreviation
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
29
Grid Cell Center Latitude
Grid Cell Center Longitude
Loc1
-34.375
-70.875
Loc2
-34.875
-71.125
Loc3
-35.875
-71.125
Loc4
-36.125
-71.625
605
Table 3 - Daily precipitation statistics for summer (DJF) and winter (JJA) periods, 1983-2007. OBS are
606
observations, ADJ are adjusted gridded meteorology.
Summer (DJF) Statistics
OBS Mean
ADJ Mean
OBS Std
ADJ Std
Mean Bias
RMSE
Correlation
(mm)
(mm)
(mm)
(mm)
(mm)
(mm)
Loc1
0.08
0.08
1.21
0.63
0.00
1.37
-0.01
Loc2
0.14
0.11
1.61
0.59
-0.03
1.71
0.02
Loc3
0.64
0.60
5.44
2.29
-0.04
5.88
0.01
Loc4
0.60
0.54
4.38
2.34
-0.07
4.99
-0.01
Winter (JJA) Statistics
607
608
609
610
611
612
613
614
615
616
30
Loc1
3.80
3.56
10.88
13.96
-0.23
17.16
0.06
Loc2
5.64
4.72
14.17
16.72
-0.92
20.99
0.09
Loc3
13.63
12.58
30.46
37.39
-1.05
46.18
0.08
Loc4
7.24
7.36
14.51
22.08
0.12
25.64
0.06
617
Table 4 - Correlation coefficients between observed and adjusted daily precipitation statistical parameters.
618
Bold values indicate the null hypothesis of equal means cannot be rejected at the 5% level based on a t-test.
R99p
R95p
Loc1
0.10
0.70
0.96
0.76
0.58
Loc2
0.40
0.65
0.92
0.65
Loc3
0.18
0.47
0.83
Loc4
0.31
0.29
0.87
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
31
PRCTOT
SDII
R5mm
R20mm
CWD
CDD
RX1d
RX5d
0.82
0.37
0.45
0.78
0.61
0.61
0.84
0.43
0.38
0.79
0.53
0.53
0.61
0.65
0.25
0.73
0.45
0.48
0.57
0.71
0.73
0.00
0.28
0.56
0.37
634
Table 5 - Contingency table summarizing the comparisons of MODIS and VIC simulated snow cover. Values
635
are relative frequencies calculated as the total number of occurrences in each category divided by the number
636
of pixels (1170).
VIC
MODIS
637
638
32
No Snow
Snow
Total
No Snow
0.58
0.06
0.64
Snow
0.07
0.29
0.36
Total
0.65
0.35
1.0
639
List of Figures
640
641
Figure 1 - Geographic location of the study area in Central Chile. From north to south the basins
642
are: Rapel, Mataquito (Mataquito river at Licanten), Maule (Claro river at Rauquen and
643
Loncomilla river at Bodega) and Itata river basins. Circles indicate the location of DGA rain
644
gauges and stars the location of the three stream gauges used in VIC simulations
645
Figure 2 - Maps of annual precipitation for the period 1951-1980. Source a) gridded global
646
observations and b) DGA. Precipitation lapse rates for latitudinal bands -35.125 S and -36.125 S
647
for c) global gridded precipitation data set and d) DGA data set.
648
Figure 3 - Scatterplots of observed and predicted average monthly precipitation for the month of
649
July. Each point represents one observation station (abscissa) and the 0.25 grid cell in which the
650
grid cell lies (ordinate).
651
Figure 4 - a) Annual adjusted global precipitation for the period 1950-2006 and b) differences
652
between the original global gridded and the adjusted global precipitation data sets.
653
Figure 5 - Location of DGA rain gauge stations and adjusted global precipitation grid points used
654
for validation of daily rainfall.
655
Figure 6 - Boxplots of statistical parameters, green represents observations and purple represents
656
adjusted precipitation for each geographic location. The bottom and top lines represent the 25th
657
and 75th percentiles and the middle line represents the median. Whiskers extend from each end
658
of the box to the adjacent values in the data within 1.5 times the Inter Quartile Range. The Inter
659
Quartile Range is the difference between the third and the first quartile, i.e., 25th and 75th
660
percentiles. Outliers are displayed with a plus sign.
33
661
Figure 7 - Probabilities of a) wet and b) dry days, and transition probabilities c) and d). Daily
662
observed (black) and adjusted gridded (grey) precipitation for the four selected locations.
663
Figure 8 - Observed and Simulated monthly flows for the Mataquito river at Licanten for the
664
calibration period (top panel) and validation period (bottom panel). Summary statistics are
665
shown in each panel.
666
Figure 9 - Monthly observed and simulated flows for the Claro river at Rauquen.
667
Figure 10 - Same as Figure 9 but for the Loncomilla river at Bodega.
668
Figure 11 - Statistical properties of observed and VIC simulated stream flows in three basins:
669
Mataquito river, Claro river and Loncomilla river. (a) Center timing, (b) water year volume, (c)
670
3-day peak flows and (d) 7-day low flows.
671
Figure 12 - Comparison of snow coverage for the period August 21-28, 2002. Shaded areas
672
indicate snow coverage. a) MODIS and b) VIC simulated Snow Water Equivalent.
673
34
674
675
Figure 1 - Geographic location of the study area in Central Chile. From north to south the basins are: Rapel,
676
Mataquito (Mataquito river at Licanten), Maule (Claro river at Rauquen and Loncomilla river at Bodega)
677
and Itata river basins. Circles indicate the location of DGA rain gauges and stars the location of the three
678
stream gauges used in VIC simulations
679
35
680
681
Figure 2 - Maps of annual precipitation for the period 1951-1980. Source a) gridded global observations and
682
b) DGA. Precipitation lapse rates for latitudinal bands -35.125 S and -36.125 S for c) global gridded
683
precipitation data set and d) DGA data set.
684
685
686
36
687
688
Figure 3 - Scatterplots of observed and predicted average monthly precipitation for the month of July. Each
689
point represents one observation station (abscissa) and the 0.25 grid cell in which the grid cell lies (ordinate).
690
691
37
692
693
Figure 4 - a) Annual adjusted global precipitation for the period 1950-2006 and b) differences between the
694
original global gridded and the adjusted global precipitation data sets.
695
696
38
697
698
Figure 5 - Location of DGA rain gauge stations and adjusted global precipitation grid points used for
699
validation of daily rainfall.
700
701
39
702
703
Figure 6 - Boxplots of statistical parameters, green represents observations and purple represents adjusted
704
precipitation for each geographic location. The bottom and top lines represent the 25th and 75th percentiles
705
and the middle line represents the median. Whiskers extend from each end of the box to the adjacent values
706
in the data within 1.5 times the Inter Quartile Range. The Inter Quartile Range is the difference between the
707
third and the first quartile, i.e., 25th and 75th percentiles. Outliers are displayed with a plus sign.
708
40
709
710
Figure 7 - Probabilities of a) wet and b) dry days, and transition probabilities c) and d). Daily observed
711
(black) and adjusted gridded (grey) precipitation for the four selected locations.
712
713
41
714
715
Figure 8 - Observed and Simulated monthly flows for the Mataquito river at Licanten for the calibration
716
period (top panel) and validation period (bottom panel). Summary statistics are shown in each panel.
717
718
42
719
720
Figure 9 - Monthly observed and simulated flows for the Claro river at Rauquen.
721
43
722
723
Figure 10 - Same as Figure 9 but for the Loncomilla river at Bodega.
724
725
44
726
727
Figure 11 - Statistical properties of observed and VIC simulated stream flows in three basins: Mataquito
728
river, Claro river and Loncomilla river. (a) Center timing, (b) water year volume, (c) 3-day peak flows and
729
(d) 7-day low flows.
730
731
45
732
733
Figure 12 - Comparison of snow coverage for the period August 21-28, 2002. Shaded areas indicate snow
734
coverage. a) MODIS and b) VIC simulated Snow Water Equivalent.
735
46
Download