Вы находитесь на странице: 1из 19

1592 J O U R N A L OF A P P L I E D M E T E O R O L O G Y AND C L I M A T O L O G Y VOLUME 52

Analysis of WRF Model Wind Estimate Sensitivity to Physics Parameterization


Choice and Terrain Representation in Andalusia (Southern Spain)
F. J. SANTOS-ALAMILLOS, D . POZO-VÁZOUEZ, J. A. RUIZ-ARIAS,
V. L A R A - F A N E G O , AND J. T O V A R - P E S C A D O R

Physics Department, University of Jaén, Jaén, Andalusia, Spain

(Manuscript reeeived 23 July 2012, infinalform 24 February 2013)

ABSTRACT
This paper reports on an evaluation of the relative roles of choice of parameterization scheme and terrain
representation in the Weather Research and Forecasting (WRF) mesoscale model, in the context of a regional
wind resource assessment. As a flrst step, 32 configurations using two different schemes for microphysies,
cumulus, planetary boundary layer (PBL), or shortwave and longwave radiation were evaluated. In a second
step, wind estimates that were obtained from various experiments with different spatial resolution (1, 3, and
9 km) were assessed. Estimates were tested against data from four stations, located in southern Spain, that
provided hourly wind speed and direction data at 40 m above ground level. Results from the first analysis
showed that wind speed standard deviation (STD) and bias values were mainly sensitive to the PBL pa-
rameterization selection, with STD differences up to 10% and bias differences between -15% and 10%. The
second analysis showed a weak influence of spatial resolution on the STD values. On the other hand, the bias
was found to be highly sensitive to model spatial resolution. The sign of the bias depended on terrain mor-
phology and the spatial resolution, but absolute values tended to be much higher with coarser spatial reso-
lution. Physical configuration was found to have little impact on wind direction distribution estimates. In
addition, these estimates proved to be more sensitive to the ability of WRF to represent the terrain mor-
phology around the station than to the model spatial resolution itself.

1. Introduction a comprehensive description of the atmospheric system,


Reliable wind resource assessment is a key issue in the using initial and boundary conditions from atmospheric
evaluation of wind energy potential and in studies re- reanalysis. On the basis of dynamical downscaling tech-
lated to integration of wind energy electricity produc- niques, current NWP models resolve local and regional
tion (GE Energy 2010). Such wind resource assessment circulation patterns, taking into account local topo-
can be divided into two steps (Sempreviva et al. 2008)— graphic surface features to a certain extent. Thus, these
a regional wind assessment, which evaluates resources of models can provide accurate meteorological informa-
a region as a flrst approach, and a site-speciflc assess- tion, such as wind speed and direction, over large regions
ment, which is a detailed evaluation of speciflc areas that at high temporal (a few minutes) and spatial (a few
exhibit high potential in the regional assessment. At this kilometers) resolution.
time, regional wind assessment is mostly based on nu- Despite the high resolution of current NWP models
merical weather prediction (NWP) models (Frank et al. (~km), there are still two main problems in the esti-
2001; Landberg et al. 2003; Lavagnini et al. 2006; Zagar mation of surface meteorological variables, especially
et al. 2006; Kanamitsu and Kanamaru 2007; Byrkjedal near-surface wind speed. These problems are 1) repre-
and Berge 2008; Jiménez et al. 2007; Hahmann et al. sentation of topography and terrain morphology in the
2010; Shimada and Ohsawa 2011). NWP modeis facilitate modeled domains and 2) representation of the numer-
ous physical processes at subgrid scale.
Topography and terrain morphology have a marked
influence on surface meteorological variables, especially
Corresponding author address: D. Pozo-Vázquez, Physics De-
partment, University of Jaén, Campus Las Lagunillas s/n Edif A3, on surface wind speed and direction. Surface winds re-
E23071, Jaén, Spain. sult from interaction between mesoscale circulation and
E-mail: dpozo@ujaen.es other more local factors, many of which are related to
DOI: 10.1175/JAMC-D-12-0204.1

© 2013 American Meteorological Society


JULY 2013 SANTOS-ALAMILLOS ET AL. 1593

topographic characteristics (elevation, aspect, and slope) evaluation of different parameterization versions of the
and terrain morphology (hills, valleys, and others). First, various schemes, for the given study area and applica-
wind speed changes with elevation in the planetary tion, is advisable.
boundary layer (PBL) as a result of reduced turbulent The WRF (Skamarock et al. 2008) model is one of
drag. Second, terrain features can accelerate wind flow. the most widely used NWP models, being a reference
As it passes a summit, an approaching air mass is often for regional wind resource assessment. Over recent
squeezed into a thinner layer and therefore speeds up. decades, many different parameterizations have been
Aside from elevation, terrain morphology (terrain aspect proposed for this model, especially for the PBL. PBL
and slope) can also modify wind flow. Over a ridge, there parameterizations are responsible for vertical subgrid-
is maximum acceleration when wind blows perpendicular scale fluxes, allowing eddy transports through the entire
to the ridge line. Terrain morphology is also important; atmospheric column. Wind distribution is affected by
wind flow accelerates through passes along valleys factors such as PBL height, wind shear, and entrainment
aligned with the flow (channeling). In a similar way, of free atmospheric air into the PBL, which determines
topography may produce local areas of reduced wind momentum, heat, and moisture exchanges at the top of
speed, such as sheltered valleys and areas in the lee of the layer (Arya 2001). These processes may vary con-
a mountain ridge. Last, it is also well known that in siderably depending on, for instance, the study region or
complex-topography areas strong horizontal tempera- spatial and temporal scales. As a consequence, many
ture gradients, caused by the sloping terrain, generate authors have reported strong sensitivity of WRF (and
mountain and valley wind regimes. If the study region is other NWP) surface wind estimates to PBL parame-
composed of flat and homogeneous land. Weather Re- terization, in many areas of the world: the United States
search and Forecasting model (WRF) terrain repre- (Zhang and Zheng 2004; Watson et al. 2009; Hu et al.
sentation may be accurate enough to provide reliable 2010; Gibbs et al. 2011), South Korea (Kim et al. 2005;
near-surface wind estimates. This representation, how- Kwun et al. 2009), Spain (Pérez-Landa et al. 2007; Borge
ever, may provide unreliable surface wind estimates in et al. 2008), and Sweden (Miao et al. 2008).
complex-topography areas, even using high spatial res- For other NWP parameterizations, the influence of
olutions. This issue is usually addressed in specific site selected schemes on surface wind speed estimates, al-
assessment, using microscale models such as the Wind though less important, could be significant. For instance,
Atlas Analysis and Application Program (WAsP; MPH schemes include explicitly resolved water vapor,
Mortensen et al. 1998). This type of model provides site- cloud, and precipitation processes. According to Rajeevan
specific wind estimates that are based on terrain in- et al. (2010), cloud microphysical processes are important
formation and wind measurements from nearby stations because of their direct influence on cold-pool strength
and are at very high spatial resolution (—100 m). Wind (from rainfall evaporation) and latent heating (from
estimates from WAsP are constrained by the availability condensation). Therefore, MPH processes can affect
of local measurement data and by the complexity of the wind fields close to the surface. SWR parameterizations
study area. In the last few years, several methods have include visible and near-visible wavelengths, whereas
been proposed for using NWP wind estimates as input to LWR parameterizations include infrared or longwave
microscale models like WAsP (Frank et al. 2001; Berge radiation absorbed and emitted by gases and surfaces. In
et al. 2007). complex topography, differences in surface radiation
Physical subgrid processes are taken into account in generate important thermal contrasts that in turn create
NWP models by the use of parameterization schemes. In local wind regimes. CMS schemes represent subgrid-
particular, most models include schemes to account for scale effects of convective and/or shallow clouds. These
cloud microphysics (MPH), cumulus formation (CMS), schemes are mainly intended to account for vertical
PBL processes, surface-layer physics, land surface layer, fluxes from unresolved updrafts and downdrafts, and
or shortwave and longwave radiation (SWR and LWR, compensating motions outside clouds. CMS parame-
respectively). Stensrud (2007) presents some of these terization schemes are largely evaluated for precip-
parameterizations. In the last decades, many different itation estimates (Kotroni and Lagouvardos 2001),
parameterization schemes have been proposed. Their and literature focusing on wind estimates is scarce
validity is highly dependent on the study region, season (Cohen 2002). Clouds have a significant effect on PBL
of the year, and spatial and temporal scales of interest, processes, however, affecting hydrologie processes
however. In fact, given a spatial and temporal configu- through redistributions of sensible and latent heats
ration, the skill of a model mainly depends on these and momentum (Arakawa 2004). Therefore, different
physical parameterizations (Lynn et al. 2004; Fernández CMS parameterization schemes may produce varying
et al. 2007). As a consequence, a prior performance surface wind estimates.
1594 J O U R N A L OF A P P L I E D M E T E O R O L O G Y AND C L I M A T O L O G Y VOLUME 52

FIG. 1. Location (inset at bottom left) and principal topographic and geographic charac-
teristics of the study region. The locations of the four validation stations are highlighted with
crosses. Grayscale colors indicate elevation above sea level as based on a 90-m-resolution
DEM. The scale at the right indicates elevation in meters above sea level.

Some authors (Jiménez and Dudhia 2012) have re- conditions that make the evaluation especially chal-
ported a positive bias in WRF wind speed estimates over lenging. Estimates are tested against data from four
plains and valleys [about lms~'at 10m above ground stations that provide hourly wind speed and direction
level (AGL)] and a negative bias over hills and mountains data at 40 m AGL. Station locations are representative
(about - 3 m s"^ at 10 m AGL). The main reason for the of different topographic and climatic conditions, thus
positive bias seems to be unresolved topographic fea- allowing a comprehensive evaluation of the WRF wind
tures, such as hills or mountains in complex terrain. These estimates.
can produce additional drag, which is not parameterized To sum up, the main objectives of this work are 1) to
in the WRF, in addition to the drag generated by vege- obtain a general model setup to be used for a long in-
tation (Beljaars et al. 2004; Jiménez and Dudhia 2012). tegration to be used toward evaluating wind resources in
The magnitude of this bias is related to the differences the study region, 2) to evaluate the extent to which this
between the actual and WRF-modeled station elevations. single setup can provide accurate wind estimates in
This work evaluates the relative roles of terrain rep- a region that is characterized by strongly variable cli-
resentation and choice of parameterization schemes in mate conditions, and 3) to assess the extent to which an
the WRF as they relate to regional wind resource as- increase in WRF spatial resolution gives better esti-
sessment. To evaluate the sensitivity of WRF wind es- mates in a region that is characterized by the existence of
timates to the choice of parameterization schemes, we areas with complex topography.
first evaluated 32 physical configurations resulting from The work is organized as follows: section 2 describes
combination of two different PBL, MPH, CMS, SWR, the study region, terrain characteristics, and data. Section 3
and LWR parameterization schemes. For each scheme, details the methods, and section 4 gives the results.
we evaluated one widely used parameterization and The main conclusions are presented in section 5.
a new one available in the recent WRF release (version
3.2; V3.2). Therefore, a by-product of this work is a
2. Study region and data
comparative evaluation of the new schemes against the
old ones. In a second step, to evaluate sensitivity of the The study region (Fig. 1) of Andalusia is in the south-
wind estimates to the terrain representation and mor- em Iberian Peninsula (IP). It extends over 87000km^.
phology in the WRF, we also evaluated wind speed and The region (35°30'-38°30'N, 7°30'-l°30'W) is in a tran-
direction estimates obtained from different experiments sition zone between temperate and subtropical climates.
at 1-, 3-, and 9-km spatial resolution. From a topographic point of view, the region may be
The work applies to the region of Andalusia (southern split into two different parts. The western part, covering
Spain). Although relatively small, this region is charac- ~40000km^, is a nearly homogeneous flat area that is
terized by varied topographic, geographic, and weather around 100 m MSL. The eastern part has very complex
JULY 2013 S A N T O S - A L A M I L L O S ET A L . 1595

TABLE 1. The main topographic characteristics at the station locations. Displayed are elevations at station locations from a 90-m-spatial-
resolution DEM and from the WRF model using three different spatial resolutions (9,3, and 1 km). The standard deviation of elevation
(denoted as cr) and the relief (denoted as mde) parameters are also shown for each station. Both parameters were derived from the DEM
and from the 1-km-spatial-resolution WRF domain and were computed over an area of 3 km x 3 km centered at each station location. All
values are presented in meters.

Station ElevWRF,
Elev DEM Elev WRF, 9 km Elev WRF 3 km ElevWRF, 1km o-DEM mde DEM o-WRF 1km mde WRF 1km
Centenar 236 229 253 253 17 152 13 36
Gomera 640 405 464 515 67 310 49 178
Almirez 1195 1052 1178 1182 58 270 55 215
Los Nietos 1056 1262 1013 1055 38 158 33 108

topography with several mountain areas, reaching were furnished by the wind division of Gamesa Corpo-
3482 m MSL in Sierra Nevada National Park (the highest ration and include hourly-averaged wind speed and di-
elevation on the IP). rection data measured at 40 m AGL for 16 days in 2005.
The atmospheric circulation over most of the IP is The 16 days at each station were selected as represen-
dominated by a semipermanent, subtropical high pres- tative of different periods in that year—data from 2 to 5
sure center over the Azores Islands, particularly so in February, from 9 to 12 April, from 28 to 31 August, and
the south part of the study region. The position and in- from 7 to 10 October were used to represent winter,
tensity of this center changes throughout the year. spring, summer, and autumn, respectively. We conducted
During winter, the high is at lower latitudes, allowing the a quality-control procedure that was based on Jiménez
region to be affected by zonal circulations from the west. et al. (2010) and found no suspicious data. We used
During summer, the high migrates northward, blocking wind data collected at 40 m AGL for the evaluation
off westerly circulations over most of the IP and par- rather than the much more common wind data at 10 m
ticularly over the study region (Pozo-Vázquez et al. AGL that are measured at most meteorological sta-
2001; Trigo et al. 2002; Castro-Diez et al. 2002). Al- tions. This procedure gives the results greater repre-
though the entire region is under the influence of the sentativeness for regional wind resource assessment,
Azores subtropical high, it has a wide range of climate mainly because wind speeds measured at the lower
conditions. This range is related to topographic fea- elevation are frequently influenced by local features
tures. The eastern part is less influenced by the Azores near the meteorological stations (buildings, vegetation,
high and is more affected by the Mediterranean Sea and others), which are definitely not captured in an
(Esteban-Parra et al. 1998; Rodríguez-Puebla et al. 1998; NWP model.
Ramos-Calzado et al. 2008). The western part is the The areas containing the four masts have very dif-
lower Guadalquivir River basin. It is open to the At- ferent terrain complexity and morphology. Since these
lantic Ocean and to direct influence by atmospheric flow factors have a marked influence on surface wind speed
controlled by the Azores high. The Guadalquivir valley and direction, a topographic characterization was at-
is exposed to mild and moist air from the Atlantic year- tempted. This characterization is important to the dis-
round except during summer. Winds are channeled cussion of the evaluation result in section 4. Many
through this valley, generating strong mean surface wind different parameters for characterizing terrain com-
speeds in the upper valley (Alsamamra 2009). The plexity have been defined in the literature (Huaxing
eastern part is partially isolated from this Atlantic in- 2008). In this work, we consider two types of parameters:
fluence by the Sierra Nevada and Cazorla Mountain the standard deviation of elevation (denoted as cr) and
ranges (Fig. 1). Annual precipitation over the region the difference between the highest and lowest elevation
consequently ranges from 300mmyr"^ in coastal, semi- in a specific region (denoted as mde). We derived both
desert southeast areas to more than 2500mmyr~^ in parameters from a 90-m-spatial-resolution digital ele-
mountains near the Strait of Gibraltar (Ramos-Calzado vation model (DEM; Farr et al. 2007). In particular, on
et al. 2008). In addition, the study region is bounded on the basis of the elevations reported by this DEM, we
the south by the Atlantic (western part) and the Medi- computed a and mde on an area of 3 km X 3 km cen-
terranean (eastern part), representing about 900 km tered at each mast location. Both a and mde represent
of coastline. One of the most important wind features subgrid-scale topographic complexity at the station lo-
is strong, semipermanent winds near the Strait of cations. Table 1 shows these parameters for the four
Gibraltar coast. stations, along with station elevations derived from the
Validation of WRF estimates was carried out by using DEM and WRF. The terrain topography and morphology
data from four meteorological masts (Fig. 1). These data surrounding the stations are represented in Figs. 2 and 3.
1596 J O U R N A L OF A P P L I E D M E T E O R O L O G Y AND C L I M A T O L O G Y VOLUME 52

FIG. 2. Topographic maps for the (a) Centenar, (b) Gomera, (c) Almirez, and (d) Los Nietos stations. Each map is
10 km X 10 km, centered at the station location, and was obtained by using a 10-m-resolution DEM. The scale at the
right indicates elevation in meters above sea level.

Figure 2 shows a detailed elevation map around the complexity. In fact, this area has the smallest values of
station locations that was obtained from a 10-m-spatial- a and mde (17 and 152 m, respectively) and therefore
resolution DEM. Figure 3 also shows elevations maps the least topographic complexity among the analyzed
around the stations, derived from a 90-m DEM and the stations (Table 1). Elevation in the WRF integrations
WRF domains using the three spatial resolutions (1, 3, ranges from 229 (WRF 9 km) to 253 (WRF 1 km) m MSL,
and 9 km). For better understanding of the terrain whereas the elevation in the DEM is 236 m MSL. This
morphology, observed wind direction distributions for WRF under/overestimation of station elevation seems to
the whole evaluation period (16 days) at the four sta- be associated with the spatial averaging, since there are
tions are also shown in Fig. 3. hills near the station (Figs. 2a and 3a). In this case, the
Centenar station (Fig. 1) is in the western part of the coarser spatial resolution better represented the terrain
study region, relatively close to the Atlantic Ocean. elevation. Nevertheless, as observed in Fig. 3d, terrain
Thus, the climatological behavior at this site is strongly morphology is more realistic at 1-km spatial resolution
influenced by mild and humid Atlantic winds through- than at coarser resolutions (Figs. 3b and 3c). Wind di-
out the year. The station is atop a hill (Fig. 2a), and rection variability at Centenar is substantially higher than
the surrounding area has relatively low topographic at the other stations (Fig. 3e).
JULY 2013 SANTOS-ALAMILLOS ET AL. 1597

Centenar
• [0.0 : 2.0]
• [2.0 : 4.0]
• [4.0 : 6.0]
B [6.0 : 8.0]
I 1 [8.0 ; 10.0]
• [10.0 : 12.0]
• [12.0 : inf]

Gomera
[0.0 ; 2.0]
[2.0 : 4.0]
[4.0 : 6.0]
[6.0 : 8.0]
[8.0 : 10.0]
[10.0 : 12.0]
[12.0 : inf]

Almirez
• • [0.0 : 2.0]
• • [2.0:4.0]
• • [4.0 : 6.0]
^m [6.0 : 8.0]
CZZl [8.0 : 10.0]
^m [10.0 : 12.0]
• • [12.0 : infl

Los Nietos
• [0.0 : 2.01
• [2.0 : 4.0]
• [4.0 : 6.0]
• [6.0 : 8.0]
3 [8.0 : 10.0]
• [10.0 : 12.0)
• [12.0 : inf]

FiG. 3. Terrain morphology and observed wind direction distribution for the four validation stations for the whole study period
(16 days). Shown are the (a) elevation map of a 50 km x 50 km area centered on Centenar station, as derived from 90-m DEM, and the
elevation maps used in the (b) 9-, (c) 3-, and (d) 1-km-spatial-resolution WRF setups for the Centenar station, also showing areas of
50 km X 50 km centered on the station. The scales at the right in (a)-(d) indicate elevation in meters above sea level, (e) Wind direction
distribution observed at Centenar station at 40 m AGL. The values in the circle indicate percentage of observations; the color scale at right
indicates ranges of wind speed (ms~'). Also shown are plots as in (a)-(e), but for the (f)-û) Gomera, (k)-(o) Almirez, and (p)-(t) Los
Nietos stations.

Gomera station is in the middle of the study region, on second greatest topographic complexity after Gomera
the southern boundary of the Guadalquivir valley (Fig. 1). (Table 1). As in the previous case, differences in station
The station is also atop a hill (Fig. 2b), at an elevation elevation from the different WRF conflgurations are
of 640 m MSL. The area around the station has the considerable—from 1052 (WRF 9 km) to 1182 (WRF
greatest topographic complexity among the analyzed 1 km) m MSL. Morphology of the surrounding topog-
stations, as derived from values of a (67 m) and mde raphy clearly dominates the wind direction distribution
(310 m) (Table 1). As a consequence and as expected, (Fig. 2o), enhancing the frequency corresponding to the
differences in the station elevation from the various southeast and northwest directions. Note that the
WRF conflgurations are considerable—from 405 (WRF strongest winds are mainly from the northwest. Unlike
9 km) to 515 (WRF 1 km) m MSL. In this case, elevation in the previous cases, station elevation (Table 1) and
and, of note, terrain morphology are much better rep- surrounding terrain morphology seem reasonably rep-
resented by the 1-km WRF integration (Fig. 3i). The resented in both the 3- and 1-km WRF integrations
hilltop location and morphology of the surrounding (Figs. 3k, 3m, and 3n). This will be conflrmed when an-
terrain clearly determine the wind direction distribution alyzing the results in section 4.
(Fig. 3j), through enhancing the southeast direction. Los Nietos station is also in the east, at high elevation
Almirez station is in the far east of the region (Fig. 1), (1056 m MSL) (Fig. 1), with climate conditions that are
on the middle of a plateau of considerable elevation very similar to those of Almirez. The main feature of this
(>1000m MSL), facilitating continental climate condi- station is its location in the middle of a valley, with rel-
tions. Station elevation is 1195 m MSL, in the foothills of atively low topographic complexity (Fig. 2d). As a con-
a mountain range and near a valley (Fig. 2c). It has the sequence, this station has the second-lowest a and mde
1598 J O U R N A L OF A P P L I E D M E T E O R O L O G Y AND C L I M A T O L O G Y VOLUME 52

FIG. 4. Spatial configuration of domains used for numerical simulations: (a) the three nested domains at 27-, 9-, and
3-km spatial resolution for Andalusia (southern Spain) that were used in the first analysis and (b) the four domains
centered on Centenar station at 27-, 9-, 3-, and 1-km spatial resolutions that were used in the second analysis. Similar
spatial configurations, but centered on the corresponding station locations, were used for the remaining stations.

values after Centenar (Table 1). Thus, as expected, direction estimates. In this endeavor, the model config-
terrain morphology (Figs. 3p, 3r, and 3s) and station uration included three nested domains, with 27-, 9-, and
elevation (Table 1) are reasonably reproduced in both 3-km spatial resolutions (Fig. 4a). Data were ultimately
the 3- and 1-km WRF configurations. Wind direction at analyzed for the third nested domain (3 km), centered
this station is clearly controlled by the valley effect, and on the study region. A two-way interaction was used in
its distribution is the narrowest of the four that were the three domains, with 36 vertical levels, 8 of which
analyzed (Fig. 3t). were within the first 1000 m above ground. The vertical
To sum up, the four station locations have different levels were configured following Hahmann et al. (2010),
climate conditions, topographic complexity, and terrain Draxl (2012), and Horvath et al. (2012). The first five
morphology. Los Nietos and Almirez stations are at el- lowest vertical levels—at approximately 4, 18, 43, 64,
evations above 1000 m MSL in the mountainous eastern and 76 m above the modeled ground level—^were used in
part of the study region and therefore have cold con- a vertical interpolation to estimate the wind speed and
tinental climates. In contrast, the Centenar station is direction at the measurement elevation (40m AGL).
at relatively low elevation in the eastern Guadalquivir Topography, land use, and land/water-mask datasets
River basin, which is an area that is characterized by were interpolated from the U.S. Geological Survey
a mild oceanic climate. For topography, the Gomera (USGS) land cover dataset (Loveland et al. 2000), with
area has the highest topographic complexity, followed appropriate spatial resolution for each domain (5', 1',
by Almirez. On the other hand, Los Nietos and, espe- and 30", respectively).
cially. Centenar are in areas of relatively low topographic The second study was on the effect of topography rep-
complexity. The effect of terrain morphology near each resentation on the wind speed and wind direction estimates.
station is important for wind direction distribution— In this effort, the model configuration included four nested
particularly for Gomera, Almirez, and Los Nietos. domains, with 27-, 9-, 3-, and 1-km spatial resolutions. In
this case, a one-way interaction was used in all domains.
3. Methods Moreover, as in the previous study, we also used 36 levels
for the vertical configuration of the domains. An example
The WRF model, version 3.2, was evaluated in this of this spatial configuration is shown in Fig. 4b (Centenar
work. Model runs were conducted for the days men- case). Topography, land use, and land/water-mask data-
tioned in section 2, and two different spatial configura- sets were interpolated from the USGS land cover, with
tions were used, depending on the study. appropriate spatial resolution for each domain (5' and 2'
The first study was on the influence of choice of pa- for the domains with 27- and 9-km spatial resolution,
rameterization scheme on the wind speed and wind respectively, and 30" for the 3- and 1-km domains).
JULY 2013 SANTOS-ALAMILLOS ET AL. 1599

TABLE 2. Schemes used to configure the 32 experiments for PBL, MPH, CMS, or SWR and LWR parameterizations. The center column
shows widely used schemes that were implemented in older versions of the WRF model and the right column shows schemes that are
newly implemented in WRF, version 3.2. Here, MYNN is Mellor-Yamada-Nakanishi-Niino.

Parameterization Routine schemes New schemes in WRF V3.2


PBL YSU (Pl; Hong et al. 2006) MYNN (P2; Nakanishi and Niino 2004)
MPH Morrison-2M (Ml; Morrison and Gettelman 2008) Thompson (M2)
CMS Kain-Fritsch (Cl; Kain and Fritsch 1993) Grell-3D (C2)
SWR Goddard (SI; Chou and Suarez 1994) RRTMG_SW (S2)
LWR RRTM (LI; Mlawer et al. 1997) RRTMG.LW (L2)

In the first study, 32 physical configurations (Table 2) variations of the PBL parameterization were consid-
were evaluated. This evaluation was conducted with a ered, keeping the rest of the parameterizations fixed.
3-km spatial resolution. These configurations were com- This approach considerably reduces the complexity of
posed as a combination of two different schemes for the the analysis. In particular, in the second study, we
PBL, MPH, CMS, SWR, and LWR parameterizations— evaluated the WRF wind estimates at three different
one widely used and referenced scheme (first column in spatial resolutions (1, 3, and 9 km) and used two alter-
Table 2) and a new one, available in a recent WRF native PBL parameterizations. In addition, for the last
model release (version 3.2; second column in Table 2). domain (1-km spatial resolution) the cumulus parame-
Selection of the old schemes for the various physical terization was disabled.
configurations was based on previous evaluations of Initial and boundary conditions for the simulation
WRF performance in the study region (Fernández et al. were generated using the Interim European Centre for
2007; Borge et al. 2008; Ruiz-Arias et al. 2008). Table 3 Medium-Range Weather Forecasts Re-Analysis (ERA-
shows the parameterization in each experiment. We Interim; Uppala et al. 2008) with 0.7° spatial resolution
note that a computer-code error ("bug") in the Yonsei and 6-h time interval. Newtonian relaxation or nudging
University (YSU) PBL scheme that was used in this is one method of four-dimensional data assimilation
study that affects the simulation of the stable boundary implemented in mesoscale models. This technique aims
layer and its consistency with thermal roughness length to relax the model state toward gridded analyses that are
has been recently reported (see the WRF V3.4.1 updates based on observations (analysis nudging) or directly
at http://www.wrf-model.org). Nevertheless, we think it toward the individual observations (observation nudg-
would not affect seriously the results presented in sec- ing). This method is attempted by adding artificial
tion 4. As explained in section 4, results of the first study forcing terms to the model predictive equations on the
show that the wind estimates were sensitive mainly to basis of the difference between both states weighted
the choice of PBL parameterization. As a consequence, by nudge coefficients (Seaman 2000). Several works in
for the sake of conciseness in the second study, only the literature have shown that the use of nudging in

TABLE 3. Nomenclature of the 32 physical configurations evaluated. Experiments E15 and E31, highlighted in boldface type, were used in
analysis of terrain representation influence.

Nomenclature Physical configuration Nomenclature Physical configuration


El Pl-Ml- Cl-Rl-Ll E17 P2-M1- Cl-Rl-Ll
E2 Pl-Ml- C1-R1-L2 E18 P2-M1- C1-R1-L2
E3 Pl-Ml- C1-R2-L1 E19 P2-M1- C1-R2-L1
E4 Pl-Ml- C1-R2-L2 B20 P2-M1- C1-R2-L2
E5 Pl-Ml C2-R1-L1 E21 P2-M1- C2-R1-L1
E6 Pl-Ml. C2-R1-L2 E22 P2-M1- C2-R1-L2
E7 Pl-Ml- C2-R2-L1 E23 P2-M1- C2-R2-L1
E8 Pl-Ml..C2-R2-L2 E24 P2-M1. C2-R2-L2
B9 P1-M2- Cl-Rl-Ll E25 P2-M2- Cl-Rl-Ll
ElO P1-M2- C1-R1-L2 E26 P2-M2- C1-R1-L2
Ell P1-M2..C1-R2-L1 E27 P2-M2. C1-R2-L1
E12 P1-M2--C1-R2-L2 E28 P2-M2- C1-R2-L2
E13 P1-M2-O-Rl-Ll E29 P2-M2- C2-R1-L1
E14 P1-M2-.C2-R1-L2 E30 P2-M2 C2-R1-L2
E15 P1-M2 .C2-R2-L1 E31 P2-M2 .C2-R2-L1
E16 P1-M2-.C2-R2-L2 E32 P2-M2--C2-R2-L2
1600 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 52

a meteorological model leads to improved meteorolog- Direction data have traditionally been analyzed using
ical simulations (Stauffer and Seaman 1994; Seaman graphical methods in the literature. In particular, the
et al. 1995; Seaman 2000; Otte 2008). The nudging to- comparison of wind direction distributions (wind-rose
ward gridded analysis was used in this study, and nudge plots) has been used for gaining an understanding of di-
coefficients for temperature and water mixing ratio were rectional characteristics. In this work, however, we wan-
selected following Otte (2008), that is, 3 X 10"'*s"' for ted to provide a quantitative measure of the performance
temperature, 1.0 X lO^^s"^ for water mixing ratio, and of the wind direction estimates, as in the wind speed case.
3 X 10"* s~^ for the horizontal wind components. Since A possibility is the deterministic forecast of multiple
the use of analysis nudging in high-resolution domains categories [described in WMO (2000)]. This method
and near the surface may inhibit the simulation of im- mainly consists of a categorical comparison that is based
portant mesoscale features, the analysis nudging was on a contingency table, which is built by ranking the wind
applied only for the first domain (27 km) and only for the direction intervals. From this analysis, several parameters
last 15 vertical levels. The simulations were conducted can be derived, such as the reliability, accuracy, and the
with a spinup of 24 h and outputs were saved every skill score (WMO 2000). In particular, the accuracy is
10 min, and then horizontal wind components were calculated as the sum of the diagonal of the contingency
hourly averaged. matrix, divided by the sum of the whole matrix. This
Following the method of Draxl (2012), the model grid parameter has limitations, however, because it could be
points that are representative for the station locations were that simulated and observed directions were in different
selected by comparing the modeled and actual elevation categories although the circular residual was low.
and distance to the coast. To this end, the model elevation To avoid this problem, in this work we propose a new
and distance to the coast of all of the grid points inside parameter for analyzing the accuracy of the wind di-
a window of 3 X 3 (9) cells centered at each station loca- rection estimates: the directional accuracy (DACC) pa-
tion were compared with the corresponding values derived rameter. This parameter is based on the circular distance,
from the DEM. For each station, the grid point that min- defined as the smaller of the two arc lengths between two
imized the accumulated squared errors was selected. points along a circumference (Jammalamadaka and
Validation of the different model configurations was Sengupta 2001). For angles a and ß , the circular dis-
done on the basis of various statistics for both wind tance is defined as
speed and direction. First, we used the standard de-
viation (STD) of residuals between observed and mod- Ae(a, ß) = mm[a - ß, 360° -(a-ß)].
eled wind speed, which estimates the amount of scatter
of the wind speed errors: Then, we define the DACC to account for the percent-
age of times in which the circular distance between the
1/2 observed and modeled wind direction is lower than
STD = a threshold, chosen as 30° here:
N or
0°<Aö,.<30°)
1/2
DACC= X 100.
relative STD = XIOO, N

Results presented in section 4 correspond to the whole


dataset (16 days). Nevertheless, these results are shown to
where A^ is the number of records and m, and o, are be generally consistent across the four periods (one per
modeled and observed wind speed values, respectively. season) of simulations. It is important to mention that this
Second, we used the mean error or bias, which quantifies work focused on the infiuence of physics parameteriza-
overall overestimation or underestimation of modeled tion and terrain configuration in the WRF model wind-
wind speed estimates:
estimate errors. Nevertheless, other sources of errors—such
as wind measurements errors, errors associated with the
¡ - o¡)
initialization/boundary conditions, land surface parame-
bias = or ter representation errors, or errors associated with an
N
unresolved horizontal thermal gradient—were not con-
sidered in the evaluation procedure. In addition, differing
relative bias = X 100. nesting parameterization schemes could be another
source of discrepancies in model predictions.
JULY 2013 SANTOS-ALAMILLOS ET AL. 1601

FlG. 5. Relative STD (%) of residuals between observed and modeled wind speed for 32 different experiments at
the (a) Centenar, (b) Gomera, (c) Almirez, and (d) Los Nietos stations. The configuration of each experiment can be
detertnined from Tables 2 and 3.

4. Results and discussion lower STD than do experiments E17-E32, which use the
P2 PBL scheme. This is confirmed in Fig. 8. In particular.
a. Analysis of sensitivity to parameterization choice
Fig. 8a shows that use of Pl PBL instead of P2 for
Figures 5-7 show, respectively, relative STD, relative Centenar reduces STD by 6% on average (37% vs43%).
bias, and DACC values obtained for the 32 experiments The other stations show smaller STD differences, of less
at each station. Figure 8 shows the sensitivity of STD, than 3%, associated with the PBL scheme choice (Figs.
bias, and DACC values to the parameterization choice 8b-d). Sensitivity of STD to the other parameterizations
for each station. To this end, the figure shows mean is smaller for all stations. Therefore, aside from the PBL,
values of these parameters resulting from the 16 ex- there is no clear and systematic effect of parameteriza-
periments in which each parameterization was involved. tion selection on STD at any of the four stations. Nev-
The STD values (Fig. 5) show considerable differ- ertheless, there are certain combinations of schemes
ences among stations and are higher for the Almirez and that perform better than others. These combinations
Centenar stations. Almirez shows the highest values, unfortunately vary among stations. For instance, for
between 48% and 55% for all experiments. In contrast, Centenar, STD varies from about 35% (E12: best ex-
Los Nietos has the lowest values, between 33% and 37% periment) to 45% (E27: worst experiment). The best
for most experiments. Dependence of STD on physical experiment for Gomera is E14 (34%) and the worst is
configuration varies considerably among stations. Cen- E28 (40%). The best experiments for Almirez and Los
tenar is the most sensitive by a large amount, followed Nietos are E l (48%) and E15 (33%), respectively.
by Gomera. Recall that these are the lowest-elevation Similar to STD, bias values (Fig. 6) show a notable
stations, in the western part of the study area. A clear dependence on the station and physical configura-
dependence of results on the PBL scheme is observed tion, ranging from - 1 5 % to 10%. There is a clear ten-
for these two stations. In particular, experiments E l - dency to underestimate wind speed at Almirez and Los
E16, which use the Pl PBL scheme, produce considerably Nietos (Figs. 6c,d). The Pl parameterization performs
1602 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 52

•••••••••••••••••••••••••••••••:
FlG. 6. As in Fig. 5, but for relative bias (%).

considerably better than P2 at these stations. At Cen- PBL scheme) show a negative bias, whereas experi-
tenar and Gomera stations (Figs. 6a,b), however, bias ments E17-E32 (using the P2 scheme) produce slightly
associated with the different experiments strongly var- positive or near zero bias. Gomera had considerably
ies. At Centenar, experiments E l - E l 6 (using the PI lower bias associated with the PI scheme. This high

FIG. 7. As in Fig. 4, but for the DACC parameter (%).


JULY 2013 SANTOS-ALAMILLOS ET A L . 1603

FIG. 8. Results of the sensitivity analysis of STD (black bars), bias (gray bars), and wind direction DACC
parameter (white bars) to choice of parameterization. Mean values of these parameters, resulting from 16 experi-
ments in which each parameterization is involved, are displayed for the (a) Centenar, (b) Gomera, (c) Almirez, and
(d) Los Nietos stations.

sensitivity of bias to the PBL parameterization is con- WRF than to the choice of parameterization. This
firmed by Fig. 8. In particular, use of PI for Centenar finding will be further analyzed in the next section.
leads to a - 1 0 % bias on average, whereas P2 leads to To conclude, the PBL parameterization has a signifi-
a value near zero. In contrast, PI performs better than cant influence on both the STD and bias, and the PI
P2 for the rest of the stations, although differences are parameterization scheme performs better overall. This
smaller than for Centenar. As with STD, the other influence is highly dependent on the station and the
schemes had little influence on bias. Again, no single statistics, however. It is notable that for the Centenar
experiment performed best for all stations, but ex- station P1/P2 provides lower/higher STD values and
periment E14 performed reasonably at Almirez and higher/lower bias values.
Los Nietos (high elevations).
b. Analysis of sensitivity to terrain representation
For DACC (Fig. 7), Gomera (between 60% and 70%)
and, especially, Los Nietos (above 75%) have the Results of the previous analysis show that the WRF
highest values. As highlighted earher, these stations wind estimates in the study region are mainly sensitive
have wind directions that are strongly conditioned by to the selection of PBL parameterization. As a conse-
nearby terrain morphology (Figs. 2 and 3). Therefore, it quence, in the analysis described in this section, we only
seems that terrain morphology at these stations is rea- consider variations in this parameterization, keeping
sonably reproduced by WRF at 3-km spatial resolution; the rest of the parameterizations fixed. This approach
this will be confirmed in the next section. Centenar and greatly reduces the complexity of this examination of
Almirez had lower DACC values of close to 60%. The terrain influence. As revealed in the previous section,
DACC parameter shows little sensitivity to the choice of none of the scheme combinations provided a clearly
parameterization. This is confirmed by Fig. 8. Only small better representation of both wind speed and direction
differences (below 2%) associated with the PBL are at the four stations. Nevertheless, among all experi-
observed. Therefore, wind direction estimation appears ments, E15 seemed to provide the best overall estimates.
to be more sensitive to the terrain representation in This experiment produced the lowest, or near the lowest.
1604 J O U R N A L OF A P P L I E D M E T E O R O L O G Y AND C L I M A T O L O G Y VOLUME 52

67.0 1.2

57.0
g
1 0.8
0.4
48.0 a 0
VI c -0.4
38.0
m .0.8
U
29.0 -1.2

47.0 1.2
0.8
40.0 _ 0.4

34.0 D
0

is ai -0.4

Gom
27.0 -0.8
20.0 -1.2

62.0 1.2

s (m/s)
0.8
53.0 0.4
m
44.0 a g N
0
£! -0.4
36.0 t;

Aimi
-0.8
27.0 -1.2

59.0 1.2
0.8
51.0
_ 1s 0.4
STD (°/i

42.0 0
S
let

.0.4
34.0 z -0.8

P1.3km Pl-lkm P2-9km P2.3km P2-lkm


25.0 a -1.2
FIG. 9. Results of the sensitivity analysis of STD to spatial resolution FIG. 10. As in Fig. 9, but for bias (ms~^).
of WRF setup: STD values of WRF wind speed estimates using the
E15 (Pl) and E31 (P2) configurations for three different spatial res- improvement (about 0.2 m s^^) of the STD was obtained
olutions (9,3, and 1 km) at the (a) Centenar, (b) Gomera, (c) Almirez, by increasing spatial resolution. At Los Nietos and
and (d) Los Nietos stations. The left-hand scale represents the STD
Almirez, STD values were slightly lower at 9 km than at
in meters per second, and the right-hand scale represents the STD in
percent with respect to the mean value of the observations. the 3- and 1-km spatial resolutions (about 0.2ms~^).
The reasons for this last result are difficult to assess and
may be associated with error sources that are not related
values of STD and bias for all stations (Figs. 5 and 6). The to the terrain representation, such as those listed at the
DACC score (Fig. 7) showed httle variation among ex- end of section 3.
periments. Therefore, we have selected E15 as a basis for A general tendency to underestimate wind speed,
the analysis presented here. We performed six experi- regardless of PBL parameterization and resolution, can
ments, involving the two physical configurations (E15 and be discerned from Fig. 10. Except for Centenar station,
E31) and the three spatial resolutions (1, 3, and 9 km), to the use of the Pl experiment provides lower bias values.
evaluate the effect of terrain representation on the WRF Figure 10 also reveals a significant and complex de-
wind estimates. This implies maintaining the parameteri- pendence of the bias on the spatial resolution. At Los
zation M2, C2, R2, and LI in both experiments, changing Nietos, the bias was positive at 9 km and was negative
only the PBL scheme. We refer to the experiment that uses at 3 and 1km (Fig. lOd). As shown in Table 1, WRF
the E15 physical configuration as Pl and to the experiment overestimated station elevation at 9 km (1262 m MSL)
that uses the E31 configuration as P2 (Table 3). while underestimating it at 3 km (1013 m MSL) and 1 km
Figures 9-11 show the STD, bias, and DACC values, (1055 m MSL)—the DEM shows 1056 m MSL. The signs
respectively, for the two physical configurations and of the biases at Los Nietos may be caused by this over-
three different spatial resolutions. The STDs (Fig. 9) estimation/underestimation of station elevation. Since
show considerable dependence on the PBL, station, and this station is in a fiat area with low topographic com-
spatial resolution. Overall, at the four stations and for plexity, according to Jiménez and Dudhia (2012), a
the three spatial resolutions, the Pl scheme generated positive bias (about lms~^ at 10 m AGL) is expected
lower STDs, with greater differences at Centenar (Fig. 9a). because of an unresolved drag effect of the terrain. Our
Dependence on spatial resolution was much more com- results are compatible with the existence of a small
plex, however. The Centenar station showed no de- positive bias of substantially lower than 1 m s~^ This is
pendence. At Gomera with the Pl scheme (Fig. 9b), slight mainly because the measuring elevation considered in
JULY 2013 S A N T O S - A L A M I L L O S ET A L . 1605

station. This underestimation is clearly observed at


9 km, for which the station elevation from WRF is close
to actual. At 1 and 3 km, the negative bias is marginally
less. This reduction is probably caused by a compensat-
ing effect in the wind speed estimates that is associated
with the overestimation of station elevation (about
17 m) at these spatial resolutions.
Several conclusions can be drawn from Fig. 11 re-
garding DACC. To aid interpretation of the results.
Fig. 12 presents wind direction distributions from WRF,
using PI at the three spatial resolutions. The observed
distribution is also shown. The DACC estimates show
a small but significant dependence on the PBL param-
eterization. The PI scheme provides DACC values at
Centenar at 1-km spatial resolution that are about 3%
higher (Fig. 11a). On the other hand, P2 provides better
DACC values at Gomera; the differences reach 8% at
1-km spatial resolution (Fig. lib). For the other stations,
differences are smaller. There are notable differences
among stations in the accuracy of wind direction esti-
mates. The highest DACC values were observed for Los
Nietos, and the lowest were observed for Centenar. The
spatial resolution had a relevant infiuence on DACC,
especially for Los Nietos station (Fig. lid). As high-
lighted in section 2, this station presents low wind di-
Pl-3km P2-9km P2-3km rection variability, with this direction highly conditioned
FIG. 11. As in Fig. 9, but for wind direction DACC parameter (%). by nearby terrain morphology. The location of Los
Nietos station, in the middle of a valley (Fig. 2d and Figs.
3p,t) clearly modulates the wind direction distribution.
this evaluation is 40 m AGL, at which height the drag
The valley morphology at this station seems reasonably
effect should be much lower than it is at 10 m AGL.
represented at 3 and 1 km but not at 9-km resolution
Results for Gomera (Fig. 10b) clearly show the effect
(Figs. 3p-s). As a result, wind direction distribution was
of elevation misrepresentation of mountains and hills on
reasonably reproduced at 3 and 1km (Figs. 12m,o,p),
the wind estimates. This is a hilltop station, and WRF
and the DACC parameter is considerable larger at these
estimates of station elevation (Table 1) range from
two spatial resolutions (Fig. lid). Gomera (Fig. 2b) and
405 m MSL (9-km resolution) to 515 m MSL (1-km
Almirez (Fig. 2c) present a low dependence of the
resolution)—the DEM gives 640 m MSL. As a conse-
DACC on the spatial resolution and also show low wind
quence, the station has a negative bias in both the PI and
direction variability (Fig. 12). Unlike the case of Los
P2 experiments that decreases in magnitude with in-
Nietos, however, the three spatial resolutions provide
creasing spatial resolution. At Almirez (Fig. 10c) in the
similar results in terms of wind direction distribution
foothill area, the elevation misrepresentation effect is
(Figs. 12f-h, j-1). As highlighted in section 2, these two
also evident. At 9-km resolution, station elevation de-
stations present the highest topographic complexity
rived from the WRF was 1052 m MSL, which is consid-
among the four analyzed. It seems that the terrain fea-
erably lower than the reference DEM elevation of
tures that constrain the wind direction variability at
1195 m MSL. This results in the negative bias that is seen
these two stations are so complex that they are not
in Fig. 10c (about -0.4m s^^). Using the 3- and 1-km
properly reproduced in the WRF, even using the 1-km
spatial resolutions, however, the WRF station elevations
spatial resolution. Therefore, low DACC values, and
are close to the DEM value, and only a slight positive
a weak dependence of this parameter on the spatial
bias (about 0.2 m s~^) is observed. Centenar station, lo-
resolution, are found for both stations. Centenar had the
cated above a smooth hill, presents low topographic
worst DACC values (Fig. 11a), with little dependence
complexity (Fig. 2a and Fig. 3a). Elevations from WRF
on WRF spatial resolution. At this station with low to-
vary from 229 m at 9 km to 253 m MSL at 1 and 3 km; the
pographic complexity, terrain morphology had little
DEM indicates 236 m MSL. Results in Fig. 10a indicate
effect on the wind direction distribution (Figs. 3a,e).
a tendency for WRF to underestimate wind speed at this
1606 J O U R N A L OF A P P L I E D M E T E O R O L O G Y AND C L I M A T O L O G Y VOLUME 52

Centenar
I [0.0 : 2.0]
I [2.0:4.0]
I [4.0 : 6.0]
I [6.0 : 8.0]
I [8.0 : 10.0]
I [10.0 : 12.0]
I [12.0 : inf]

Gomera
[0.0 : 2.5]
[2.5 : 5.0]
[5.0:7.5]
[7.5 : 10.0]
[10.0: 12.5J
[12.5 : 15.0]
[15.0 : inf]

Almirez
^ [0.0: 2.0]
^ [2.0: 4.0]
^m [4.0 : 6.0]
• • [6.0 :8.0]
[ZZl [8.0 : 10.0]
^m [10.0 : 12.0]
^m [12.0 : inf]

Los Nietos
I [0.0 : 2.0]
I [2.0 : 4.0]
I [4.0 : 6.0]
I [6.0 : 8.0]
] [8.0 : 10.0]
I [10.0 : 12.0]
I [12.0 : inf]

FIG. 12. Modeled [from the best-performing WRF configuration (E15) and using different spatial resolutions] and observed wind
direction distributions, at the 40-m AGL measurement elevation of four stations. Shown are the (a) observed and (b)-(d) modeled (using
9-, 3-, and 1-km spatial resolutions, respectively) wind distributions for the Centenar station. Also shown are plots as in (a)-(d), but for the
(e)-(h) Gomera, (i)-(l) Almirez, and (m)-(p) Los Nietos stations. The values in the circle indicate the percentage of observations; the
color scale at right indicates ranges of wind speed (m s~').

Therefore, differences in terrain representation at the channeling effect of surrounding topography (Figs. 2d
three WRF resolutions (Figs. 3b-d) had little effect on and 3p,t). The wind distribution was also reasonably re-
the wind distribution estimates (Figs. 12b-d). produced at Gomera station (Figs. 12e,h), but with a bias
To sum up, from Figs. 9-12, it is concluded that the of -25° toward the southeast. At Almirez (Fig. 121), the
WRF setup that is based on experiment E15 (Pl) and principal wind directions were fairly well reproduced,
1-km spatial resolution provides the best wind speed and with no bias; not so the distribution. The model indicated
direction estimates overall at the four stations. Never- more cases than were observed in the south through
theless, there are significant differences among stations in east directions and fewer cases than were observed in
the reliability of these estimates. The STD ranged from the north through west directions. The poorest esti-
about 2 (Centenar) to 2.8 (Gomera) m s ~ \ The bias was mation of the wind direction distribution is for Centenar
from -0.4 (Centenar) to 0.3 (Almirez) m s~\ and DACC (Figs. 12a,d). The model was able to estimate some of the
ranged from about 57% at Centenar to 85% at Los Nietos. predominant wind directions, but some of the modeled
Wind direction distribution was fairly reproduced at directions are not found in the observed data.
Los Nietos (Figs. 12m-p). The predominant southeast
wind direction was clearly reproduced by the model,
5. Summary and conclusions
although there is overestimation of values for the north
through west directions. As highlighted earlier, this In this work, we evaluated the relative roles of pa-
predominant wind direction can be explained by the rameterization choice and terrain representation in the
JULY 2013 S A N T O S - A L A M I L L O S ET A L . 1607

Weather Research and Forecasting mesoscale model, as version (RRTMG) (SWR) parameterization schemes,
they pertain to regional wind resource assessment. In performed the best. Since results of the flrst analysis
a flrst part, to evaluate sensitivity of WRF wind esti- showed that WRF wind estimates in the study region
mates to parameterization choice, 32 conflgurations were mainly sensitive to the PBL parameterization se-
(experiments) of the microphysies, cumulus, planetary lection, only variations of this parameterization were
boundary layer, or shortwave and longwave radiation considered in the second analysis. Six experiments were
parameterization schemes were assessed. These experi- conducted, with three spatial resolutions (1,3, and 9 km)
ments, with 3-km spatial resolution and 1-h output fre- and two PBL parameterizations. The remaining pa-
quency, were conducted for the four seasons in 2005. In rameterizations were chosen on the basis of the best-
a second part, to evaluate the sensitivity of the wind es- performing experiment from the flrst analysis. Although
timates to terrain representation in WRF, experiments the spatial conflguration and, in particular, the nesting
were conducted with three different spatial resolutions— approaches in the flrst and second part of this study were
9,3, and 1 km. The study region was Andalusia (southern different, we consider that results from the flrst part are
Spain). Although relatively small, this region is charac- still applicable for the analysis of the second part, mainly
terized by varied topographic, geographic, and weather because we believe that the weak influence that the
conditions that made the evaluation especially chal- choice of the MPH, CMS, SWR, and LWR parameter-
lenging. Estimates were obtained with data collected at izations schemes showed on the wind estimates in the
four stations, the locations of which are representative flrst analysis should not change signiflcantly just by
of different climatic conditions and varied topographic changing the nesting approach.
complexity and morphology in the study region. The In the second analysis, STDs showed a dependence on
four stations provided hourly wind speed and direction the PBL parameterization, with the PI (YSU) scheme
data at 40 m AGL. The WRF estimates were assessed performing better (differences up to 0.3 ms~^). On the
using different statistics that represented both wind other hand, the influence of the spatial resolution on
speed and wind direction. The standard deviation of the STD values was found to be considerably weak
residuals between observed and modeled wind speed and highly dependent on the evaluation station. Biases
and the bias were used to evaluate wind speed pre- in the WRF wind speed estimates were very sensitive
dictions. For evaluating wind direction, we used an ac- to model spatial resolution. This was mainly because higher
curacy parameter (DACC) that considers the percentage resolution improved the representation of terrain eleva-
of times for which the circular distance between observed tion. The sign of the bias depended on terrain morphology
and modeled wind direction is less than 30°. and the spatial resolution, but absolute values tended to be
Results of the flrst analysis show STD values of the much higher with coarser spatial resolution (9 km).
wind speed estimates to be sensitive mostly to the PBL Physical conflguration was found to have little im-
parameterization choice. For some of the stations, dif- pact on wind direction distribution estimates. These
ferences up to 10% were observed, depending on the estimates proved to be more sensitive to the repre-
PBL parameterization. Sensitivity of the wind speed sentation of terrain morphology around the station
STD values to other parameterizations was not signif- than to the model spatial resolution. In this regard, for
icant. Nevertheless, some combinations of schemes stations that are located in areas of high topographic
(experiments) performed better than others. These complexity, the use of 3- or even 9-km grid resolution in
combinations unfortunately differed among stations. the model may provide similar (low) accuracy in the
The bias values of wind speed showed considerable de- wind direction estimates to that from 1-km grid reso-
pendence on the station and choice of PBL parameter- lution. In other cases, when the terrain is not so com-
ization. Values ranged from - 1 5 % to 10%, depending plex, the increment in the spatial resolution provide
on the station and experiment. The DACC values a substantial improvement in the accuracy of the wind
showed a large sensitivity only to the choice of the PBL direction estimates.
parameterization scheme, although only small differ- Experiment E15 at 1-km grid resolution provided the
ences (below 2%) were observed. On the other hand, overall best estimates of wind speed and direction at
there were substantial differences among stations in the four stations. Nevertheless, at the stations where the
DACC values, indicating that WRF wind direction es- terrain morphology and elevations are fairly reproduced
timates are highly sensitive to terrain characteristics. at 3-km grid resolution (Almirez and Los Nietos), re-
Among all 32 experiments, experiment number 15, sults for this resolution are similar to the results obtained
which used the YSU (PBL), Morrison two-moment for the 1-km grid resolution. In addition, there were
(Morrison-2M) (MPH), Grell-3D (CMS), Rapid Radia- signiflcant differences among stations in the reliability
tive Transfer Model (RRTM) (LWR), and RRTM GCM of these estimates. The STD ranged from about 2 to
1608 J O U R N A L OF A P P L I E D M E T E O R O L O G Y AND CLIMATOLOGY VOLUME 52

more than 2.8 m s . The bias was from - 0.4 to 0.3 m s Ris0-PhD-84(EN), 104 pp. [Available online at http://orbit.
and the DACC ranged from approximately 57% to dtu.dk/files/10246050/Ris_PhD_84.pdf.]
85%. Wind direction distribution was reliably repro- Esteban-Parra, M. J., F. Rodrigo, and Y. Castro-Diez, 1998: Spatial
and temporal patterns of precipitation in Spain for the period
duced with this WRF setup for the stations that present 1880-1992. Int. J. Climatol., 18,1557-1574.
marked terrain morphology features (viz., the Los Farr, T. G., and Coauthors, 2007: The shuttle radar topography mis-
Nietos, Almirez, and Gomera stations). Nevertheless, sion. Rev. Geophys., 45, RG2004, doi:10.1029/2005RG000183.
at some stations the model showed bias in the pre- Fernández, J., J. P. Montávez, J. Saenz, J. F. González-Rouco,
dominant directions, and at other stations it output wind and E. Zorita, 2007: Sensitivity of the MM5 mesoscale model
to physical parameterizations for regional climate studies:
directions that were not observed. Annual cycle. J. Geophys. Res., 112, D04101, doi:10.1029/
2005JD006649.
Acknowledgments. The Consejería de Innovación, Frank, H. P., O. Rathmann, N. G. Mortensen, and L. Landberg,
Ciencia y Empresa (CICE) of Junta de Andalucía (Spain) 2001: The numerical wind atlas—The KAMM/WAsP method.
(Project P07-RNM-02872) and FEDER funds financed Ris0 National Laboratory Tech. Rep. Ris0-R-1252(EN), 60 pp.
[Available online at http://www.risoe.dk/rispubl/VEA/veapdf/
this study. This work has been also supported by the ris-r-1252.pdf.]
TEP-220 research group. The authors thank Gamesa GE Energy, 2010: Western wind and solar integration study. Na-
Co. for providing the wind measurements. We also tional Renewable Energy Laboratory Tech. Rep. NREL/SR-
thank the reviewers for their very helpful comments. 550-47434, 536 pp. [Available online at http://www.nrel.gov/
docs/fyl0osti/47434.pdf.]
Gibbs, J. A., E. Fedorovich, and A. Eijk, 2011: Evaluating Weather
REFERENCES Research and Forecasting (WRF) model predictions of tur-
Alsamamra, H., 2009: Using geostatistical techniques to obtain bulent flow parameters in a dry convective boundary layer.
high spatial resolution climate maps for Andalusia (southern J. Appl. Meteor. Climatol., 50,2429-2444.
Spain). Ph.D. thesis. University of Jaén, 262 pp. [Available Hahmann, A., D. Rostkier-Edelstein, F. Vandenberghe, Y. Liu,
online at http://refbase.iecolab.es/files/husainalsamamra/2009/ S. Swerdlin, T. Warner, and R. Babarsky, 2010: A reanalysis
1147_HusainAlsamamra2009.pdf.] system for the generation of mesoscale climatographies.
Arakawa, A., 2004: The cumulus parameterization problem: Past, /. Appl. Meteor. Climatol., 49, 954-972.
present, and future. J. Climate, 17, 2493-2525. Hong, S., Y. Noh, and J. Dudhia, 2006: A new vertical diffusion
Arya, S., 2001: Introduction to Micrometeorology. Academic Press, package with an explicit treatment of entrainment processes.
420 pp. Mon. Wea. Rev., 134,2318-2341.
Beljaars, A. C. M., A. R. Brown, and N. Wood, 2004: A new pa- Horvath, D., D. Koracin, R. Vellore, and J. Jiang, 2012: Sub-
rametrization of turbulent orographie form drag. Quart. kilometer dynamical downscaling of near-surface winds in
J. Roy. Meteor. Soc, 130,1327-1347. complex terrain using WRF and MM5 mesoscale models.
Berge, E., R. E. Bredesen, and K. MoUestad, 2007: Combining /. Geophys. Res., Ill, D l l l l l , doi:10.1029/2012JD017432.
WAsP with WRF mesoscale model: Evaluation of wind re- Hu, X. M., J. W. Nielsen-Gammon, and F. Zhang, 2010: Evaluation
source assessment for three Norwegian wind farm areas. Proc of three planetary boundary layer schemes in the WRF Model.
EWEC2007, Milan, Italy, European Wind Conference and /. Appl. Meteor. Climatol, 49,1831-1844.
Exhibition. Huaxing, L. U., 2008: Modelling terrain complexity. Advances in
Borge, R., V. Alexandrov, J. Del Vas, J. Lumbreras, and Digital Terrain Analysis, Springer, 159-176.
E. Rodriguez, 2008: A comprehensive sensitivity analysis of Jammalamadaka, S., and A. Sengupta, 2001: Topics in Circular
the WRF model for air quality applications over the Iberian Statistics. World Scientific, 322 pp.
Peninsula. Atmos. Environ., 42, 8560-8574. Jiménez, B., F. Durante, B. Lange, T. Kreutzer, and J. Tambke,
Byrkjedal, 0., and E. Berge, 2008: The use of WRF for wind re- 2007: Offshore wind resource assessment with WAsP and
source mapping in Norway. Preprints, Ninth Annual WRF MM5: Comparative study for the German Bight. Wind En-
User's Workshop, Boulder, CO, NCAR, P9.18., 4 pp. [Avail- ergy, 10,121-134.
able online at http://viww.mmm.ucar.edu/wr&users/workshops/ Jiménez, P. A., and J. Dudhia, 2012: Improving the representation
WS2008/abstracts/P9-18.pdf.] of resolved and unresolved topographic effects on surface wind
Castro-Diez, Y., D. Pozo-Vázquez, F. Rodrigo, and M. J. Esteban- in the WRF model /. Appl. Meteor. Climatol., 51, 300-316.
Parra, 2002: NAO and vnnter temperature variability in southern , J. F. González-Rouco, J. Navarro, J. P. Montávez, and
Europe. Geophys. Res. Lett., 29, doi:10.1029/2001GL014042. E. Garcia-Bustamante, 2010: Ouality assurance of surface
Chou, M.-D., and M. J. Suarez, 1994: Technical report series on wind observations from automated weather stations. J. Atmos.
global modeling and data assimilation. Volume 3: An efficient Oceanic Technoi, 27,1101-1122.
thermal infrared radiation parameterization for use in general Kain, J., and J. Fritsch, 1993: Convective parameterization for
circulation models. NASA Tech. Memo. 104606, Vol. 3,85 pp. mesoscale models: The Kain-Fritsch scheme. The Represen-
[Available online at http://archive.org/details/nasa_techdoc_ tation of Cumulus Convection in Numerical Models, Meteor.
19950009331.] Monogr., No. 46, Amer. Meteor. Soc, 165-170.
Cohen, C, 2002: A comparison of cumulus parameterizations in Kanamitsu, M., and H. Kanamaru, 2007: Fifty-seven-year California
idealized sea-breeze simulations. Mon. Wea. Rev., 130,2554- reanalysis downscaling at 10 km (CaRDlO). Part I: System detail
2571. and validation with observations. / Climate, 20,5553-5571.
Draxl, C, 2012: On the predictability of hub height winds. Tech- Kim, Y., J. H. Jeong, J. H. Bae, S. K. Song, and J. W. Seo, 2005: Study
nical University of Denmark (DTU) Wind Energy PhD-Rep. on planetary boundary layer schemes suitable for simulation
JULY 2013 SANTOS-ALAMILLOS ET AL. 1609

of sea surface wind in the southeastern coastal area. /. Environ. Ramos-Calzado, P., J. Gómez-Camacho, P. Pérez-Bernal, and
Sei., 14,1015-1026. M. Pita-Lopez, 2008: A novel approach to precipitation series
Kotroni, V., and K. Lagouvardos, 2001; Precipitation forecast skill completion in climatological datasets: Application to Anda-
of different convective parameterization and microphysical lusia. Int. J. ClimatoL, 28,1525-1534.
schemes: Application for the cold season over Greece. Geo- Rodriguez-Puebla, C, A. H. Encinas, S. Nieto, and J. Garmendia,
phys. Res. Lett, 28,1977-1980. 1998: Spatial and temporal patterns of annual precipitation
Kwun, J., Y. Kim, J. Seo, J. Jeong, and S. You, 2009: Sensitivity of variability over the Iberian Peninsula. Int. J. ClimatoL, 18,
MM5 and WRF mesoscale model predictions of surface winds 299-316.
in a typhoon to planetary boundary layer parameterizations. Ruiz-Arias, J., D. Pozo-Vázquez, N. Sánchez-Sánchez, J. Montávez,
Nat. Hazards, 51, 63-77. A. Hayas-Barrú, and J. Tovar-Pescador, 2008: Evaluation of
Landberg, L., L. Myllerup, O. Rathmann, E. Petersen, B. J0rgensen, two MM5-PBL parametrizations for solar radiation and tem-
J. Badger, and N. Mortensen, 2003: Wind resource estimation perature estimaUon in the south-eastern area of the Iberian
an overview. Wind Energy, 6,261-271. Peninsula. Nuovo Cimento C, 31, 825-842.
Lavagnini, A., A. Sempreviva, C. Transedci, C. Accadia, M. Casaioli, Seaman, N. L., 2000: Meteorological modeling for air-quality as-
S. Mariani, and A. Speranza, 2006: Offshore wind climatology sessments. Atmos. Environ., 34, 2231-2259.
over the Mediterranean basin. Wind Energy, 9, 251-266. , D. R. Stauffer, and A. M. Lario, 1995: A multiscale four-
Loveland, T. R., J. F. Reed, D. O. Brown, J. Ohlen, Z. Zhu, dimensional data assimilation system applied in the San
L. Yang, and J. W. Merchant, 2000: Development of a global Joaquin Valley during SARMAP. Part I: Modeling design
land-cover characteristics database and IGBP DISCover from and basic performance characteristics. /. Appl. Meteor., 34,
1-km AVHRR data. Int. J. Remote Sens., 21,1303-1330. 1739-1761.
Lynn, B. H., and Coauthors, 2004: Sensitivity of present and future Sempreviva, A., R. Barthelmie, and S. Pryor, 2008: Review of
surface temperatures to precipitation characteristics. Climate methodologies for offshore wind resource assessment in
Res., 28, 53-65. European seas. Surv. Geophys., 29,471^97.
Miao, J.-F., and Coauthors, 2008: Evaluation of MM5 mesoscale Shimada, S., and T. Ohsawa, 2011: Accuracy and characteristics
model at local scale for air quaUty applications over the of offshore wind speeds simulated by WRF. SOLA, 1, 21-
Swedish west coast: Influence of PBL and LSM parameteri- 24.
zations. Meteor. Atmos. Phys., 99, 77-103. Skamarock, W., and Coauthors, 2008: A description of the Ad-
Mlawer, E., S. Taubman, P. Brown, M. Iacono, and S. Clough, 1997: vanced Research WRF version 3. NCAR Tech. Note NCAR/
Radiative transfer for inhomogeneous atmospheres: RRTM, TN-4751-I-STR, 125 pp.
a validated correlated-fc model for the longwave. /. Geophys. Stauffer, D. R., and N. L. Seaman, 1994: Multiscale four-
Äe.s., 102,16663-16682. dimensional data assimilation. J. AppL Meteor., 33,416-434.
Morrison, H., and A. Gettelman, 2008: A new two-moment bulk Stensrud, D., 2007: Parameterization Schemes: Keys to Under-
stratiform cloud microphysics scheme in the Community At- standing Numerical Weather Prediction Models. Cambridge
mosphere Model, version 3 (CAM3). Part I: Description and University Press, 459 pp.
numerical tests. /. Climate, 21, 3642-3659. Trigo, R., T. Osborn, and J. Corte-Real, 2002: The North Atlantic
Mortensen, N. G., L. Landberg, I. Troen, and E. L. Petersen, 1998: oscillation influence on Europe: Climate impacts and associ-
Wind Atlas Analysis and Application Program (WAsP), ated physical mechanisms. Climate Res., 20, 9-17.
Vol. 1: Getting started. Riso National Laboratory Tech. Rep. Uppala, S., D. Dee, S. Kobayashi, P. Berrisford, and A. Simmons,
Ris0-I-666(EN), 30 pp. 2008: Towards a climate data assimilation system: Status up-
Nakanishi, M., and H. Niino, 2004: An improved Mellor-Yamada date of ERA-Interim. ECMWF Newsletter, No. 155, ECMWF,
level-3 model with condensation physics: Its design and veri- Reading, United Kingdom, 12-18.
fication. Bound.-Layer Meteor., Ill, 1-31. Watson, L., W. Bauman, and B. Hoeth, 2009: Weather Research
Otte, T. L., 2008: The impact of nudging in the meteorological and Forecasting Model Wind Sensitivity Study at Edwards Air
model for retrospective air quality simulations. Part II: Eval- Force Base, CA. Preprints, 13th Conf. on Integrated Observing
uating collocated meteorological and air quaUty observations. and Assimilation Systems for Atmosphere, Oceans, and Land
J. Appl. Meteor. ClimatoL, 47,1868-1887. Surface (lOAS-AOLS), Phoenix, AZ, Amer. Meteor. Soc,
Pérez-Landa, G., P. Ciáis, M. Sanz, B. GioU, F. Miglietta, J. L. 6.1. [Available online at https://ams.confex.com/ams/89annual/
Palau, G. GangoiU, and M. M. Millán, 2007: Mesoscale cir- techprogram/paper_l 44873.htm.]
culations over complex terrain in the Valencia coastal region, WMO, 2000: Guidelines on performance assessment of public
Spain—Part 1: Simulation of diurnal circulation regimes. At- weather services. World Meteorological Organization Tech.
mos. Chem. Phys., 7,1835-1849. Rep. WMO/TD 1023, 67 pp. [Available online at http://www.
Pozo-Vázquez, D., M. Esteban-Parra, F. Rodrigo, and Y. Castro- wmo.int/pages/prog/amp/pwsp/pdf/TD-1023.pdf.]
Diez, 2001: A study of NAO variability and its possible non- Zagar, N., M. Zagar, M. Cedilnik, C. Gregori, and J. Rakovec,
linear influences on European surface temperature. Climate 2006: Validation of mesoscale low-level winds obtained by
Dyn., 17,701-715. dynamical downscaling of ERA40 over complex terrain. Tel-
Rajeevan, M., A. Kesarkar, S. Thampi, T. Rao, B. Radhakrishna, lus, 58A, 445^55.
and M. Rajasekhar, 2010: Sensitivity of WRF cloud microphysics Zhang, D., and W. Zheng, 2004: Diurnal cycles of surface winds
to simulations of a severe thunderstorm event over Southeast and temperatures as simulated by five boundary layer pa-
India. Ann. Geophys., 28, 603-619. rameterizations. /. AppL Meteor., 43,157-169.
Copyright of Journal of Applied Meteorology & Climatology is the property of American
Meteorological Society and its content may not be copied or emailed to multiple sites or
posted to a listserv without the copyright holder's express written permission. However, users
may print, download, or email articles for individual use.

Вам также может понравиться